content
stringlengths 1
15.9M
|
---|
\section{Introduction}
It is well known that classical Newtonian dynamics fails on galactic
scales. There is astronomical and cosmological evidence for a
discrepancy between the dynamically measured mass-to-light ratio of
any system and the minimum mass-to-light ratios that are compatible
with our understanding of stars, of galaxies, of groups and clusters
of galaxies, and of superclusters. Observations on the rotation curves have turn out
that galaxies are not rotating in the same manner as the Solar
System. If the orbits of the stars are governed solely by
gravitational force, it was expected that stars at the outer edge of
the disc would have a much lower orbital velocity than those near
the middle. In fact, by the Virial theorem the total kinetic energy
should be half the total gravitational binding energy of the
galaxies. Experimentally, however, the total kinetic energy is found
to be much greater than predicted by the Virial theorem. Galactic
rotation curves, which illustrate the velocity of rotation versus
the distance from the galactic center, cannot be explained by only
the visible matter. This suggests that either a large portion of the
mass of galaxies was contained in the relatively dark galactic halo
or Newtonian dynamics does not apply universally.
The dark matter proposal is mostly referred to Zwicky \cite{Zwicky}
who gave the first empirical evidence for the existence of the
unknown type of matter that takes part in the galactic scale only by
its gravitational action. He found that the motion of the galaxies
of the clusters induced by the gravitational field of the cluster
can only be explained by the assumption of dark matter in addition
to the matter of the sum of the observed galaxies. Later, It was
demonstrated that dark matter is not only an exotic property of
clusters but can also be found in single galaxies to explain their
flat rotation curves.
The second proposal results in the modified Newtonian dynamics
(MOND), proposed by Milgrom, based on a modification of Newton's
second law of motion \cite{MOND}. This well known law states that an
object of mass $m$ subject to a force $F$ undergoes an acceleration
$a$ by the simple equation $F=ma$. However, it has never been
verified for extremely small accelerations which are happening at
the scale of galaxies. The modification proposed by Milgrom was the
following
\begin{equation}
F=m\mu(\frac{a}{a_0})a,
\end{equation}
\begin{equation}
\mu(x)=\left \{ \begin{array}{ll} 1 \:\: \mbox{if}\:\: x\gg 1
\\
x \:\: \mbox{if}\:\: \|x|\ll 1,
\end{array}\right.
\end{equation}
where $a_0=1.2\times10^{-10} ms^{-2}$ is a proposed new constant.
The acceleration $a$ is usually much greater than $a_0$ for all
physical effects in everyday life, therefore $\mu(a/a_0)$=1 and
$F=ma$ as usual. However, at the galactic scale outside the central bulge
where $a \sim a_0$ we have the modified dynamics $F=m(\frac{a^2}{a_0})$. Using this new law of dynamics for the gravitational force $F=\frac{GmM}{r^2}$
we obtain
\begin{equation}
\frac{GM}{r^2}=(\frac{a^2}{a_0}),
\end{equation}
which results in the constant rotational velocity
\begin{equation}\label{v}
v^2=\sqrt{GMa_0}.
\end{equation}
Another interesting model in this direction has been recently
proposed by Sanders. In this model, it is assumed that gravitational
attraction force becomes more like $1/r$ beyond some galactic
scale\cite{Sanders}. A test particle at a distance $r$ from a large
mass $M$ is subject to the acceleration
\begin{equation}
a = \frac{GM}{r^2} g(r/r_0),
\end{equation}
where $G$ is the Newtonian constant, $r_0$ is of the order of the
sizes of galaxies and $g(r/r_0)$ is a function with the asymptotic
behavior
\begin{equation}
g(r/r_0)=\left \{ \begin{array}{ll} 1 \:\: \mbox{if}\:\: r\gg r_0
\\
r/r_0 \:\: \mbox{if}\:\: r\ll r_0.
\end{array}\right.
\end{equation}
Dark matter as the manifestation of Mach principle has also been
considered as one of the solutions for the dark matter problem.
According to Mach principle \cite{Mach} the distant mass
distribution of the universe has been considered as being
responsible for generating the local inertial properties of the
close material bodies. Borzeszkowski and Treder have shown that the
dark matter problem may be solved by a theory of Einstein-Mayer type
\cite{Treder}. The field equations of this gravitational theory
contain hidden matter terms, where the existence of hidden matter is
inferred solely from its gravitational effects. In the
nonrelativistic mechanical approximation, the field equations
provide an inertia-free mechanics where the inertial mass of a body
is induced by the gravitational action of the cosmic masses. From
the Newtonian point of view, this mechanics shows that the effective
gravitational mass of astrophysical objects depends on $r$ such that
one expects the existence of new type of matter, the so called dark
matter. In what follows we will introduce a new model for the dynamical structure of a galaxy, based on a model for the structure of the nucleus, as an alternative to dark matter and MOND to explain the rotation curve.
\section{Liquid-Droplet model for the nucleus}
In nuclear physics, Weizs\"acker's formula or the semi-empirical mass formula (SEMF) is used to approximate the mass and various features of atomic nucleus
\cite{Weizs}. The theoretical basis of this formula is the liquid drop model. This model in nuclear physics, first proposed by George Gamow and developed by Niels Bohr and John Archibald Wheeler, treats the nucleus as a drop of incompressible nuclear fluid which is made of nucleons (protons and neutrons) binding together by the strong nuclear force. It predicts the binding energy of a nucleus in terms of the numbers of protons and neutrons it contains. This statement
manifests in the semi-empirical mass formula which has five terms on its right hand side. These correspond to i) volume binding energy of all the nucleons inside the
nucleus, ii) surface binding energy of all the nucleons on the surface of
nucleus, iii) the electrostatic mutual repulsion of the protons, iv) an asymmetry term and v) a pairing term (both derivable from quantum mechanical considerations).
The first two terms concern about the {\it attractive} strong nuclear force.
We shall focus on these two terms and leave the other terms irrelevant for
our purposes. Assuming an approximately constant density for the nucleus one can calculate the nuclear radius by using that density as if the nucleus were a drop of a uniform liquid. The liquid droplet model of the nucleus takes into account the fact that the forces on the nucleons on the surface is different from those associated with the interior ones. This is something similar to taking into account of surface tension as a contributor to the energy of a tiny liquid drop.
In the case of volume binding energy, when an assembly of nucleons of the same size is packed together into the smallest volume, each interior nucleon has a certain number of other nucleons in close contact with it. Actually, the number of pairs that can be taken from $A$ particles is $A(A-1)/2$, so one may expect a term for binding energy proportional to $A^2$ interacting pairs. However, the strong force has a very limited range, and a given nucleon may only interact strongly with its nearest neighbors. In other words, the nuclear force is saturated and each nucleon contributes an almost constant energy to the binding of the nucleus. Therefore, the number of pairs of particles that effectively interact is roughly proportional to $A$ and not $A^2$, and the nuclear binding energy is proportional to the volume of the nucleus. The volume term suggests that each nucleon interacts with a constant number of surrounding
nucleons, independent of $A$.
In the case of surface binding energy however, each nucleon at the surface of a nucleus interacts with fewer other (interior and surface) nucleons than one in the interior of the nucleus. Therefore, the binding energy associated with each surface nucleon is less than that of the interior one. In fact, the surface term makes correction to the volume term by its negative contribution as follows
\begin{equation}\label{1}
B=a_V A-a_S A^{2/3},
\end{equation}
where $B$ accounts for the binding energy and $a_V, a_S$ are some
appropriate constants with the same order of magnitude. One may focus more
on the volume term to make maximum use of its physical justification. The
physics behind this term in the case of a real liquid-droplet is as follows: To evaporate one liquid-droplet we have to use a definite heat energy as $Q_v M_m A$ where $Q_v$ accounts for the {\it vaporization latent heat}, $M_m$ is the mass of each molecule and $A$ is the total number of molecules in the
droplet. This heat energy is necessary to overcome all the attractive interactions
between the molecules. Hence, it is exactly equal to the binding energy of the droplet. This justifies the first term as
\begin{equation}\label{2}
B_V=Q_v M_m A=a_V A
\end{equation}
which results in the identification $a_V=Q_v M_m$. Therefore, we come to the conclusion that the binding energy per each molecule, namely $(B/A)$, corresponding to the volume term is independent of the total number of molecules $A$. This has a simple reason: the number of nearest neighbours of each interior
molecule is independent of the total size of the system (droplet), then $(B/A)$ becomes independent of $A$. {\it This is characteristic feature of every system in which the range of interaction between the constituent particles is short in comparison with the size of the system}. Of course, this feature undergoes a surface correction due to the finite size of the system according to (\ref{1}).
Moreover, if we consider the repulsive coulomb force between the protons,
then its negative contribution proportional to $Z^2$ ($Z$ is the atomic
number) affects the ratio $(B/A)$ in a more complicate way. One may add
also the contributions of an asymmetry term and a pairing term. In general,
all of these contributions to the volume term except for the pairing term
are negative.
The result is that the curve of $(B/A)$ with respect to $A$
is almost flat except for the small and large values of $A$. The decrease for small values of $A$, is due to the fact that the
full binding of each nucleon is accomplished when that nucleon is fully encircled
by other neighbor nucleons. However, this does not apply for the surface
nucleons; then for light nucleus having larger portion of surface nucleons
the binding energy for each nucleon in the nucleus is decreased. On the other
hand, the decrease for large values of $A$ results from the coulomb repulsive
force between each protons which is proportional to $Z^2$ and hence increases
more rapidly than $A$, resulting in a decrease of binding energy per each
nucleon. It is common to associate the almost flat portion of the curve with
the saturation property of the nuclear force which itself is emerged due
to the short range of this force. If this force would not be saturated
and the nuclear force would be a long range force, then each nucleon would
interact with all other nucleons with a term proportional to the size of
the nucleus or $A^2$. The more large size of the nucleus, the more
stabilized system, a result which is not favored by the nature. Hence, it
seems the saturation property or short range of the nuclear force is related
to the fact that nature avoids the structures with unlimited large size.
\section{Liquid-Droplet model for a typical galaxy }
To model the structure of the galaxy on the basis of Liquid-Droplet model
we need to revisit the concept of the {\it range} of gravitational interaction
in a large size structure of a galaxy. In fact, both the range of an interaction and the size of structure in which such interaction is governed are deeply related to the causal structure of space-time. Newton believed in the action of distance so that every gravitational effect, whatsoever far from a point, can be felt immediately at that point. The concept of a local gravitational field obeying the Lorentz invariance however changed this view: no gravitational effect at a large distance can be experienced at a point unless after a time interval that a signal with light velocity takes to enter the future light cone at that point. Even, in the Machian viewpoint where it is assumed that each gravitational constituent {\it feels} gravitationally all other constituents,
the concept of local gravitational field obeying the Lorentz invariance is
considered. This lorentz invariant causal structure then pose a crucial limit
on the {\it effective} range of gravitational interaction in a large size structure. In fact, it is reasonable to assume that the {\it effective} range of any interaction in an structure is modeled by the size of that structure. For example, the nucleus with its small size determines an effective short range of Yukawa nuclear force.
A deeper analysis on the basis of causal space-time structure may reveal
that the large size of a galaxy determines the effective range for gravitational
interaction inside this structure. Suppose, in a similar way that one calculates
the binding energy in a nucleus by Liquid-Droplet model, we calculate the binding energy in a galaxy. In the case of nucleus, we learned that each nucleon has full interaction just with neighbouring nucleons, due to short range of nuclear interaction. In the case of galaxy, we know gravitational interaction has long range. But, if we vaporize a small portion of a galaxy
to free all the stellar mass in this portion from their mutual gravitational interactions, then all other stellar mass far enough from this portion of the galaxy will not feel this vaporization until its signal (with light velocity)
can be reached by them. This exactly means that each local portion of a galaxy has an {\it effective} short range interaction compared with the large size of the whole galaxy. In other words, each small portion of the galaxy has just effective short range gravitational interaction with its limited neighborhood portions due to the limited velocity of light. This feature in the galaxy is similar to the feature in a massive nucleus where the range of effective nuclear force between close nucleons is short in comparison to the size of the nucleus. Therefore, if we accept in the context of Liquid-Droplet model that the effective range of gravitational interaction in a galactic structure is modeled by the effective size and mass of that structure, (an example of which is the short range nuclear force in the nucleus), then we may propose a model of gravitational interaction for the galaxy as follows
\begin{equation}\label{3}
U(r)=-G\frac{[k(r)m][Nm]}{R},
\end{equation}
where $U$ is the {\it effective} gravitational potential in a galaxy containing
$N$ gravitational constituents of average mass $m$, $R$ is the radius (effective size) of the galactic bulge , $G$ is the gravitational constant, and $k(r)$
is the $r$-dependent number of neighbor constituents for each individual constituent interacting with them.
This form of gravitational potential may be partly interpreted as {\it Machian }. We know the mass is a positive quantity and it is not possible to screen
the gravitational effects. Therefore, we assume that in any two (close) body process in the galaxy the extra influence of all other matter distributed over the effective size of the galaxy should be taken into account.
Accordingly, we can replace the whole gravitational effect of all effective
interactions at the position of a constituent of mass $m$ by the effect of a spherical shell of the {\it effective} mass $M$ and the effective radius $R$. This shell acts like a gravitational Faraday cage inside of which a gravitational potential exists
\begin{equation}\label{3'}
\Phi=-\frac{GM}{R}.
\end{equation}
From the Liquid-Droplet model, we shall assume that each constituent
gravitationally interacts with some limited neighbor constituents $k$. But,
this number will depend on the distance $r$ of that constituents from the center of the galaxy, as we shall explain later. By this kind of interaction we mean a {\it local} interaction. On the other hand, we assume a Machian
(or nonlocal) interaction, as explained above, contributing to $\Phi$ along with the local interaction. Therefore, if we consider the contribution of both local and nonlocal interactions to the gravitational potential (\ref{3'}) as the manifestation of effective mass $M$, and compare it with the gravitational potential energy (\ref{3}) trough $U=m \Phi$, then we obtain the effective mass $M$ as an {\it active} gravitational mass
\begin{equation}\label{4}
M(r)=k(r)Nm.
\end{equation}
This is a key feature of the present model in which the active gravitational
mass is different from the additive mass of the galaxy, namely $Nm$.
Moreover, we may associate the mean contribution $-Gm^2/R$ to each pair of
constituents so that the potential energy at the location $r$ will contain
$k(r)N$ terms where $k(r)$ and $N$ account for the local and nonlocal sectors
of interactions, respectively.
We expect that the linear dependence of $U$ on $N$ will result in the saturation of the gravitational force, similar to the saturation of the nuclear force in the nucleus. In fact, similar to the volume binding energy of a nucleus which is calculated using the Liquid-Droplet model, the {\it volume} binding energy of the galaxy is given
\begin{equation}\label{5}
B_V=a_V N,
\end{equation}
where the vaporization latent heat $a_V$ per each constituent in the galactic
bulge is given by
\begin{equation}\label{6}
a_V=\int_0^R G\frac{k(r)m^2}{R}dr.
\end{equation}
\section{Rotation curve }
In this section, we study the rotation curve of a typical galaxy upon
the Liquid-Droplet model. For each gravitational constituent located at a
distance $r$ from the center of galaxy having equilibrium state, one may approximately relate the order of magnitudes of the kinetic and potential energies as
\begin{equation}\label{7}
T \sim |U(r)|.
\end{equation}
Now, we introduce the very important role of the variable $k$ in the potential
energy $U$. In the Liquid-Droplet model, $k$ is the number of neighbor constituents interacting effectively with the individual constituent in consideration.
In the case of a real Liquid droplet or a nucleus this number is a constant which is determined by the geometric shape of packed molecules or nucleons. However, this is not the case for the galaxy where there are large voids between the constituents. Assuming a spherical symmetry for the high density mass distribution in the central galactic bulge, it is easy to realize that the number of neighbor constituents around each constituent $i$ located in a volume element grows with the distance $r$ similar to the volume element $dV= 4\pi r^2 dr $. Therefore, in the central galactic bulge we have
\begin{equation}\label{8}
k_{in}=\alpha r_i^2,
\end{equation}
where $r_i$ is the distance of the $i^{th}$ constituent from the center of galaxy
and $\alpha$ is a constant. The potential (\ref{3}) is then rewritten as
follows
\begin{equation}\label{9}
U(r_i)=-A r_i^2,
\end{equation}
where $A=\alpha{G Nm^2}/{R}$ is another constant. By substituting (\ref{9})
into the equation (\ref{6}) we obtain
\begin{equation}\label{10}
\frac{1}{2}m v_i^2 \sim A r_i^2,
\end{equation}
which leads to the familiar result in that section of rotation curve which
corresponds to the mass distribution in the galactic bulge, as
\begin{equation}\label{11}
v_i \sim r_i.
\end{equation}
For the outside of the central bulge where the mass is smoothly distributed with very low density (compared with the bulge) over each volume element, it is plausible to assume that each constituent located in the middle of
a volume element is surrounded by some of the other interacting constituents whose number is a very slow function of the distance $r_i$. A suitable choice is
\begin{equation}\label{12}
k_{out} = \ln{r_i},
\end{equation}
for which we obtain
\begin{equation}\label{13}
U(r_i)=-B \ln{r_i},
\end{equation}
\begin{equation}\label{14}
\frac{1}{2}m v_i^2 \sim B \ln{r_i},
\end{equation}
where $B=GNm^2/R$, or
\begin{equation}\label{15}
v_i \sim \sqrt{\ln{r_i}}.
\end{equation}
This result gives an almost flat rotation curve for the outside of the central bulge, in complete agreement with current observations.
We may repeat the calculations in a more dynamical way to obtain the almost
same results by resorting to Newton's second law. In the case of galactic bulge the Newton's law becomes
\begin{equation}\label{16}
\frac{m v_i^2}{r_i}=-\frac{dU}{dr_i}=2Ar_i,
\end{equation}
which gives the velocity proportional to the distance as
\begin{equation}\label{17}
v_i =\sqrt{\frac{2A}{m}}r_i.
\end{equation}
For the outside of the central bulge the Newton's law becomes
\begin{equation}\label{18}
\frac{m v_i^2}{r_i}=-\frac{dU}{dr_i}=\frac{B}{r_i},
\end{equation}
which gives a constant velocity
\begin{equation}\label{19}
v_i =\sqrt{\frac{B}{m}}.
\end{equation}
In order to match the velocities at $r_i=R$ we need to equate the followings
\begin{equation}\label{20}
\sqrt{\frac{2A}{m}}R=\sqrt{\frac{B}{m}},
\end{equation}
which results in
\begin{equation}\label{21}
\alpha=\frac{1}{2R^2}.
\end{equation}
Substituting this into Eq.(\ref{8}) leads to the important result that each constituent inside the central bulge is at most ($r_i=R$) effectively interacting with one half of another neighbor constituent. This shows that in the Liquid-Droplet model each constituent is {\it locally} interacting with a fraction $k$ of other neighbor individual constituent, while globally interacts with $N$ distant constituent. This fraction is very small close to the center of galaxy but approaches to ``one half" at the radius $R$ of the bulge.
Note that all the above results come out due to our special and simple assumptions
about the forms of the quantity $k(r)$. It is worth noting that the quantity $k(r)$ inside the galactic bulge seems to be almost the same for all galaxies, but outside the galactic bulge it may depend on some characteristic features of each galaxy like its shape, size, ellipticity, mass distribution , etc. Therefore, it is expected that in this model the rotation curve for the outside galactic bulge is not universally flat for all the galaxies, rather it differs for different galaxies according to their characteristic features; a result
which is observed in some galaxies.
\section{Virial theorem }
For a system of N point particles, the virial $G$ satisfies the following
equation \cite{Virial}
\begin{equation}\label{22}
\frac{dG}{dt}=2T+\sum_{i=1}^{N}{\bf F}_i \cdot {\bf r}_i,
\end{equation}
where $T$ is the total kinetic energy of the system, ${\bf F}_i$ is the net force on the $i^{th}$ particle at the position ${\bf r}_i$, and $G$ is defined
as
\begin{equation}\label{23}
G=\sum_{i=1}^{N}{\bf P}_i \cdot {\bf r}_i.
\end{equation}
We have realized that in the Liquid-Droplet model, the total force on each particle $i$ is acted by a fraction of only a neighbor particle. Moreover,
this fraction depends on the distance of that particle from the center of
galaxy. Therefore, the mutual interaction between all particles which is
usually considered in a gravitational system has now been replaced by this new type of limited interaction. If we assume the force to be derived from a potential energy, then we should accept that this potential is just a function of the distance of each particle from the center of galaxy $r_{i}$ rather than the relative distance $r_{ji}$ between the particles, namely ${\bf F}_{i}=
-{\bf \nabla}_{r_{i}}U$. Therefore, we have
\begin{equation}\label{25}
\sum_{i=1}^{N}{\bf F}_i \cdot {\bf r}_i=-\sum_{i=1}^{N}
\frac{dU}{dr_{i}}r_{i}.
\end{equation}
Hence, Eq.(\ref{22}) becomes
\begin{equation}\label{26}
\frac{dG}{dt}=2T-\sum_{i=1}^{N}
\frac{dU}{dr_{i}}r_{i}.
\end{equation}
For the central galactic bulge with $U_{in}\sim r_i^2$ we obtain
\begin{equation}\label{27}
\frac{dG}{dt}=2T-2\sum_{i=1}^{N_{in}}
U_{in}(r_{i})=2T-2U_{in},
\end{equation}
where $U_{in}$ is the total potential energy of the system in the galactic bulge
consisting of $N_{in}$ particles. On the other hand, outside the central bulge
of the galaxy consisting of $N_{out}$ particles with $U_{out}=-B \ln{r_i}$, we have
\begin{equation}\label{28}
\frac{dG}{dt}=2T+\sum_{i=1}^{N_{out}}B
=2T+N_{out}B.
\end{equation}
The virial theorem states that the time average $\langle dG/dt\rangle$ vanishes,
hence we obtain
\begin{equation}\label{29}
\langle T \rangle_{in}=\langle U_T \rangle_{in},
\end{equation}
for the inside, and
\begin{eqnarray}\label{30}
\langle T \rangle_{out}=-\frac{1}{2}N_{out}B ,
\end{eqnarray}
for the outside of the galactic bulge . Therefore, the total time average of the kinetic energy of the whole galaxy is obtained as
\begin{eqnarray}\label{31}
\langle T \rangle_{tot}&=&\langle U_T \rangle_{in}-\frac{1}{2}N_{out} B\\ \nonumber
&=&\frac{GNm^2}{R}\left(\sum_i^{N_{in}} {\langle k_{in}\rangle}-\frac{1}{2}N_{out} \right).
\end{eqnarray}
By interpreting $N_{out} B$ as $\langle U_T \rangle_{out}$, we obtain
\begin{eqnarray}\label{32}
\langle E \rangle_{in}=\langle T \rangle_{in}+\langle U_T \rangle_{in}=2\langle T \rangle_{in},
\end{eqnarray}
\begin{eqnarray}\label{33}
\langle E \rangle_{out}=\langle T \rangle_{out}+\langle U_T \rangle_{out}=-\langle T \rangle_{out},
\end{eqnarray}
\begin{eqnarray}\label{34}
\langle E \rangle_{tot}=2\langle T \rangle_{in}-\langle T \rangle_{out}.
\end{eqnarray}
In a stable gravitationally bound system we require $\langle E \rangle_{tot}<0$ which leads to
\begin{eqnarray}\label{35}
\langle T \rangle_{in}<\frac{1}{2}\langle T \rangle_{out}.
\end{eqnarray}
This is an interesting result which states that the total average kinetic energy in the galactic bulge is less than the half total average kinetic energy in the outside of galactic bulge.
\section{Tully-Fisher relation }
The Tully-Fisher relation is an empirical relationship between the intrinsic luminosity (proportional to the stellar mass) of a spiral galaxy and its amplitude of its rotation curve (which sets the total gravitational mass)
\cite{TF}. It has been turned out that on large scales most astronomical systems like
a spiral galaxy have much larger mass-to-light ratios than the central parts.
This fact may be explained by the relation (\ref{4}) as follows. In the present model, the gravitational mass is the active mass $M$. On the other hand,
the luminosity which is proportional to the stellar mass becomes proportional
to the additive mass $Nm$. Therefore, the mass-to-light ratio becomes proportional
to
\begin{eqnarray}\label{36}
\frac{M}{Nm}=k(r).
\end{eqnarray}
Since in both the galactic bulge and outer region, $k(r)$ is an increasing function of the distance from the center of galaxy $r$ (see Eqs.(\ref{8}), (\ref{12})), hence we have much larger mass-to-light ratios than the central parts which is in complete agreement
with the Tully-Fisher relation.
From another point of view we may study the Tully-Fisher relation in the
present model. We may rewrite the equation (\ref{19}) as
\begin{eqnarray}\label{37}
v^4=\left(\frac{GNm}{R}\right)^2.
\end{eqnarray}
If we consider $R$ as the effective size of the galaxy and $Nm$ as the
approximate stellar mass, then assuming a roughly constant mass density we
may write $Nm \sim R^3$ or $R^2$ according to whether we assume a spherical shape or disk shape for the whole galaxy, respectively. Therefore, Eq.(\ref{37}) reads
\begin{eqnarray}\label{38}
v^{\alpha} \sim Nm,
\end{eqnarray}
where $\alpha=3$ or 4 corresponding to spherical or disk shapes for the galaxy, respectively. This is an interesting result which again accounts
for the empirical Tully-Fisher relation.
\section{Liquid-Droplet model and MOND}
According to MOND, the rotational velocity is given by $v=(GNma_0)^{\frac{1}{4}}$.
However we find that in the present model this velocity is obtained by
$v=\left(\frac{GNm}{R}\right)^{\frac{1}{2}}$. Comparison of these two velocities
leads to the following result
\begin{eqnarray}\label{39}
a_0=\frac{GNm}{R^2}.
\end{eqnarray}
This means the constant acceleration $a_0$ which has no clear physical explanation in the MOND is explained in the present model through (\ref{39}) indicating
that this acceleration is nothing but the {\it effective} gravitational strength
of each galaxy. To justify this relation
we consider for example the Andromeda galaxy whose stellar mass $Nm$ and bulge radius $R$ are approximately $7\times 10^{11} SM$ and $10^5 LY$, respectively. Putting these values into (\ref{39}) gives $a_0 \sim 10^{-10}$ which is in good agreement with the value obtained by Milgrom. It seems similar calculation
for other galaxies gives almost the same order of magnitude for the value
of $a_0$.
\newpage
\section{Conclusion}
The problem of rotation curves for the galaxies has been investigated in the
context of a new model which may be considered as an alternative to the dark
matter and MOND. This model is motivated by the so called Weizs\"acker's
{\it Liquid-Droplet} model which had been used to describe the structure of the nucleus, according to which the nuclear force is saturated inside
the nucleus to exhibit a short range force and a flat curve of binding energy. Similarly, we have tried to show that the reason for the flat rotation curve
of a galaxy may have its origin in the fact that the gravitational force
exhibits an effective local short range interaction in the galactic structure
while representing characteristic features of Mach principle. This kind of gravitational interaction at the galactic scale gives the rotation curve in good agreement with observations. We have also developed the Virial theorem and also obtained the Tully-Fisher relation. Moreover, we have given a physical explanation for the so called {\it constant} acceleration in the MOND.
|
\section{Introduction}
\label{sec:intro}
The mechanism by which angular momentum is lost by the material in accretion discs is a classic problem in astrophysics. While gravitational torques associated with a binary companion (e.g.~\citealt{SMIH87}) or gravitational instabilities in massive discs \citep{ARS89} may play an occasional role, magnetic tension appears likely to be the most common source of the torque required to remove angular momentum from the material in the disc and permit it to spiral towards the central object.
Magnetically-mediated accretion disc models fall into two broad paradigms. In disc-jet models, a large-scale poloidal field threads the disc and the magnetic tension of the field lines anchored to the disc accelerates material outward and upward along the field lines away from the disc surfaces to form a collimated bipolar jet \citep{BP82}. The net effect is to transfer angular momentum from the disc material to the accelerated outflow. In the absence of an outflow the disc may still be braked through torsional magnetic coupling to an extended envelope or surrounding cloud \citep[e.g.][]{KK02}.
In models invoking magnetically-driven turbulence, on the other hand, tension in a weak magnetic field virulently destabilises the orbital shear flow in the disc via the magnetorotational instability (MRI; \citealp{V59}; \citealp{C60}; \citealp{BH91}), driving MHD turbulence that transports angular momentum radially outwards (e.g.\ \citealp{BH92}; \citealp{SHGB96}). Generally, one expects that the jet models apply when the magnetic and thermal pressures are comparable at the disc mid-plane, while substantially sub-thermal magnetic fields are subject to the MRI and are incapable of driving strong jets. It is conceivable that these two transport mechanisms may operate at the same radius for intermediate strength fields, with the MRI operating for a range of $z$ adjacent to the mid-plane of the disc and vertical transport dominating at greater heights (e.g.~\citealt{SKW07}). Similarly, as the field strength is expected to decrease with radius, there could also be an inner region where jets dominate, with the MRI operating at larger radii \citep{CF08}.
In this paper we focus on the efficacy of the MRI-driven turbulence in protostellar/protoplanetary discs. In this context, apart from transporting angular momentum, MRI-driven turbulence selectively heats the disc, effectively mixes chemical species
\citep[e.g.][]{SWH06}, and affects the transport, aggregation, and sedimentation of dust grains \citep{JK05, TWBY06, FP06}, the assembly blocks of planetesimals in the `core accretion' scenario of planet formation. Furthermore, the effective viscosity of the disc associated with MHD turbulence determines the ability of protoplanets to open gaps in the disc and also contributes to the delicate imbalance of gravitational torques that determines the rate and direction of planetary migration e.g.~\citep{MP03, NP04, JGM06}.
It is of great importance, therefore, to determine the location, extent and behaviour of the magnetically-active regions of protoplanetary discs, which are restricted by the very low fractional ionisation beyond $\sim 0.1$ au from the central star, where thermal ionisation is ineffective \citep{G96}. The dominant sources of ionisation -- X-rays or UV radiation from the forming star, and possibly interstellar cosmic rays -- are unable to penetrate to the disc mid-plane except perhaps at larger radii where the column density is low. At intermediate radii a very low level of ionisation is maintained by cosmic rays or, if these are absent, the decay of radioactive elements mixed with the gas (\citealt{SWH04, GFMW05}).
MRI--driven MHD turbulence is therefore thought to be restricted to the surface layers \citep{G96} over the range $\sim 0.3$--$20$ au where shielding of external ionisation sources results in very low levels of ionisation.
Assessments of the location and extent of magnetically-active zones (or equivalently, their complementary magnetic `dead zones') typically invoke only damping of the linear MRI by Ohm diffusion (e.g.~\citealt{H81,J96,FTB02,MP03,IN06b,TS08}). However, Ohm diffusion is the dominant magnetic diffusivity only at the very high densities achieved close to the disc mid-plane within 1 au of the central star. Instead, Hall diffusion\footnote{Despite the lack of dissipation associated with Hall drift we adopt the term ``Hall diffusion'' to maintain consistency with the notion of a tensor diffusivity appearing in the induction equation.} dominates at the intermediate mid-plane densities between 1 and 30 au \citep{WN99,SS02a}, and ambipolar diffusion beyond about 30\,au. Note, however, that as the density decreases towards the disc surfaces, different diffusivities may dominate in different layers. The neglect of ambipolar diffusion in assessing magnetic activity is less drastic, as it dominates near the disc surface where the fractional ionisation is relatively high and magnetic diffusivity may be insufficient to stabilise the MRI (e.g. \citealt{PC11}).
Hall diffusion provides a dissipation-free pathway for the magnetic field that either enables the MRI to grow despite the damping effect of Ohm or ambipolar diffusion or suppresses it entirely \citep{W99,BT01}, depending on the orientation of the magnetic field. Based on the linear growth rates Hall diffusion may drastically extend or restrict the reach of magnetic activity in protoplanetary discs \citep{W07} \emph{and likely modifies the transport and dissipative properties of the resulting turbulence.} However, pioneering shearing-box calculations including both Ohm and Hall diffusion did not appear to confirm this expectation, suggesting instead that Hall diffusion was unable to counter the damping effect of Ohm diffusion \citep{SS02a,SS02b}. As a result, subsequent evaluations of the extent of dead zones in protoplanetary discs are based either on the damping effects of Ohm diffusion (e.g. \citealt{TS08,TD09,KL10}), or on Ohm and ambipolar diffusion \citep{PC11,B11c}.
The apparent failure of fully-developed turbulence to respect the underlying linear instability in the Hall-diffusion-dominated limit is surprising, as it does so in the Ohm (e.g.~\citealt{TS08}) and ambipolar diffusion regimes \citep{BS11}. However, closer inspection of the parametrisation of the diffusive effects adopted by \citet{SS02a,SS02b} reveals that their simulations did not probe the important regime in which (a) Hall diffusion dominates Ohm diffusion \emph{and} (b) both dominate inductive effects. Indeed, none of the results of \citet{SS02a,SS02b} conflict with expectations based on the linear analysis. The confusion that has arisen can largely be attributed to the adoption of a magnetic Reynolds number $Re_M$ and a Hall parameter $X$ to characterise the magnitudes of Ohm and Hall diffusion respectively. Then the criteria for Ohm and Hall diffusion to dominate the inductive term are $Re_M<1$ and $X>2$, respectively. Crucially, the ratio of Hall to Ohm diffusion is then $\frac{1}{2} XRe_M$. This is never $\gg 1$ when $Re_M<1$ in the simulations, apart from the zero-net-flux simulations S14--S16 of \cite{SS02b} in which Hall diffusion suppresses the MRI in the $B_z<0$ regions, as one might expect based on the linear analysis.
The purpose of this paper is to emphasise the importance of Hall diffusion for MRI-driven turbulence in protoplanetary discs. We do this by considering the simplest example of the MRI -- perturbations with wave vectors parallel to an initially vertical magnetic field -- and demonstrating that the column density of the MRI-unstable region at 1\,au in the minimum mass solar nebula increases or decreases (depending on the sign of $B_z$) by an order of magnitude when Hall diffusion is included (see Fig. \ref{fig:active-sigma}). We begin by describing the field-line drifts induced by ambipolar, Hall, and Ohm diffusion in \S2 and give a qualitative discussion of their effect on the MRI. In \S3 we present an overview of the results of a linear analysis of the MRI in the presence of diffusion and give a coherent survey of its dependence on Pedersen (ie Ohm+ambipolar) and Hall diffusion by considering contours of the growth rate and wave number of the fastest growing mode and the range of MRI-unstable wave numbers in \S4. We then compare previous notation with the diffusivity notation we now favour and discuss the numerical simulations of \citet{SS02a,SS02b}. In \S5 we estimate the column density of the MRI-unstable layer at 1\,au in the MMSN using the criterion that there should exist a local unstable MRI mode with wave number satisfying $kh>1$. The results are dramatic, indicating that Hall diffusion increases the column density of the MRI-unstable surface layers by an order of magnitude depending on whether $B_z$ is positive or negative, respectively. Finally, we summarise our results and conclusions in \S6.
\section{Magnetic diffusion and the magnetorotational instability}
\label{sec:diffusion}
\subsection{Magnetic diffusion}
The magnetic field evolves according to the induction equation
\begin{equation}
\delt{\bm{B}} = \nabla \cross(\bm{v}\bm{\times}\bm{B}) - c\nabla \cross\E'
\,.
\label{eq:induction}
\end{equation}
where $\E'$ is the electric field in the local instantaneous rest frame of the fluid, which satisfies the generalised Ohm's Law
\begin{equation}
c \E' = \eta_\mathrm{A} (\nabla \cross\bm{B})_\perp + \eta_\mathrm{H}(\nabla \cross\bm{B})\bm{\times}\hat{\bm{B}}+ \eta_\mathrm{O}(\nabla \cross\bm{B})\,,
\label{eq:E-fluid-P}
\end{equation}
where we use subscripts $\parallel$ and $\perp$ to refer to the orientation with respect to the local magnetic field, and $\eta_\mathrm{A}$, $\eta_\mathrm{H}$, and $\eta_\mathrm{O}$ are the ambipolar, Hall, and Ohm diffusivities, respectively.
It will prove useful to recast the induction equation in a form that makes explicit the drift of the magnetic field through the fluid, i.e.
\begin{equation}
\delt{\bm{B}} =
\nabla \cross\left[(\bm{v}+\bm{v}_{\rm B})\bm{\times}\bm{B} \
-\eta_\mathrm{O}(\nabla \cross\bm{B})_\parallel\right]
\,,
\label{eq:induction-v2}
\end{equation}
where
\begin{eqnarray}
\bm{v}_{\rm B} & = & c\;\frac{\E'\bm{\times}\bm{B} }{B^2} \nonumber\\
& = & \eta_\mathrm{P}\,\frac{(\nabla \cross\bm{B})_\perp\bm{\times}\hat{\bm{B}}}{B}
\,-\,\eta_\mathrm{H}\,\frac{(\nabla \cross\bm{B})_\perp}{B} \,.
\label{eq:vB}
\end{eqnarray}
and the Pedersen diffusivity $\eta_\mathrm{P}$ is given by
\begin{equation}
\eta_\mathrm{P} = \eta_\mathrm{A}+\eta_\mathrm{O},
\label{eq:etaP}
\end{equation}
This form makes explicit the role of the diffusivities in transporting flux, and emphasizes that only Ohm diffusion is capable of destroying magnetic flux (i.e.\ via the $\eta_\mathrm{O}(\nabla \cross\bm{B})_\parallel$ term).
The diffusivities are determined by the abundances of charged particles and their collision cross-sections with the neutrals (e.g.~\citealt{C57,WN99}), and -- when the fractional ionisation is not too small -- with each other. In general, $\eta$, $\eta_\mathrm{A}$, and $\eta_\mathrm{P}$ are all positive, whereas $\eta_\mathrm{H}$ is positive or negative depending on whether positively- or negatively-charged species, respectively, are on average more decoupled from the magnetic field by collisions. For example, for a weakly-ionised ion-electron-neutral plasma we obtain
\begin{equation}
\eta = \frac{c^2 m_e\gamma_e\rho}{4\,\pi e^2 n_e}\,,
\label{eq:eta}
\end{equation}
\begin{equation}
\eta_\mathrm{A} = \frac{B^2}{4\,\pi\,\gamma_i\,\rho\rho_i}\,,
\label{eq:etaA}
\end{equation}
and
\begin{equation}
\eta_\mathrm{H} =\frac{c\,B}{4\,\pi\,e\,n_e}\,.
\label{eq:etaH}
\end{equation}
(e.g~\citealt{K89,BT01}).
The diffusivities exhibit more complex dependencies on density, field strength and electron abundance when charged grains are present. In this case the Hall diffusivity is negative at relatively low densities when neutral collisions decouple grains from the magnetic field, but not the electrons and ions, and is positive at higher densities when ions are also decoupled from the magnetic field \citep[e.g.][]{WN99}. The diffusivities exhibit significant spatial gradients if the ionisation rate and gas density do so, as in protoplanetary discs (e.g.\ \citealt{W07}). Fortunately, for a linear analysis the response of the diffusivities to perturbations is not needed as $\nabla \cross\bm{B}$ vanishes in the initial equilibrium state. Thus in our local analysis we can regard the diffusivities as specified constants, but we shall return to them when considering the application to protoplanetary disc s in \S\ref{sec:ppds}.
\subsection{Effect on the MRI}
Before deriving the dispersion relation for the magnetorotational instability\ in diffusive
media we first give an intuitive, physical description of the
effect of magnetic diffusion on its growth and properties.
We first examine the growth of the instability in a Keplerian accretion
disc under ideal-MHD conditions and assuming an initially vertical magnetic field (see Fig.~\ref{fig:MRI-sketch}, top-left panel). We consider, in particular, the situation where alternate layers of fluid are displaced from their equilibrium position, such that point 1 moves radially inwards (and forwards in azimuth) while point 2 is shifted radially outwards (and backwards in azimuth). In this case the magnetic field is buckled because the field lines are frozen into the fluid (top-right panel). The deformation of the lines creates magnetic tension forces which are directed outwards and backwards in azimuth (at point 1), or inwards and forwards in azimuth (at point 2), as depicted in the lower panel of Fig.\ 1. At point 1 then the magnetic tension provides some radial support against gravity and supplies a negative torque. This means that the fluid element at point 1 will lose angular momentum and spiral inwards. The opposite takes place at point 2, with the fluid element there spiralling outwards. As a result, the magnetic field buckling and associated tension increases, leading to runaway growth. The fastest growing mode has growth rate $\frac{3}{4}\,\Omega$ on scales with wave numbers $k\sim \Omega/v_{\rm A}$, where $\Omega = v_{\rm K}/r $ is the Keplerian frequency and $v_{\rm A}$ is the Alfv\'en speed \citep{BH91}.
\begin{figure}
\centering
\includegraphics[scale=0.9,trim= 0 0 65 0]{fig1.pdf}
\caption{A sketch of the development of the magnetorotational instability (MRI) in an initially vertical magnetic field $B_0$ embedded in an unstratified, Keplerian disc for a vertical wave vector. Top panel shows a poloidal view of (\emph{left}) the initial field configuration and (\emph{right}) the perturbed field and associated current density ($\bm{J}$), Lorentz force ($\bm{J} \bm{\times} \bm{B}$) and deviation from local Keplerian velocity. The lower panel shows in plan view the perturbed fluid elements at points 1 and 2. At point 1 the magnetic tension provides some radial support against gravity and supplies a negative torque. As a result the material here is sub- Keplerian and moving inwards, as indicated by the green vector which indicates the departure from Keplerian rotation. The field transfers the angular momentum to the fluid at locations such as 2 where the field is buckled outwards, and the fluid there spirals out. The differential radial motion of the fluid that is losing and gaining angular momentum at 1 and 2 respectively enhances the buckling, leading to runaway growth.
\label{fig:MRI-sketch}
}
\end{figure}
Magnetic diffusion modifies the above picture by allowing slippage between the field lines and the fluid, so that there is no longer a direct connection between the relative displacement of the fluid layers and the buckling of the magnetic field lines. One might expect that this effect should always reduce the instability because the displacement of the field would lag that of the fluid. However, this intuition is based on the limit of Ohm or ambipolar diffusion, in which the drift is in the direction of the magnetic stresses and tends to straighten up field lines. Hall diffusion, by contrast, creates a drift orthogonal to the tension forces and may, therefore, enhance or suppress the radial buckling depending on the situation. It is this feature that gives Hall diffusion its unique properties and that we aim to explain here.
In the ambipolar diffusion limit, the magnetic field is frozen into the ions and electrons, which drift together through the neutral component of the fluid. Collisions with the neutrals then transmit magnetic stresses to the bulk of the gas. As the fractional ionisation is low, the ion and electron inertia and thermal pressure are negligible and the drift velocity of the ions, electrons and field is determined by the balance between the Lorentz force on the ions and the collisional drag with the neutrals,
\begin{equation}
\bm{v}_{\rm P} = \bm{v}_i - \bm{v} = \frac{\bm{J}\bm{\times}\bm{B}}{c\gamma_i\rho_i\rho}\,,
\label{eq:ion-mom}
\end{equation}
and so is parallel to $\bm{J}\bm{\times}\bm{B}$. In eq. (\ref{eq:ion-mom}),
$\rho$ and $\rho_i$ are the neutral and ion density, respectively, and
\begin{equation}
\gamma_i = \frac{\langle \sigma v \rangle_i}{m_i + m} \,,
\end{equation}
where $\langle \sigma v \rangle_i$ measures the rate coefficient of momentum exchange via collisions with the neutrals, taken to have mean mass $m$, and $m_i$ is the ion mass. When $\bm{J}\cdot\bm{B}=0$, such as envisaged here, Ohm diffusion also produces a drift in the same direction, so we have adopted the subscript P to denote the Pedersen drift (see eq \ref{eq:vB}).
Fig.\ \ref{fig:MRI-drift-AD} shows the field line drift (red vectors) for the fluid configuration depicted in Fig.\ \ref{fig:MRI-sketch}, but now incorporating the effect of ambipolar or Ohm diffusion. The black vectors indicate the total drift of the magnetic field in the local Keplerian frame, and correspond to the vector sum $\bm{v}+\bm{v}_{\rm B}$ (see eq \ref{eq:vB}).
In this case the net effect of the drift is to reduce the radial and
azimuthal stretching of the field, and therefore -- as one might guess -- reduce the degree of instability.
\begin{figure}
\centering
\includegraphics[scale=0.8,trim= 0 0 0 0]{fig2.pdf}
\caption{Plan view of the effect of ambipolar or Ohm diffusion
on the development of the MRI. The current density and associated
magnetic stress on the fluid at points 1 and 2 of Fig.
\ref{fig:MRI-sketch} are indicated in blue. Because of finite
diffusivity the field lines drift through the fluid at velocity $\bm{v}_\textrm{B}$ (\emph{red}
vectors) in the direction parallel to $\bm{J}\bm{\times}\bm{B}$ (see eq
\ref{eq:ion-mom}). The net drift of the field line with respect to
Keplerian rotation, $\bm{v}+\bm{v}_\textrm{B}$ is indicated by the \emph{black}
vector. In this case the net drift is marginally inwards
at point 1. At
point 2, the current density and magnetic stresses are reversed, so
the field drifts in the opposite direction. As a result the field
lines drift inwards and outwards at reduced rates at points 1 and
2 respectively, and the instability proceeds more slowly than in
ideal MHD.}
\label{fig:MRI-drift-AD}
\end{figure}
In the Hall limit, collisions with the neutral gas are frequent and strong
enough to decouple the ions from the field, but the electrons -- which have
a smaller collision cross section with the neutrals and a higher
charge-to-mass ratio -- remain well coupled. The magnetic field is then
frozen into the electrons, and these drift together through the neutrals
and ions, which remain tightly-coupled by collisions. In this limit, the
field drift speed through the neutrals is given by the ion-electron drift
and hence is antiparallel to the current density,
\begin{equation}
\bm{v}_{\rm H} = \bm{v}_e - \bm{v}_i = - \frac{\bm{J}}{e n_e} \,.
\label{eq:vB-hall}
\end{equation}
In eq. (\ref{eq:vB-hall}), $n_e$ ($= n_i$) is the electron number
density and $e$ is the elementary electric charge.
More generally, when all diffusion mechanisms are taken into account the drift
velocity $\bm{v}_{\rm B}$ of the field lines through the fluid is the sum of
these two orthogonal contributions (see eq \ref{eq:vB} below). The
implications for the magnetorotational instability\ are sketched in Fig.\
\ref{fig:MRI-drift-a}, which now includes a Hall drift antiparallel to
$\bm{J}$.
\begin{figure}
\centering
\includegraphics[scale=0.8,trim= 0 0 0 0]{fig3.pdf}
\caption{Plan view of the effect of magnetic diffusion
[incorporating the ambipolar, Ohm and Hall ($> 0$) terms] on the
development of the MRI. The current density and associated magnetic
stress on the fluid at points 1 and 2 of Fig. \ref{fig:MRI-sketch} are
indicated in blue. At point 1 the magnetic tension provides some
radial support against gravity and supplies a negative torque. As a
result, the material here is sub-Keplerian and moving inwards, as
indicated by the green vector (which denotes departure from Keplerian
rotation). Because of
the
finite diffusivity, the field lines drift
through the fluid (\emph{red} vectors) with a velocity determined by
the magnitudes of the ambipolar, Ohm and Hall diffusivities (see eq
\ref{eq:vB}). Ambipolar and/or Ohm diffusion contributes a drift
$\bm{v}_{\rm P}$ parallel to $\bm{J}\bm{\times}\bm{B}$, while the Hall drift $\bm{v}_{\rm
H}$ is antiparallel to the current density $\bm{J}$. The net drift of the
field line with respect to the Keplerian rotation, $\bm{v}+\bm{v}_{\rm B}$,
is indicated by the \emph{black} vector. In this case the net drift is
inwards. At point 2, the current density and magnetic stresses are
reversed, so the field drifts in the opposite direction. As a result
the field lines drift inwards at point 1 and outwards at point 2 and the
instability proceeds.}
\label{fig:MRI-drift-a}
\end{figure}
The generic geometry of the MRI means that the Hall drift at point 1
is directed inwards and retrograde in azimuth, and it is outwards and
prograde at point 2. The net effect is to exacerbate the radial
buckling of the field, but to reduce the buckling in the azimuthal
direction. We show in the next section that it is the radial effect
that is critical, as azimuthal field is always created out of the
underlying Keplerian shear of the disc. So Hall diffusion in this
case
(e.g. when the magnetic field is aligned with the angular velocity vector of the disc)
tends to be destabilising.
This is not always the case, however, because the direction of the current
density, and therefore of the Hall diffusion, reverses upon global reversal of
the magnetic field. This situation is illustrated in Fig.\
\ref{fig:MRI-drift-b}.
\begin{figure}
\centering
\includegraphics[scale=0.8,trim= 0 0 0 0]{fig4.pdf}
\caption{As for Fig.\ \ref{fig:MRI-drift-a}, but after a global
reversal of the magnetic field, as would be expected if the
initial magnetic field was antiparallel to the disc's angular
velocity vector. In this case, the magnetic stresses are unaffected --
and so, therefore, are the fluid velocity and Ohm/ambipolar
drift current density. However, the current density is reversed,
and therefore so is the Hall drift. The magnetic field now drifts
outwards at 1 and inwards at 2, reducing the radial buckling. The MRI is
partly or entirely suppressed in these circumstances (see text).}
\label{fig:MRI-drift-b}
\end{figure}
The magnetic stresses, the fluid velocity, and the field-line drift
associated with Ohm and ambipolar diffusion are not affected by this
reversal.
Note, however, that Hall diffusion now acts to stabilise the disc by acting against
the radial buckling of the field that would otherwise be driving the fluid
motions.
To summarise, Ohm and ambipolar diffusion are stabilising effects,whereas Hall diffusion may be destabilising or stabilising depending on whether the initial vertical magnetic field is parallel or antiparallel to the rotation axis, respectively. This asymmetry reflects the fundamental asymmetry in the microscopic properties of the positive and negative charged species in the fluid in this limit.
\section{Linear Analysis of the MRI}
\label{sec:formulation}
We consider a small region of an axisymmetric, geometrically thin and nearly Keplerian disc, threaded by a vertical magnetic field, with sound and Alfv\'en speeds ($c_{\rm s}$ and $v_{\rm A}$) at the mid-plane that are both small compared to the local Keplerian speed $v_{\rm K}$. We assume that radial gradients are on the scale of $r$ and neglect vertical stratification of the initial equilibrium state, so that our analysis only holds near the mid-plane at heights $z\ll c_s /\Omega$. The initial state is in Keplerian rotation with a uniform density, pressure, and vertical magnetic field $\bm{B}=sB\zh$ (where $s=\pm1$). We linearise the equations around this state and seek solutions for axisymmetric perturbations of the form $\exp(\nu t - ikz)$. The equations for perturbations in density, pressure and $v_z$ form a separate system that describes vertically-propagating sound waves. The system of linear equations in the remaining perturbations involve fluctuations in the $r$ and $\phi$ components of $\bm{B}$, $\bm{v}$ and $\bm{v}_{\rm B}$.
A derivation of the dispersion relation is outlined in Appendices
\ref{app:formulation} and \ref{app:analysis}. This dispersion relation was first derived by \cite{W99}, and later extended to more general geometries in the Hall-Ohm limit \citep{BT01}, and then including ambipolar diffusion \citep{D04}; the ambipolar diffusion limit was also considered by \citet{KB04}. Our emphasis here is on the dependence of the instability on the Pedersen and Hall diffusivities $\eta_\mathrm{P}$ and $\eta_\mathrm{H}$; a detailed analysis is presented in the appendices. An overview is provided by Fig.\ \ref{fig:locii}, which illustrates the qualitative changes in the growth rate vs wave number curves for different choices of the diffusivities.
First, note that the ideal-MHD limit holds at the origin (i.~e.~$\eta_\mathrm{H}=\eta_\mathrm{P}=0)$, the Hall MHD limit holds along the horizontal axis ($\eta_\mathrm{P}=0$) and the Ohm or ambipolar diffusion limits hold along the vertical line ($\eta_\mathrm{H}=0$). Recall also that only the half plane $\eta_\mathrm{P} \geq 0$ is physically relevant. It turns out that there are three distinct forms of the resulting $\nu(k)$ curve, corresponding to regions labelled I, II and III in the plane, as illustrated in Fig.\ \ref{fig:locii}. In region I wave numbers less than a cutoff $k_c$ are unstable with the maximum growth rate attained at an intermediate wave number $k_0$ (see eqs \ref{eq:kc}, \ref{eq:nu0-1}, and \ref{eq:k0}). In region II all wave numbers are unstable with the maximum growth rate still occurring at finite wave number. Finally, in region III all wave numbers are unstable and the maximum growth rate is approached asymptotically as $k\rightarrow\infty$.
\begin{figure*}
\centering
\includegraphics[scale=0.8,trim= 30 0 0 0]{fig5.pdf}
\caption{Schematic dependence of the behaviour of the MRI on the Hall and Pedersen (Ohm+ambipolar) diffusivities, $\eta_\mathrm{H}$ and $\eta_\mathrm{P}$, for an initial magnetic field $sB\zh$ in a Keplerian disc, where $s=\pm1$ is the sign of $B_z$. The field is assumed to be weak so that stratification can be neglected. Physical values of the Pedersen conductivity $\eta_\mathrm{P}$ are non-negative, and there are no unstable modes for Hall diffusivities $s\eta_\mathrm{H}\Omega/v_A^2\leq -2$. The unstable region $s\eta_\mathrm{H}\Omega/v_A^2 > -2$ is subdivided according to the dependence of the growth rate on wave number $k$, which is sketched in the insets labelled I--III. In region I (outside the red locus given by eq.\ \ref{eq:kcinf-locus}), wave numbers less than a cutoff $k_c$ are unstable with the maximum growth rate $\nu_0$ attained at $k_0$ (see eqs \ref{eq:kc}, \ref{eq:nu0-1}, and \ref{eq:k0}). Between the red and blue loci (region II) all wave numbers are unstable with the maximum growth rate still occurring at finite wave number. Within the blue locus given by eq.\ \ref{eq:k0inf-locus} (region III) all wave numbers are unstable and the maximum growth rate is approached asymptotically as $k\rightarrow\infty$.}
\label{fig:locii}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[scale=0.8]{fig6.pdf}
\caption{Contours of the maximum growth rate of the MRI (solid black lines, units
$\Omega$) and corresponding wave number (dashed blue lines, units $\Omega/v_A$) as a function of Hall and Pedersen (Ohm+ambipolar) diffusivities. The innermost blue contour
shows where $k_0$ becomes infinite, corresponding to the boundary between regions II and III in Fig.\ \ref{fig:locii}. Within this region the fastest growth rate is approached asymptotically as $k\rightarrow\infty$; this transition is responsible for the curvature of the black contours within this region (see eqs \ref{eq:nu0-1} and \ref{eq:nu0-2}).}
\label{fig:contours}
\end{figure*}
Having delineated these three regions, we now consider how the critical
wave number $k_c$, fastest growth rate $\nu_0$ and corresponding wave
number $k_0$ vary across the entire $\eta_\mathrm{P}$--$\eta_\mathrm{H}$ plane.
Contours of the growth rate and wave number of the most unstable mode are plotted in this plane in Fig.\ \ref{fig:contours}.
The growth rate increases clockwise, from $0\,\Omega$ along the vertical line $s\eta_\mathrm{H}=-2$ up to $0.75\,\Omega$ for the horizontal line $\eta_\mathrm{P} =0$ for $s\eta_\mathrm{H}>-4/5$. In the absence of Hall diffusion, the maximum growth rate $\nu_0$ declines with increasing (Ohm and/or ambipolar) diffusivity (e.g. moving vertically upwards through the $s\eta_\mathrm{H} = 0$ point in the horizontal axis), with $\nu_0 \approx\textstyle{\frac{3}{4}\eta_\mathrm{P}^{-1}}$ for $\eta_\mathrm{P}\gg 1$. The most important effect of Hall diffusion, apparent from Fig.\ref{fig:contours}, is that \emph{the growth rate of the MRI exceeds $0.3\,\Omega$ for $s\eta_\mathrm{H}\ga \eta_\mathrm{P}$, even for arbitrarily large $\eta_\mathrm{P}$}. More generally, the addition of Hall diffusion at fixed $\eta_\mathrm{P}$ increases the growth rate if $s\eta_\mathrm{H}>0$ and decreases it when $s\eta_\mathrm{H}<0$. For large values of $\eta_\mathrm{P}$, eq (\ref{eq:nu0-1}) shows that $s\eta_\mathrm{H}/\eta_\mathrm{P} \approx 24\nu_0/(9-16\nu_0^2)$. It is this fact that has the potential to modify the extent of dead zones in protoplanetary discs, as we explore in \S \ref{sec:ppds}.
The wave number of the fastest growing mode (blue contours in Fig.\ \ref{fig:contours}) decreases as the diffusivity is increased. Again, the contours are not arranged so that the highest wave numbers occur in the ideal-MHD limit, but to the
$s\eta_\mathrm{H} < 0$ side, within the boundary between regions II and III (traced by the $k_0=\infty$ contour).
Contours of constant $k_c$
are semicircles, as plotted in Fig.\ \ref{fig:kcrit}.
While the range of unstable wave numbers is reduced for large values of
$s\eta_\mathrm{H}$ and $\eta_\mathrm{P}$, as one might expect, the range is not maximised in the
ideal limit (ie. at the origin) but in regions II and III, bounded by the
$k_c=\infty$ contour.
In general we note that increasing $\eta_\mathrm{P}$ decreases the maximum growth rate and the characteristic wave numbers, whereas increasing $s\eta_\mathrm{H}$ above $-2$ increases the maximum growth rate and may either increase (when $s\eta_\mathrm{H}+1\la\eta_\mathrm{P}$) or decrease (when $s\eta_\mathrm{H}+1\ga\eta_\mathrm{P}$) the corresponding wave number.
Overall, these patterns place the ideal, Ohm (or ambipolar) and Hall regimes in context, and for the first time we see an overview of the effect of magnetic diffusivity on the linear MRI. In particular, there is nothing special about the Ohm/ambipolar limit, e.g.\ the behaviour of the instability in the presence of diffusion is not qualitatively different for $s\eta_\mathrm{H}=2$ vs $\eta_\mathrm{H}=0$. Even the ideal-MHD limit does not stand apart as remarkable, although it still holds a special place conceptually because flux freezing holds and it is easier to visualise.
What does stand out is the part of the plane in the lower left, regions II and III, characterised by high wave numbers and the spraying out of the growth contours. In this part of the plane, the instability operates in the ``cyclotron limit'' $\eta_\perp \sim 1$ and $k^2\gg 1$ (see appendix \ref{app:limits}), in which the instability arises in the competition between magnetic diffusion and advection of the field by the fluid; generation of $B_\phi$ from $B_r$ by the Keplerian shear flow is negligible.
The
lack of any $k$ dependence in this regime occurs because both the magnetic diffusion and the magnetic stresses on the fluid (which are responsible for the fluid displacement) scale as $k^2$.
This short-wavelength, low-frequency limit corresponds to the cyclotron mode of the magnetised fluid, which has frequency $\omega_H = v_A^2/|\eta_\mathrm{H}|$ \citep{WN99,PW08}. This mode couples effectively to the Keplerian rotation as long as the sense of circular polarisation matches that of epicyclic motion, i.e. as long as $B_z\eta_\mathrm{H} < 0$. The other short-wavelength mode, the high-frequency whistler ($\omega \approx k^2 v_A^2/\omega_H$), is unable to couple effectively to the rotation \citep{W99}.
The maximum growth rate contours emerging from regions II and III continue on to attain another limit when $\eta_\mathrm{P}^2+\eta_\mathrm{H}^2 \gg 1$, in which the field evolves in response to shear and diffusion, without significant feedback from the perturbations that it induces in the fluid flow. Instability in this case relies on the keplerian shear flow generating $B_\phi$ from $B_r$, and the tendency of Hall diffusion to convert $B_\phi$ back into $B_r$. This brings the potential destabilising effect of Hall diffusion in shear flows \citep{K08} to the fore and shows that it is quite independent of rotational effects -- i.e.~the Coriolis and centripetal acceleration -- that drive the MRI. In Appendix \ref{app:limits} we obtain simple analytic expressions for growth rate in plane-parallel shear flows as a function of $k$ for arbitrary diffusivity, the results are plotted in Fig. \ref{fig:nu-k-diffusion}.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{fig7.pdf}
\caption{Contours of the critical wave wave number $k_c $ below
which instability sets in (units $\Omega/v_A$). The red innermost contour shows where
$k_c$ becomes infinite and all wave numbers are unstable,
corresponding to the boundary between regions I and II in Fig.\
\ref{fig:locii}. }
\label{fig:kcrit}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=8cm]{fig8.pdf}
\caption{Growth rate versus wave number for the Hall-driven diffusive
instability for a mutually orthogonal magnetic field, shear velocity, and velocity
gradient $v'$. For Keplerian rotation and cylindrical geometry, the effective
velocity gradient is $v' = \textstyle \frac{3}{2} \Omega$. The curves correspond to
different values of the ratio
of the Pedersen and Hall diffusivities.}
\label{fig:nu-k-diffusion}
\end{figure}
For more general field and wave vector configurations, ambipolar diffusion plays a similar role, albeit hindered by dissipation \citep[Pandey \& Wardle, in prep]{KB04,D04,K08}.
\section{Comparison with numerical simulations}
\label{sec:MRI-onset}
In this section we explore the relationship of our linear calculations to the nonlinear, unstratified shearing box simulations of MRI-driven turbulence by \citet{SS02a, SS02b}, which explicitly included both Ohm and Hall diffusion. These authors' results could be interpreted as implying that the presence of Hall currents has little effect on the critical (minimum) degree of magnetic coupling (the coupling between the neutral gas and the magnetic field) for the instability to operate\footnote{This field-matter coupling parameter is often referred to as the magnetic Reynolds number $R_{\rm EM}$. As we expound below, however, it really corresponds to the Elsasser number $\Lambda$ (see discussion below and Table \ref {table:numbers}).}. We show, however, that their calculations do not yet probe deeply enough into the Hall regime for Hall diffusion to be able to impact on the development and properties of the instability. In fact, their solutions are in agreement with expectations from local and stratified linear analyses for the same values of the parameters describing the initial fluid conditions. Finally, we use the ratios of the different terms in the induction equation in order to delineate the region of parameter space where Hall diffusion is expected to substantially modify the growth rate, and spatial scale, of the MRI-unstable modes. First, however, and for the sake of clarity, we discuss the notation used in the literature to characterise the fluid, and its magnetic activity, in diffusive environments.
\subsection{Fluid parameters}
\label{subsec:fluid}
\begin{table*}
\caption{Typical dimensionless numbers used to characterise the degree of coupling between the magnetic field and the neutral gas: Magnetic Reynolds No., Lundquist
No., and Elsasser No. For comparison, we also show the Reynolds No.~(although this parameter combination does not involve the magnetic diffusivity and, therefore,
does not measure the field-matter coupling). The fluid variables used to characterise the required scalings are the flow velocity ($v$), the Alfv\'en velocity ($v_
{\rm A}$), the kinematic viscosity ($\nu$) and the magnetic diffusivity $\eta$. As usual, $\Omega$ is the Keplerian angular frequency and $l$ is a typical,
unspecified, length scale of the flow.}
\begin{tabular}{|c|l|l|c|c|c|c|}
\hline
Symbol & \multicolumn{1}{|c|}{Dimensionless} & \multicolumn{1}{|c|}{Ratio of} & \multicolumn{3}{|c|}{Scaling} & Definition \\
& \multicolumn{1}{|c|} {number} & \multicolumn{1}{|c|}{physical quantities} & $V$ & $L$ & $D$ & $VL/D$ \\
\hline\\
$R_{\rm e}$ & Reynolds & Inertial to viscous forces & $v$ & $l$ & $\nu$ & $vl/\nu$ \\[3pt]
$R_{\rm EM}$ & Magnetic Reynolds & Inertial to resistive time scales & $v$ & $l$ & $\eta$ & $vl/\eta$ \\[3pt]
$S$ & Lundquist & Resistive diffusion to Alfv\'en time scales & $v_{\rm A}$ & $l$ & $\eta$ & $v_{\rm A} l/\eta$ \\[3pt]
$\Lambda$ & Elsasser & Lorentz to Coriolis forces & $v_{\rm A}$ & $v_{\rm A}/\Omega$ & $\eta_\perp$ & $v_{\rm A}^2/\eta_\perp
\Omega$ \\
\hline
\end{tabular}
\label{table:numbers}
\end{table*}
The growth and structure of the MRI are strongly dependent on the magnetic field strength, geometry, and the nature -- and magnitude -- of the magnetic diffusivity. Different parametrizations have been used in the literature to characterise these properties, depending on the formulation of the problem and the adopted form of the induction equation. Some authors use a tensor diffusivity, and the induction equation then takes the form
\begin{eqnarray}
\delt{\bm{B}} &=& \nabla \cross \left (\bm{v} \bm{\times} \bm{B}\right) - \nabla \cross \Big[
\eta \nabla \cross \bm{B}
\nonumber\\[6pt]
& & \quad \quad + \;\; \eta_{\rm H} (\nabla \cross \bm{B}) \bm{\times} \hat{\bm{B}} + \eta_{\rm A}
(\nabla \cross \bm{B})_{\perp} \Big]
\label{eq1:induction_drift}
\end{eqnarray}
(e.g.~\citealt{W07}). Others prefer to use separate equations for the charge carriers and the neutrals. In the latter `multifluid' approach, when the ionised species are ions and electrons only (denoted by the subscripts `i' and `e', respectively), the induction equation becomes (e.g. \citealt{K89,BT01,SS02a})
\begin{eqnarray}
\delt{\bm{B}} &=& \nabla \cross \left (\bm{v} \bm{\times} \bm{B}\right) - \nabla \cross \Bigg[
\frac{c m_e \nu_{en} \bm{J}}{n_e e^2}
\nonumber\\[6pt]
& & \quad\quad +\;\; \frac{\bm{J} \bm{\times} \bm{B}}{e n_e} - \frac{(\bm{J} \bm{\times} \bm{B}) \bm{\times} \bm{B}}{c
\gamma_i \rho \rho_i} \Bigg] \,.
\label{eq:induction_drift}
\end{eqnarray}
Comparison of equations (\ref{eq1:induction_drift}) and (\ref{eq:induction_drift}) yields the expressions given in (\ref{eq:eta}) to (\ref{eq:etaH}) for the diffusivities. On the r.h.s.~of equation (\ref{eq1:induction_drift}), as well as in the second line of equation (\ref{eq:induction_drift}), the terms (from left to right) denote the inductive (I), Ohm (O), Hall (H) and ambipolar diffusion (A) contributions to the evolution of the magnetic field, respectively.
The following parameters have typically been used to characterise the magnetic properties of the fluid:
\begin{enumerate}
\item \emph{Field strength}.
This property is commonly measured either by the ratio of the Alfv\'en speed to the isothermal sound speed ($v_{\rm A}/c_{\rm s}$), or by the plasma beta parameter $\beta = (2/\gamma) c_{\rm s}^2/v_{\rm A}^2$, where $\gamma$ is the adiabatic index of the fluid.
\item \emph{Field-matter coupling}.
The degree of coupling between the neutral matter and the magnetic field is, most generally, measured by a ratio of the type $V L/ D$ where $V$, $L$ and $D$ represent characteristic speed, length and diffusion scales of the flow, respectively. In MRI studies, it is appropriate to take $V = v_{\rm A}$, $L = v_{\rm A}/\Omega$ (the characteristic wavelength of MRI-unstable modes in ideal-MHD conditions) and $D = \eta_{\perp}$, the total perpendicular diffusivity [e.g. equation (\ref{eq:etaperp})]. The resulting ratio is the Elsasser number
\begin{equation}
\Lambda \equiv \frac{v_{\rm A}^2}{\Omega \eta_{\perp}} \,.
\label{eq:Lambda}
\end{equation}
This dimensionless number is formally defined as the ratio of the Lorentz force ($\bmath{J} \times \bmath{B})/c$ to the Coriolis force $\propto \rho (\bmath{v} \times \bmath{\Omega})$, which results in an expression of the form $v_{\rm A}^2/(LV)\Omega$. Adopting the magnetic diffusivity $\eta_\perp$ as the typical magnitude of the LV factor in the denominator yields expression (\ref{eq:Lambda}).
The Elsasser number is widely used in geophysics, as in a geodynamo process the magnetic field is thought to amplify until $\Lambda$ becomes of order unity (see, e.g.~\citealt{C10} and references therein). This ratio was denoted $\chi$ in the work of \citet{W99}, who expressed it as
\begin{equation}
\chi \equiv \frac{B^2 \sigma_{\perp}}{\rho c^2 \Omega} \equiv
\frac{\omega_c}{\Omega} \,,
\label{eq:chi}
\end{equation}
where $\sigma_{\perp} = c^2/(4 \pi \eta_\perp)$ and $\omega_{\rm c}$ is the critical frequency above which flux-freezing conditions break down, so that the generic non-ideal MHD term in the induction equation dominates over the inductive term. When the field-matter coupling is characterised by the Elsasser number, values $\gg 1$ ($\ll 1$) correspond to strong (weak) coupling. Similarly, from equation (\ref{eq:chi}) it is clear that when $\Lambda \gg 1$, the magnetic field is strongly coupled to the fluid at frequencies of order $\Omega$, the frequencies of interest for the study of the MRI.
Note that, in general, the Elsasser number is different from the combination $v_{\rm A} l / \eta$, which compares the resistive diffusion time $\tau_{\rm R} \propto l^2/\eta$ to the Alfv\'en time $\tau_{\rm A} \propto l/v_{\rm A}$ (e.g. the Lundquist number) and from the ratio $v l/\eta$ (the magnetic Reynolds number $R_{\rm EM}$, where $v$ is the fluid velocity). However, both terms have been used in the literature to refer to $\Lambda$. Table \ref{table:numbers} summarises the definitions and typical scalings associated with the dimensionless numbers discussed above, namely, the magnetic Reynolds, Lundquist and Elsasser numbers. For clarity we also list the Reynolds number, although this does not measure magnetic coupling,
as it deals with the viscosity of the fluid instead of its magnetic diffusivity.
Finally, sometimes $c_{\rm s}$ is adopted as the characteristic velocity scale of the flow instead of $v_{\rm A}$, resulting in the ratio $c_{\rm s}^2/\eta \Omega$ (e.g.~\citealt{FSH00}).
\item \emph{Diffusivity regime}.
This property characterises the importance of the different non-ideal MHD
terms in the induction equation. The ratios of each of the diffusive terms to the inductive term are typically used, which in a multifluid formulation are expressed as
(e.g. \citealt{BT01}, \citealt{SS02a}):
\begin{equation}
\frac{O}{I} = \frac{\eta \Omega}{v_{\rm A}^2} \equiv \tilde{\eta} \,,
\label{eq:OI}
\end{equation}
\begin{equation}
\frac{H}{I} = \frac{X}{2} \,,
\label{eq:HI}
\end{equation}
and
\begin{equation}
\frac{A}{I} = \frac{\Omega}{\gamma_i \rho_i} \,.
\end{equation}
Note that the inverse of the normalised Ohm resistivity $\tilde{\eta}^{-1}$ in equation (\ref{eq:OI}) has been referred to as the magnetic Reynolds number. However, as discussed
above, this parameter combination is akin to the Elsasser number, with the difference that in the definition of equation (\ref{eq:Lambda}) we used the total
perpendicular diffusivity $\eta_\perp$, instead of $\eta$. In equation (\ref{eq:HI}),
\begin{equation}
X \equiv \frac{c B \Omega}{2 \pi e n_e v_{\rm A}^2}
\end{equation}
is the so-called ``Hall parameter", not to be confused with its
namesake, the ratio of the gyrofrequency to the collision frequency of ionized species $j$ (here either ions or electrons) with the neutrals, given by
\begin{equation}
\beta_j \equiv \frac{eB}{m_jc}\frac{1}{\gamma_j \rho} \,.
\end{equation}
Similarly, the relative importance of the
non-ideal MHD terms can be measured by the following ratios (e.g. \citealt{BT01}, \citealt{SS02a})
\begin{equation}
\frac{H}{O} = \frac{eB}{m_e c} \frac{1}{\gamma_e \rho} \equiv \beta_e
= \frac{X}{2 \tilde{\eta}} \,,
\end{equation}
and
\begin{equation}
\frac{A}{H} = \frac{eB}{m_i c} \frac{1}{\gamma_i \rho} \equiv \beta_i \,.
\end{equation}
Using the tensor diffusivity notation, we can write
\begin{equation}
\frac{O+A}{I} = \frac{\eta_\mathrm{P} \Omega}{v_{\rm A}^2} \,,
\label{eq:etaOAI}
\end{equation}
\begin{equation}
\frac{H}{I} = \frac{\eta_\mathrm{H} \Omega}{v_{\rm A}^2} \,,
\label{eq:etaHI}
\end{equation}
and the relative magnitudes of the non-ideal MHD terms are, simply
\begin{equation}
\frac{H}{O} = \frac{\eta_\mathrm{H}}{\eta}
\label{eq:etaHO}
\end{equation}
and
\begin{equation}
\frac{A}{H} = \frac{\eta_\mathrm{A}}{\eta_\mathrm{H}} \,.
\label{eq:etaAH}
\end{equation}
From expressions (\ref{eq:etaOAI}) and (\ref{eq:etaHI}) it is clear that the Elsasser number is a measure of the ratio of the inductive term to the total non-ideal MHD terms in the induction equation.
Furthermore, for the vertical field geometry adopted here, the ambipolar diffusivity acts as a
field-dependent resistivity\footnote{In other words, for this field geometry the current density $\bm{J}$ is perpendicular to $\bm{B}$ and the term $(\bm{J} \bm{\times} \bm{B}) \bm{\times}
\bm{B} \equiv (\bm{J} \cdot \bm{B}) \bm{B} - B^2 \bm{J} = B^2 \bm{J}$ (see equation (\ref{eq:induction_drift})).} \citep[see ][]{BT01}. This property will be used in the next section to treat the Ohm and
ambipolar diffusivities as an `Ohm-like' term when comparing the impact of the `Ohm-like' and Hall resistivities on the MRI.
\end{enumerate}
\subsection{Criterion for Hall diffusion to affect the MRI}
\begin{figure}
\centering
\includegraphics[scale=0.8]{fig9.pdf}
\caption{Regions of parameter space where the MRI is expected to grow, and dominant diffusion mechanism, in a (O+A)/I versus H/I (or, equivalently, the
normalized Pedersen versus Hall diffusivities) plane. Note that for the vertical field geometry considered here, the ambipolar and Ohm diffusivities can be combined
into a generalized `Ohm-type' diffusivity (see text). In the lower-left (ochre) panel, magnetic diffusion is weak and the instability grows at a rate comparable to the ideal-MHD rate ($\sim
\Omega$). Conversely, in the upper-left (grey) and lower-right (purple) panels, only one diffusivity term (Ohm-type and Hall, respectively) dominate over the
inductive term in the induction equation. Ohm-type diffusivity suppresses the instability in the top-left quadrant (see footnote \ref{foot:bound}, however), but
since Hall diffusion can also be destabilising (Section \ref{sec:diffusion}) the instability can still grow in the lower-right panel. Finally, both diffusivity
components are important in the upper-right quadrant of the figure. In the top (grey) portion of this panel, Ohm-type diffusion overcomes Hall diffusion [H/(O+A) $<
1$] and the MRI is expected to be suppressed. Hall diffusivity, however, is dominant [H/(O+A) $> 1$] in the bottom (purple) portion of the panel. In this region of
parameter space, H/(O+A) \emph{and} H/I are both $> 1$ and the instability may proceed despite $\Lambda$ being less than unity. }
\label{fig:hall_dom}
\end{figure}
\begin{figure}
\centering
\includegraphics[scale=1.0]{fig10.pdf}
\caption{Growth rate of the fastest-growing MRI-unstable mode as a function of the ratio of the Hall to Ohm diffusion terms in the induction equation [H/O $= X/
(2 \tilde{\eta})$]. Calculations correspond to a stratified disc with $\beta \equiv (2/\gamma) c_{\rm s}^2/v_{\rm A}^2 = 3200$ (or $v_{\rm A}/c_{\rm s} = 0.02$).
Each curve corresponds to a different value of $\tilde{\eta}^{-1}$ (= I/O):
0.1 (blue), 1.0 (red) and 10 (green). A solid (dotted) line is used to plot solutions for which the ratio H/I $> 1$ ($< 1$). Note that solutions for which Hall
diffusion is dominant lie in the solid-line portion of the curves (H/I $> 1$) \emph{and} to the right of the vertical dashed line (H/O $> 1$). Superimposed (filled
circles) are the linear growth rates corresponding to the solutions presented in Fig.~9 of \citet{SS02b} for $X = 4$. Note that the calculation for $\tilde{\eta}^
{-1} = 0.1$ and $X = 4$ satisfies H/O $= 0.2$ and is \emph{not} in the Hall-dominated region of parameter space. As a result, the MRI is significantly damped by Ohm
diffusion. On the contrary, the instability can grow at about the ideal-MHD rate when H/O $> 1$ even when $\tilde{\eta}^{-1} = 0.1$.}
\label{fig:Sano_comp}
\end{figure}
We now use the ratios given in the previous subsection to constrain the region
of parameter space where the MRI is expected to grow, as well as to
determine a criterion for Hall diffusion to substantially modify its properties. First,
however, we discuss the results of the work of \citet{SS02b} on
this topic. These authors characterised the Ohm and Hall
terms by their magnitudes relative to the inductive term in the
induction equation (e.g.~via eqs.~\ref{eq:OI} and \ref{eq:HI}). Furthermore, they
considered that Hall diffusion was dominant when the ratio of the Hall to
inductive terms was larger than unity (H/I $> 1$). Fig.~9 of \citet{SS02b}
shows the saturation level of the Maxwell stress (normalised by the initial
gas pressure), as a function of the initial (subscript `0') of the inverse Ohm resistivity ($\tilde{\eta_{\rm 0}}^{-1}$) -- called the magnetic Reynolds number --
for different values of the initial
plasma beta and Hall parameters ($\beta_{\rm 0}$ and $X_{\rm 0}$,
respectively).
Their results indicate that when $\tilde{\eta_{\rm 0}}^{-1} \geq 1$,
the normalised saturated value of the Maxwell stress is of the order of
$0.1$ (with some scatter depending on the adopted $\beta_{\rm 0}$ and
$X_{\rm 0}$), and it is fairly independent of the actual value of
the resistivity. The actual saturated magnitude of the stress is larger
in the models with $X_{\rm 0} = 4$ with respect to the ones with $X_{\rm 0}
= 0$ (e.g.~no Hall diffusivity) or $X_{\rm 0} = -2$ (negative Hall
diffusivity, when the magnetic field is counter-aligned with the disc
angular velocity vector). On the contrary, when the initial inverse resistivity $\tilde{\eta_{\rm 0}}^{-1}$ is less than
unity, the saturation level of the Maxwell stress decreases by 1-2
orders of magnitude with respect to the $\tilde{\eta_{\rm 0}}^{-1} > 1$ case. This trend of the Maxwell stress at saturation with
$\tilde{\eta_{\rm 0}}^{-1}$ seems to be unaffected by the presence (or magnitude) of
the Hall diffusivity. These results have been interpreted to show that the
Hall diffusivity does not change the critical (maximum) $\tilde{\eta}$
required for the instability to grow ($\tilde{\eta}_{\rm crit} \sim 1$), but it
enhances the saturated level of the Maxwell stress by a factor of a few.
Note that the calculations of \citet{SS02b} probe the region of parameter
space where Ohm and Hall diffusivity terms dominate over the inductive term
(${\rm I}/{\rm O} = \tilde{\eta}_{\rm 0}^{-1} < 1$ and H/I $= X_{\rm 0}/2 > 1$).
However, $|X_{\rm 0}| \leq 4$ in all the presented calculations. As a result,
for $\tilde{\eta}_{\rm 0} > 2$, the ratio H/O $= X_{\rm 0}/(2 \tilde{\eta}_{\rm 0})$ is $< 1$ and the
dominant diffusion mechanism is Ohm-type. For the calculations with
$X_{\rm 0} = 4$ and $\tilde{\eta}_{\rm 0}^{-1} = 0.1$, in particular, the ratio ${\rm H}/{\rm O} = 0.2$, which implies that the Hall term is too weak
to overcome the damping effect introduced by Ohm diffusion. It is not
surprising, therefore, that the large drop in the saturated value of the
Maxwell stress, with respect to that associated with the solutions
satisfying $\tilde{\eta}_{\rm 0}^{-1} > 1$, is not significantly modified by the
presence of Hall diffusion. Naively, and as confirmed by the linear
analysis, the Hall effect should substantially modify the growth rate of
the magnetorotational instability\ provided that Hall diffusion dominates over the inductive term (i.e.\
$|X_{\rm 0}|\ga 2$) \emph{and also dominates Ohm diffusion} (i.e.\ $X_{\rm 0}\ga
2\tilde{\eta}_{\rm 0}$). According to these criteria, the Hall term would be strong enough to
modify the presented results if $X_{\rm 0} \geq 20$ (for the $\tilde{\eta}_{\rm 0}^{-1} =
0.1$ case), a value likely to be well beyond what has been computationally
feasible so far.
The above considerations are summarised in Fig.~\ref{fig:hall_dom}, which
shows the regions of parameter space where the instability is expected to
grow -- and, if so, the dominant diffusion mechanism -- in a [(O+A)/I - H/I]
or, equivalently, the normalised Pedersen and Hall diffusivities, plane. As discussed at the end of Section \ref{subsec:fluid}, we use a generalised `Ohm' term, which should be understood to mean `Ohm + ambipolar'.
Diffusivity effects are weak in the lower-left (ochre) quadrant of the
figure, as the inductive term dominates over both the Ohm and Hall
diffusion terms in this region of parameter space. The MRI is then
expected to grow at a rate not significantly reduced from the ideal-MHD
rate\footnote{Note, of course, that the boundaries between these regions
are not sharp as this simplified diagram might suggest, and the MRI is expected to grow at reduced rate when
$\eta_\mathrm{P} v_{\rm A}^2/\Omega \sim 1$ (e.g.~near the border between the lower and upper-left
panels of the figure). This also applies along the diagonal line
separating the Hall and Ohm-dominated regions in the upper-right
panel.\label{foot:bound}}. Conversely, in the upper-left (grey) and
lower-right (purple) panels, only one diffusivity term is dominant over the
inductive term (the Ohm and Hall term, respectively). When Ohm diffusion
dominates, the instability's growth is damped, as this diffusion mechanism
is always stabilising (See the discussion in Section \ref{sec:diffusion} and Fig.~\ref{fig:MRI-drift-AD}). However, the Hall term can be either stabilising or
destabilising (see Figs.~\ref{fig:MRI-drift-a} and \ref{fig:MRI-drift-b}), so when it dominates the instability can still potentially
grow . Finally, the upper-right
quadrant of the figure contains the region of parameter space where both
the Ohm and Hall diffusivity terms dominate over the inductive term.
Although both diffusivity terms are important with respect to the ideal-MHD
term in this region, only in the lower (purple) portion of the panel Hall
diffusion is strong enough to potentially overcome the damping effect
introduced by Ohm diffusivity, as {\it only} here the ratios H/(O+A) and H/I
are both $> 1$. Note that this region of parameter space was not probed by
\citet{SS02b}, as their solutions incorporating Hall diffusion lie in the
upper (grey) portion of the panel, where Hall diffusion is weak in
comparison to Ohm diffusion. The potentially destabilising effect of a
sufficiently strong Hall diffusivity (the purple region of the upper-right
quadrant in Fig.~\ref{fig:hall_dom}), therefore, remains to be explored by
numerical simulations.
In order to test these assertions, we computed the growth rate of the most
unstable MRI mode in a stratified, geometrically thin and axisymmetric
accretion disc, using the method described in \citet{SW03}, to which we
refer the reader for additional details. Our results are depicted in
Fig.~\ref{fig:Sano_comp} for $\beta = 3200$ (or $v_{\rm A}/c_{\rm s} =
0.02$), as a function of the ratio H/O ($= X/2\tilde{\eta}$). Each curve
corresponds to a different value of the ratio I/O ($= \tilde{\eta}^{-1}$), as
follows: $ \tilde{\eta}^{-1}$ = 0.1 (blue), 1.0 (red), and 10 (green). The range of
solutions shown with solid lines satisfy $X > 2$, so
that the Hall term dominates over the inductive term (H/I $> 1$). The solutions depicted with a dotted line correspond to $X < 2$ (H/I $<
1$). Note also that to the left of the vertical dashed line the ratio H/O
= $X/2\tilde{\eta} < 1$ and Ohm diffusion is the dominant non-ideal MHD term.
From the considerations in the previous paragraphs, it is clear that the
solutions that probe the Hall-dominated region of parameter space should
lie in the solid segment of each curve {\it and} to the right of the
vertical dashed line. Note, in particular, that even for $\tilde{\eta}^{-1} = 0.1$, the instability
grows at essentially the ideal-MHD rate ($\nu_{\rm max}/\Omega \approx
0.75$) if the Hall term is strong enough (H/O $> 1$, or to the
right of the vertical line). For comparison, we superimpose filled circles on the curves
showing the linear growth rate of the most unstable mode
corresponding to $X = 4$ and $\tilde{\eta}^{-1} = 0.1$, 1 and 10 (the set of
parameters tested by \citealt{SS02b}). Note that the solution with $X = 4$
and $\tilde{\eta}^{-1} = 0.1$ lies to the \emph{left} of the H/O $= 1$ line, and Ohm
diffusion suppresses the growth of the instability. We conclude, therefore, that the relative saturation levels of the MRI simulations presented in \citet{SS02b} are entirely consistent with expectations based on a linear analysis, and did not probe the regime $H>\{I,O,A\}$ in which the hall term strongly modifies the linear MRI.
\section{Application to protoplanetary discs}
\label{sec:ppds}
We now illustrate the implications of Hall diffusion for the extent of MRI-driven activity in protoplanetary discs by applying the linear analysis from
appendix \ref{app:analysis} to the minimum mass solar nebula at 1\,au
While a linear analysis cannot hope to give a good sense of the properties of the nonlinear turbulence the MRI drives, numerical simulations in the Ohm limit have shown that it does predict when such turbulence exists, and appears to be an excellent predictor of the vertical extent of dead zones in stratified simulations of protoplanetary discs \citep{TS08}. As noted in the previous section, the Hall-Ohm simulations of \cite{SS02a,SS02b} are also consistent with the expectations based on the linear dispersion relation. Thus, we use the local dispersion relation to examine the role of the magnetic diffusivity in determining the extent of MRI-active regions in protoplanetary discs.
The tendency of the MRI to manifest on successively larger scales in the presence of increasing diffusivity is limited by the finite thickness of the disc: the wave numbers of interest are bounded from below by requiring $kh>1$ where $h = c_s / \Omega$ is the disc scale height. Thus we adopt as a local criterion for growth of the MRI that the dispersion relation (\ref{eq:dispersion_relation}) yields modes with $\nu >0$ and $kh>1$. From this we infer that the diffusivities must lie within the semicircular locus obtained from eq (\ref{eq:disp-circle}) with $\nu=0$ and $k=h^{-1}$, i.e.
\begin{equation}
\left(\frac{s\eta_\mathrm{H}\Omega}{v_A^2} +
\frac{5}{4}-\frac{3}{4}\frac{c_s^2}{v_A^2}\right)^{2} +
\left(\frac{\eta_\mathrm{P}\Omega}{v_A^2}\right)^2 <
\;\frac{9}{16}\left(1+\frac{c_s^2}{v_A^2}\right)^2
\,.
\label{eq:disp-critical}
\end{equation}
where $s=sign(B_z)$.
This criterion applies to any near- Keplerian disc threaded by a vertical magnetic field. Its apparent simplicity belies the fact that the diffusivities are complicated (but calculable) functions of magnetic field strength, gas density, temperature, and the abundances of charged species. Note also that the ratio $c_s^2/v_A^2$ plausibly ranges between 0 and the equipartition value
$2$ at the disc mid-plane but may be significantly higher away from the mid-plane if the field is anchored at lower heights within the disc.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{fig11.pdf}
\caption{Fractional abundances of charged particles in the minimum mass solar nebula at 1\,au from the Sun as a function of height above the mid-plane, assuming that grains have radius 1\,\micron. The upper, centre, and lower panels correspond to dust to gas ratios $10^{-2}$, $10^{-4}$, and 0 by mass, respectively. The black curve in the lower panel shows the height-dependence of the ionisation rate per hydrogen nucleus due to stellar x-rays and interstellar cosmic rays (see text) assumed for all three models. The other curves give the fractional abundances of electrons (\emph{red}), light ions (H$^{+}$, H$_3^{+}$, He$^{+}$, C$^{+}$) representative molecular (m$^{+}$), and metal (M$^{+}$) ions (\emph{blue}), and grains (\emph{green}, labelled by charge state).}
\label{fig:x1au}
\end{figure}
We specialise to protoplanetary discs by adopting the ionisation models of \citet{W07} for a standard minimum mass solar nebula disc ionised by cosmic-rays and stellar x-rays at 1\,au from the central star. These models assume a column density of $1700\,\mathrm{g\,cm^{-2}} $
and temperature $T=280$\,K independent of height, with x-ray ionisation rate computed by \citet{IG99} and a standard interstellar cosmic-ray ionisation rate of $10^{-17}$\,s$^{-1}$\,H$^{-1}$ that is exponentially attenuated with depth as $\exp(-\Sigma / 96\,\mathrm{g\,cm^{-2}} )$. We characterise the grain population by assuming a single 1\,\micron\ radius and varying the dust-to-gas mass ratio, crudely mimicking the effect of the settling of grains to the disc mid-plane.
The abundances of charged species for dust-to-gas mass ratios of $10^{-2}$, $10^{-4}$ and 0 are presented in Figure \ref{fig:x1au}.
The ionisation rate as a function of depth is plotted in the lower panel -- interstellar cosmic rays are the dominant source below 2 scale heights, above this ionisation by stellar x-rays dominates. In the no-grain case, electrons and metal ions are the dominant charged species because the metal ions have the smallest recombination rate coefficient. The top panel shows the effect of adding a population of 1\,\micron-radius grains with total mass 1\% of the gas mass: grains acquire a charge via sticking of electrons and ions from the gas phase. Above $z/h\approx 2.5$ the grain charge is determined by the competitive rates of sticking of ions and electrons, with the Coulomb repulsion of electrons by negatively-charged grains offsetting their greater thermal velocity compared to ions \citep{Spit41, DS87}. This leads to a gaussian grain charge distribution with mean charge (in units of $e$)
$\langle Z_g \rangle \approx -4akT/e^2 \approx-67$ and standard deviation
$\approx \langle Z_g \rangle ^{1/2}\approx 8$.
Most recombinations still occur in the gas phase. The abundance of metal ions and electrons is reduced over the grain-free case for $z/h\la 5$ where the charge stored on grains becomes comparable to the electron abundance. For $z/h\approx 2$--2.5, the abundances of ions and electrons have declined to the point that the majority of electrons stick to grain surfaces before they can recombine in the gas phase, and most neutralisations occur when ions stick to negatively charged grains \citep{NU80}. Closer to the mid-plane, the ionisation fraction is so low that most grains are lightly charged, so that Coulomb attraction or repulsion of ions and electrons by grains is negligible. Then ions and electrons stick to any grain that they encounter, with recombinations occurring on grain surfaces. The middle panel of Fig.\ \ref{fig:x1au} shows what happens if 99\% of the grains are removed (e.g.~by settling to the mid-plane). The capacity of the grain population to soak up electrons from the gas phase is reduced a hundredfold, and the height below which grains substantially reduce the free electron density below the ion density moves downwards to the lowest scale height.
Unlike Ohm resistivity, the Hall and ambipolar diffusivities depend on the magnetic field strength as well as the charged particle abundances, so we consider field strengths ranging between $10^{-3}$--$10^2$\,G, encompassing the plausible range of values
in the solar nebula (see \citealp{W07}). For each choice of magnetic field we use the ionisation calculations to compute the diffusivities as a function of height. Typically, ambipolar diffusion dominates at the surface and, for weak magnetic fields, Ohm diffusion dominates near the mid-plane. Between these regimes Hall diffusion dominates \citep{W07}. With the diffusivity profiles in hand, we apply eq (\ref{eq:disp-critical}) to identify the fastest growing MRI mode that has $kh>1$ at a given height; the results for dust-to-gas mass ratios $0$ and $10^{-2}$ are displayed in Figs~\ref{fig:maxnu-f=0} and \ref{fig:maxnu-f=1} respectively.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{fig12}
\caption{Colour shading shows the growth rate of the fastest growing local MRI mode at height $z$ above the mid-plane in the minimum mass solar nebula at 1\,au as a function of the strength of an initially vertical magnetic field $B$. Dust grains are assumed to have settled to the mid-plane (see lower panel of Fig.\ \ref{fig:x1au}). The vertical wave number $k$ of the modes are required to satisfy $kh>1$ where $h$ is the disc scale height. The unshaded regions show stable combinations of height and field strength. The top and middle panels correspond to the cases where the initial field and rotation axis are parallel ($B_z>0$) or antiparallel ($B_z<0$), respectively. The lower panel shows the effect of artificially suppressing ambipolar and Hall diffusion, so including only Ohm diffusion. In this case there is no dependence on the sign of $B_z$. The dashed line in each panel indicates the value of the local equipartition magnetic field as a function of height above the mid-plane.\label{fig:maxnu-f=0}}
\end{figure}
First consider the case in which grains are absent (Fig.~\ref{fig:maxnu-f=0}). The bottom panel shows the fastest growth rate (subject to
$kh>1$) as a function of magnetic field strength and height if Hall and ambipolar diffusion are neglected and only Ohm diffusion is included. In this case the departure from ideal MHD is measured by $\eta\Omega/v_A^2$ (the inverse of the Elsasser number; see section \ref{subsec:fluid}), which strongly decreases with height as the fractional ionisation and Alfv\'en speed both increase
sharply away from the midplane. For the ionisation profile plotted in Fig.\ \ref{fig:x1au} it turns out that $\eta\Omega/v_A^2$ is small near the surface (so that ideal MHD holds) and large near the mid-plane (so that Ohm damping is severe). As a result, near the surface the largest achievable growth rate is close to the ideal value 0.75\,$\Omega$, but declines rapidly below the height where $\eta\Omega/v_A^2 \sim 1$. The height of the transition between these regimes declines with increasing field strength, simply because
$\eta\Omega/v_A^2\propto B^{-2}$. A second consideration is that the range of MRI-unstable wave numbers is bounded above by $k_c$ (see eq \ref{eq:kc}), and this must be larger than $1/h$ if \emph{any} unstable modes are to exist with $kh>1$. It is this criterion that provides the upper and lower envelopes to the unstable region in Figs.\ \ref{fig:maxnu-f=0} and \ref{fig:maxnu-f=1}. Near the surface where
$\eta\Omega/v_A^2 \ll 1$, $k_c$ is approximately $\sqrt{3}\Omega/v_A$, and as $v_A$ rapidly increases with height the unstable range shifts to wave numbers with $kh<1$ and it is no longer possible to find unstable modes that fit within a scale height. This occurs when the magnetic pressure roughly exceeds the gas pressure. By contrast, close to the mid-plane where $\eta\Omega/v_A^2 \gg 1$, $k_c \approx \sqrt{3}v_A/\eta$ and the wave numbers are pushed out of the relevant range by the increasing diffusivity and the declining Alfv\'en speed as the mid-plane is approached.
Next, we add the remaining magnetic diffusion terms, ie.\ ambipolar and Hall diffusion. The latter depends on the sign of $B_z$, which we take to be positive or negative in the upper and middle panels of Fig.\ \ref{fig:maxnu-f=0} respectively. Ambipolar diffusion dominates in the surface layers (see Fig.\ 5 of \citealt{W07}), and it acts just like Ohm diffusion when, as assumed here, the initial magnetic field is vertical
\footnote{This is not the case in more general geometries, e.g.~ambipolar diffusion may be destabilising when the field has vertical and toroidal components and the wave number has radial and vertical components (see \citealt{KB04} and \citealt{D04}).}, tending to damp the growth of the MRI and pushing the unstable
modes to longer wavelengths. Hall diffusion tends to dominate the lower layers, except for weak fields when Ohm diffusion dominates closer to the mid-plane.
For $B_z > 0$ Hall diffusion is destabilising and pushes the region where the fastest growth rate is close to 0.75\,$\Omega$ to greater depths. The rapid decline to low growth rates (ie.\ from green to blue shading in Fig. \ref{fig:maxnu-f=0}) traces the transition from Hall-dominated to Ohm-dominated diffusion. Below this, Hall diffusion acts to significantly extend the region of slow growth by modifying $k_c$, and this extends all the way to the mid-plane for fields in excess of 10\,mG.
On the other hand, when $B_z<0$ (ie.\ $\bm{B}$ is antiparallel to the rotation axis) Hall diffusion tends to stabilise the MRI while extending the unstable wavelengths to shorter wavelengths (see Fig.\ \ref{fig:contours}). As a result, super-equipartition fields are unstable near the mid-plane. We emphasise that \emph{irrespective of the magnitudes of the other diffusivities Hall diffusion is completely stabilising when $\eta_\mathrm{H} > 2 v_A^2/\Omega$.}This is responsible for the sharp cutoff at the lower boundary in the middle panel of Fig.\ \ref{fig:maxnu-f=0}.
When a full complement of 1\,\micron\ dust grains are present (i.e.~dust-to-gas mass ratio $10^{-2}$; charged species as in lower panel of Fig.\ \ref{fig:x1au}), the diffusivities are greatly increased near the mid-plane because electrons are locked up by grains and rendered immobile. Fig.~\ref{fig:maxnu-f=1} displays the same trends as in the zero-grain case, but the MRI-unstable region is now restricted to the upper layers of the disc, and in all cases the bulk of the disc is stable to the MRI. While the differences between the three panels might appear less severe in this case, the strong density stratification means that there are orders of magnitude differences in the column density of the MRI-unstable region.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{fig13}
\caption{As for Fig.\ \ref{fig:maxnu-f=0}, but now including a
population of 1$\,\mu$m radius grains with total mass 1\% of the gas mass
(see upper panel of Fig.\ \ref{fig:x1au}).\label{fig:maxnu-f=1}}
\end{figure}
Fig.~\ref{fig:active-sigma} shows the magnetically-active column density as a function of field strength for different assumptions regarding the diffusivity when 99\% of grains are assumed to have settled (i.e.\ dust to gas ratio $10^{-4}$).
\begin{figure}
\centering
\includegraphics[width=8.5cm]{fig14.pdf}
\caption{Column density of the MRI-active region at 1\,au in the minimum mass solar nebula as a function of magnetic field strength for different assumptions about magnetic diffusion (see text). Grains are assumed to have radius 1\micron\ with 99\% having settled to the disc mid plane (i.e.~total mass only $10^{-4}$ relative to gas). \emph{Solid black} and \emph{dashed blue} curves are for the magnetic field oriented parallel or antiparallel to the disc's rotation axis respectively. \emph{Long-dashed red} curves indicate the effect of accounting only for Ohm or Ohm+ambipolar (i.e.~Pedersen) diffusion. \label{fig:active-sigma}}
\end{figure}
The active column density varies by 1-2 orders of magnitude depending on the accuracy of the treatment of magnetic diffusion. First, neglect Hall diffusion and consider either Ohm diffusion alone or Ohm and ambipolar diffusion (i.e.~Pedersen diffusion) operating in concert (red long-dashed curves). Ohm diffusion dominates ambipolar diffusion except for strong magnetic fields and low densities, so there is little difference between these two cases except for magnetic fields in excess of 1\,G where the additional damping due to ambipolar diffusivity suppresses the MRI. Hall diffusion either increases or decreases the active column density by an order of magnitude depending on whether the initial magnetic field is pointing up or down; the maximum unstable field strength also varies considerably.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{fig15.pdf}
\caption{
As for Fig. \ref{fig:active-sigma} but for dust-to-gas mass ratio varying from $0$ to $10^{-2}$. Note that active column density is capped at the total column of the minimum mass solar nebula at 1\,au, i.e.~1700\,g\,cm$^{-2}$.\label{fig:active-sigma-dust}}
\end{figure}
Fig.~\ref{fig:active-sigma-dust} shows how the active column depends on grain abundance and magnetic field orientation. In the absence of grains the entire disc cross-section is magnetically active for an upwardly-directed field in the range 0.01--5\,G. On the other hand, when the field is pointing downwards, a minuscule fraction of the disc is active for weak fields, but the active column increases rapidly and encompasses the entire
cross section of the disc
for field strengths in the range 20--80\,G. As one would expect, grains sharply reduce the active column density because of the reduction in mobile charge carriers. The continued extreme sensitivity to field alignment demonstrates that Hall diffusion still plays a critical role in determining the extent of magnetic activity regardless of the grain abundance.
The reduction in the active column when $B_z < 0$ occurs because Hall diffusion stabilises the disc against the MRI when $s\eta_\mathrm{H}<-2 v_A^2/\Omega$.
When the grains and ions are strongly coupled to the neutrals and the electrons are coupled to the magnetic field, the Hall diffusivity is given by eq (\ref{eq:etaH}) and we obtain a field-dependent criterion for stability on the fractional ionisation
\begin{equation}
\frac{n_e}{n_H} \, \la \;\frac{1}{2} \, \frac{1.4 m_H c \, \Omega}{eB} \approx \frac{1.5\times 10^{-11}}{B(\mathrm{G})} \,\frac{(M/\mathrm{M_\odot})^{1/2}}{ (r/\mathrm{au})^{3/2}}\,.
\label{eqn:xe-crit}
\end{equation}
This critical fractional ionisation is typically larger than the corresponding criterion for Ohm damping, so sets the lower boundary of the magnetically active region when $B_z < 0$.
Note that once the grain abundance is low enough not to affect the ionisation fraction, the lower boundary reaches the midplane and the active column density saturates.
These results suggest that Ohm estimates of the column density of the magnetically-active layers in protoplanetary discs are in error by about an order of magnitude, systematically under- or overestimating the active column if the magnetic field is directed upwards or downwards, respectively.
This conclusion rest on several assumptions. Our neglect of stratification in the linear analysis is unlikely to be serious given that in the Ohm limit this approach yields excellent predictions of the extent of the depth of the turbulent layers \citep{TS08}. The restriction to a simple and somewhat degenerate geometry -- i.e~ to vertical fields and wave vectors -- is perhaps more suspect. However, when ambipolar diffusion is unimportant, a toroidal component of the magnetic field and a radial component of the wave vector can easily be accommodated: a rescaling of the full dispersion relation yields the identical dispersion relation (Pandey \& Wardle, in preparation). The situation is less clear when ambipolar diffusion is involved, as in this case it stabilises or destabilises the MRI
\citep{KB04,D04}
in an analogous manner to the Hall contribution for the vertical field considered here (Pandey \& Wardle, in preparation). However, at $\sim1$\,au, ambipolar diffusion only dominates near the disc surface where the ionisation fraction is so high that diffusion is small in any case. Finally, although the grain model is crude, the key property of the grain population is its capacity to soak up electrons, which is proportional to $\int a\,n(a)\,da$ \citep{W07}.
This can be used to scale our results to any grain size distribution. The real uncertainty is the small-radius tail of the grain size distribution which may still be sufficient to reduce the electron and ion fractions at a few scale heights despite the tendency of grains to aggregate and settle towards the mid-plane.
The robust point is that the relative magnitudes of the diffusivities are independent of the ionisation fraction, being controlled by the ratio $B/n_\mathrm{H}$, and that Hall diffusion dominates the other mechanisms over much of protoplanetary discs \citep{WN99,BT01,SS02a}.
As X-rays and cosmic rays both penetrate to the Hall-dominated region just below the disc surface, it is absolutely essential that Hall diffusion be included in modelling the extent of MRI-driven turbulence in protoplanetary discs. The real uncertainty is that, notwithstanding the pioneering effort of \citet{SS02a,SS02b}, the saturation of MRI-driven turbulence in the Hall-dominated regime
as yet remains unexplored.
\section{Summary \& Conclusions}
\label{sec:summary}
In this paper we re-examined the role of Hall diffusion in suppressing or enhancing MRI-driven magnetic turbulence in Keplerian discs.
We first undertook a local, linear analysis of the magnetorotational instability, for simplicity restricting our attention to a vertical, weak magnetic field subject to axisymmetric perturbations with a purely vertical wave vector. While this is not new, our approach and presentation differed from previous analyses \citep{W99,BT01} in two critical ways. First, we characterised the departure from ideal MHD using magnetic diffusivities rather than conductivities or characteristic frequencies, allowing us to make a clearer connection with the $\E\bm{\times}\bm{B}$ field-line drift implicit in the diffusive MHD induction equation. Second, we presented a clear perspective of the dependence on the Hall and Pedersen (Ohm+ambipolar) diffusivities by examining how the properties of the instability vary over a Hall-Pedersen diffusivity plane (see Fig.~\ref{fig:locii}). This clearly delineates when the range of unstable wave numbers is finite or infinite, and whether the fastest growth occurs for finite wave number or asymptotically as $k\rightarrow\infty$. We also discussed the limiting forms of the dispersion relation, making a connection with the diffusive plane-parallel shear instabilities of \cite{K08}.
Next, we reviewed the alternative parameterisations of non-ideal MHD that have appeared in the literature, and emphasised that existing simulations of the nonlinear development and saturation of the instability in the Hall-Ohm case \citep{SS02a,SS02b} are consistent with expectations based on the simple linear analysis and have not yet probed the Hall-dominated regime characteristic of protoplanetary discs.
Finally, we illustrated the critical effect of Hall diffusion on the size of dead zones in protoplanetary disc s by applying a local criterion for growth of the magnetorotational instability \ to a simple model of minimum-mass solar nebula at 1\,au , including x-ray and cosmic-ray ionisation and a population of 1\,\micron\ grains.
Our key results can be summarised as follows.
\begin{enumerate}
\item The radial diffusion of perturbed field lines through the fluid is directly related to the criterion for marginal stability. The MRI is suppressed if the radial drift of field lines against the infall of the fluid is sufficient to restore the field to its equilibrium radial position.
\item The behaviour of the MRI for our adopted geometry is determined by the dimensionless Pedersen and Hall diffusivities $\eta_\mathrm{P} \Omega/v_A^2$ and $s\eta_\mathrm{H}\Omega/v_A^2$, where $s = \mathrm{sign}(B_z)$. For $s\eta_\mathrm{H}<-2 v_A^2/\Omega$ Hall diffusion suppresses the MRI irrespective of the value of $\eta_\mathrm{P}$. For $s\eta_\mathrm{H}>-2 v_A^2/\Omega$ the $\eta_\mathrm{P}-s\eta_\mathrm{H}$ half plane is divided into three regions in which either (I) there is a finite range of unstable wave numbers with a fastest growing mode; (II) instability for all wave numbers, with a unique fastest growing mode and a slower asymptotic growth rate as $k\rightarrow\infty$, or (III) instability at all wave numbers with fastest growth asymptotically achieved as $k\rightarrow\infty$ (see Fig.~\ref{fig:locii}).
\item For fixed $\eta_\mathrm{P}$, the maximum growth rate increases with increasing $s\eta_\mathrm{H}$, from 0 at $s\eta_\mathrm{H}\Omega/v_A^2 = -2$ to the maximum 0.75\,$\Omega$ as $s\eta_\mathrm{H} \rightarrow +\infty$. For fixed $\eta_\mathrm{H}$ with $s\eta_\mathrm{H}>-2 v_A^2 \Omega$, the growth rate is a maximum for $\eta_\mathrm{P}=0$ and declines as $\eta_\mathrm{P}\rightarrow\infty$ (see Fig.~\ref{fig:contours}).
\item In the highly diffusive limit the instability reduces to the Hall-diffusion-driven instability in plane-parallel shear flow discussed by \cite{K08}. Diffusion is so severe in this limit that the perturbations in the fluid velocity do not affect the field evolution, which is driven purely by diffusion and Keplerian shear. Our restriction to vertical initial fields and perturbation wave numbers enabled us to extend the results of \cite{K08} to include the damping by Pedersen diffusion and to derive pleasant analytic expressions for the upper cutoff to the unstable wavenumber range and the growth rate and wave number of the most unstable mode (see eqs \ref{eq:kc-shear}, \ref{eq:nu0-shear}, and \ref{eq:k0-shear}, respectively, and also Fig.~\ref{fig:nu-k-diffusion}).
\item We argued that simulations of MRI-driven MHD turbulence in the presence of Hall and Ohm diffusion \citep{SS02a,SS02b} have not yet probed the ``deep'' Hall regime $s\eta_\mathrm{H} \ga \eta_\mathrm{P} \ga v_A^2/\Omega$, where the linear analysis suggests that Hall diffusion allows the instability to proceed when it otherwise would not.
\item We found that at 1\,au in the minimum-mass solar nebula, Hall diffusion changes the magnetically active column density by an order of magnitude. This change is either an increase or decrease depending on whether B is parallel or antiparallel to the rotation axis, respectively. Hall diffusion likely plays a critical role in determining the radial extent of dead zones and the thickness of magnetically active layers in protoplanetary discs, and estimates based on damping by Ohm diffusion are probably wildly inaccurate.
\end{enumerate}
The simplifications adopted in our analysis engender three significant uncertainties in our conclusions that are worth some final discussion.
First, the restriction to the stability of vertical magnetic fields to perturbations with vertical wave vectors does not capture the destabilisation by ambipolar diffusion that arises when toroidal field and radial wave vector components are also present \citep{KB04,D04}. However, we have captured the analogous Hall diffusion-driven destabilisation of the MRI, and as Hall diffusion typically dominates this restriction is unlikely to have a great impact.
Second, while the use of the linear analysis to predict the boundary of the manifestly nonlinear active region, appears to be justified for the Ohm case \citep{TS08}, it is not known whether this applies in the Hall-dominated regime that we tout here.
Finally, we caution that the MRI may simply be irrelevant in protoplanetary discs. A minor population of small dust grains would remove so many electrons from the gas phase that the MRI-active column density becomes so small as to be irrelevant. Instead, magnetic activity -- if any -- may be due to fields lying within the disc and varying on length scales of order $r$ \citep{TS08} or a strong poloidal magnetic field brought in during the formation of the disc.
\section{Acknowledgements}
\label{sec:acknowledgments}
We are indebted to Catherine Braiding, Matthew Kunz, BP Pandey, Takayoshi
Sano, and Jim Stone for stimulating discussion and suggestions, and are
grateful for the hospitality provided by the Isaac Newton Institute for
Mathematical Sciences at Cambridge University and the School of Physics at
the University of Sydney, where some of this work was conducted. This
research was supported by the Australian Research Council through Discovery
Project grant DP0881066, a Visiting Fellowship from the Isaac Newton Institute at the University of Cambridge, and an Outside Studies
Program grant from Macquarie University.
\bibliographystyle{mn2e}
|
\section{Introduction}
Throughout this paper a {\em topological dynamical system}
(TDS for short) is a pair $(X, T)$,
where $X$ is a non-vacuous compact metric space with a metric $d$ and
$T$ is a continuous map from $X$ to itself.
A non-vacuous closed invariant subset $Y \subset X$ (\ie, $TY\subset Y$)
defines naturally a {\em subsystem} $(Y, T)$ of $(X, T)$.
\medskip
Let $\mathbb Z$, $\mathbb Z_+$ and $\mathbb N$ denote the sets of the integers,
the non-negative integers, and the positive integers, respectively.
\medskip
Let $(X,T)$ be a TDS, for two opene (standing for non-empty open) subsets $U,V$ of $X$, set
\[N(U,V)=\{n\in\mathbb Z_+: T^nU\cap V\neq\emptyset\}=
\{n\in\mathbb Z_+: U\cap T^{-n}V\neq\emptyset\}.\]
We call $N(U,V)$ {the \em hitting time set of $U$ and $V$}.
Recall that a system $(X,T)$ is called {\em topologically transitive}
(or just {\em transitive}) if for every two opene subsets $U,V$ of $X$
the hitting time set $N(U,V)$ is infinite.
A system $(X,T)$ is called {\em weakly mixing}
if the product system $(X\times X, T\times T)$ is transitive;
{\em strongly mixing} if for every two opene subsets $U,V$ of $X$,
the hitting time set $N(U,V)$ is cofinite, \ie, there exists some $N\in \mathbb N$
such that $N(U,V)\supset \{N, N+1,\ldots\}$.
\medskip
A system $(X,T)$ is called {\em minimal} if it contains no proper subsystem.
Each point belonging to some minimal subsystem of $(X,T)$ is called a {\em minimal point}.
Let $(X,T)$ be a transitive system,
$(X,T)$ is called a {\em P-system} if it has dense periodic points;
an {\em M-system} if it has dense minimal points;
an {\em E-system} if it has an invariant measure with full support;
a {\em topologically ergodic system} if for every two opene subsets $U,V$ of $X$,
the hitting time set $N(U,V)$ is syndetic (bounded gaps).
\subsection{Furstenberg families}
Before going on, let us recall some notations related to a family (for more details see \cite{A97}).
For the set of nonnegative integers $\mathbb{Z}_+$,
denote by $\mathcal{P}=\mathcal{P}(\mathbb{Z}_+)$ the collection of all subsets of $\mathbb{Z}_+$.
A subset $\mathcal{F}$ of $\mathcal{P}$ is called a {\em Furstenberg family} (or just {\em family}),
if it is hereditary upward, \ie, $F_1\subset F_2$ and $F_1\in\mathcal F$ imply $F_2\in \mathcal F$.
A family $\mathcal F$ is called {\em proper} if it is a nonempty proper subset of $\mathcal P$,
\ie, neither empty nor all of $\mathcal P$. Any nonempty collection $\mathcal A$ of
subsets of $\mathbb Z_+$ naturally generates a family
\[\mathcal F(\mathcal A)=\{F\subset \mathbb Z_+:\, F\supset A \text{ for some } A \in\mathcal A\}.\]
For a family $\mathcal F$, the {\em dual family} of $\mathcal F$, denoted by $\kappa\mathcal F$, is
\[\{F\in \mathcal P: F\cap F'\neq \emptyset, \forall F'\in \mathcal F\}.\]
Sometimes the dual family $\kappa\mathcal F$ is also denoted by $\mathcal F^*$.
Let $\mathcal F_{inf}$ be the family of all infinite subsets of $\mathbb Z_+$.
It is easy to see that its dual family $\kappa\mathcal F_{inf}$
is the family of all cofinite subsets, denoted by $\mathcal F_{cf}$.
{\bf All the families considered in this paper are assumed
to be proper and contained in $\mathcal F_{inf}$.}
Let $F$ be a subset of $\mathbb Z_+$, the {\em upper Banach density of $F$} is
$$BD^*(F)=\limsup_{|I|\to\infty}\frac{|F\cap I|}{|I|}$$
where $I$ is taken over all nonempty finite intervals of $\mathbb Z_+$
and $|\cdot|$ denote the cardinality of the set.
Denote by $\F_{pubd}$ the family of sets with
positive upper Banach density.
A subset $F$ of $\Z$ is called {\em thick} if it contains arbitrarily long runs of positive integers,
\ie, for every $n\in\N$ there exists some $a_n\in\Z$ such that $[a_n, a_n+n]\subset F$;
{\em syndetic} if there is $N\in\mathbb N$ such that $[n,n+N]\cap F\neq \emptyset$ for every $n\in\Z$;
{\em piecewise syndetic} if it is the intersection of a thick set and a syndetic set.
The families of all thick sets, syndetic sets and piecewise syndetic sets
are denoted by $\mathcal F_{t}$, $\mathcal F_{s}$ and $\mathcal F_{ps}$, respectively.
It is easy to see that $\kappa \mathcal F_s=\mathcal F_t$.
For a sequence $\{p_i\}_{i=1}^{\infty}$ in $\N$, define the finite sums of $\{p_i\}_{i=1}^{\infty}$ as
$$FS\{p_i\}_{i=1}^{\infty}=\left\{ \sum\nolimits_{i\in \alpha}p_i:\,
\alpha \textrm{ is a nonempty finite subset of }\N\,\right\}.$$
A subset $F$ of $\mathbb Z_+$ is called an {\em IP set}
if there exists a sequence $\{p_i\}_{i=1}^{\infty}$ in $\N$
such that $ FS\{p_i\}_{i=1}^{\infty}\subset F$.
We denote by $\mathcal F_{ip}$ the family of all IP sets.
For a family $\mathcal F$, the {\em block family of $\F$}, denoted by $b\F$, is the
family consisting of sets $F\subset\Z$ for which there exists some $F'\in \F$ such that for
every finite subset $W$ of $F'$ one has $m+W \subset F$ for some $m\in \Z$ (see \cite{HLY09}).
It is easy to see that
$b\F=\allowbreak\{F\subset\Z:\ \exists F'\in\F, (\forall n\in\N)(\exists a_n\in\Z)
\textrm{ s.t.\@ } a_n +(F' \cap [0, n])\subset F\}$,
$b(b\F)=b\F$, $b\F_{cf}=\F_t$, $b\F_{s}=\F_{ps}$ and $b\F_{pubd}=\F_{pubd}$.
For a subset $F$ of $\Z$, let $F-F$ be $\{a-b: a,b\in F$ and $a>b \}$ if $0\not\in F$,
or $\{a-b: a,b\in F$ and $a\geq b \}$ if $0\in F$.
For a family $\mathcal F$, define {\em the difference family of $\F$} as
\[\Delta(\F)=\{F\subset\Z: \exists F'\in\F, \textrm{ s.t. } F'-F'\subset F\}\]
and $\Delta^*(\F)=\kappa\Delta(\F)$.
\begin{lem}
Let $\F$ be a family, then $\Delta(\F)=\Delta(b\F)$.
\end{lem}
\begin{proof}
Since $\F\subset b\F$, $\Delta(\F)\subset \Delta(b\F)$.
If $F\in \Delta(b\F)$, there exists $F_1\in b\F$ such that $F_1-F_1\subset F$.
Since $F_1\in b\F$, there exists $F_2\in \F$ and $\{a_n\}_{n=1}^\infty $ in $\Z$ such that
$a_n+F_2\cap[0,n]\subset F_1$. Then $F_2-F_2\subset F_1-F_1$ and $F\in \Delta(\F$).
Hence, $\Delta(\F)=\Delta(b\F)$.
\end{proof}
\subsection{Three ways to classify transitive systems}
It is well known that the study of transitive systems and its
classification play a big role in topological dynamics.
There are several ways to classify transitive systems.
The first one was started with Furstenberg
by the hitting time sets of two opene subsets.
Let $\F$ be a family,
we call $(X,T)$ is {\em $\F$-transitive} if for every two opene subsets
$U, V$ of $X$ the hitting time set $N(U,V)\in \F$.
In his seminal paper \cite{F67}, Furstenberg showed
that a TDS $(X,T)$ is weakly mixing if and only if it is $\{\textrm{thick sets}\}$-transitive.
The second way is by weak disjointness.
Two TDSs are called {\em weakly disjoint} if their product system is transitive \cite{P72}.
Then a system is weakly mixing if and only if it is weakly disjoint from itself.
The third way is by the complexity of open covers
which was introduced in \cite{BHM00}.
For a TDS $(X, T)$ and a finite open cover $\alpha$ of $X$,
let $N(\alpha)$ denote the number of sets in a finite subcover of $\alpha$ with small cardinality.
For an infinite set $A=\{a_1<a_2<\cdots \} \subset \Z$ put
\[ c_A(\alpha,n)= N(T^{-a_1}\alpha\vee T^{-a_2}\alpha\vee \cdots\vee T^{-a_n}\alpha).\]
We call $c_A(\alpha,n)$ the {\em complexity function} of the cover $\alpha$ along $A$.
An open cover $\alpha =\{U_1,U_2,\cdots,U_k\}$ of $X$ is called {\em non-trivial}
if $U_i$ is not dense in $X$ for each $1\le i \le k$.
Let $\F$ be a family,
we call $(X,T)$ is {\em $\F$-scattering} if for every $A\in \F$ and non-trivial open
cover $\alpha$ of $X$ the complexity function $c_A(\alpha,n)$ is unbounded.
Recently, the authors in \cite{A97, AG01,BHM00, G04, HY02, HY04, SY04}
have successfully classified transitive systems by the above three ways.
We summarize the results in the following table:
\begin{longtable}{|c|c|c|c|}
\hline
Transitive properties & $\F$-transitivity & $\F$-scattering & \multicolumn{1}{c|}{Weakly disjointness}\\
\hline
\multirow{2}{*}{Transitivity} & \multirow{2}{*}{$\F_{inf}$} & \multirow{2}{*}{No?}& Mild mixing \\
&&&or trivial system\\
\hline
Total transitivity & $\kappa\F_{rs}$ & No? & Periodic system \\
\hline
\multirow{2}{*}{Weakly scattering} & \multirow{2}{*}{?} & \multirow{2}{*}{No?} & Minimal equi- \\
&&& continuous system \\
\hline
\multirow{2}{*}{Scattering} & \multirow{2}{*}{$\Delta^*(\F_{ps})$} & \multirow{2}{*}{$\F_{ps}$ or $\{\Z\}$}
&Minimal system \\
&&& or M-system \\
\hline
Strong scattering & $\Delta^*(\F_{pubd})$ & $\F_{pubd}$ or $\F_{pud}$ & E-system \\
\hline
\multirow{2}{*}{Extreme scattering} & \multirow{2}{*}{?} & \multirow{2}{*}{?} & Topologically\\
&&&ergodic system \\
\hline
Weak mixing & $\F_t$ & ? & Itself \\% $\times$
\hline
Mild mixing &$\Delta^*(\F_{ip})$ & $\F_{ip}$ & Transitive system \\
\hline
Full scattering & ? & $\F_{inf}$ & \multicolumn{1}{c|}{No}\\
\hline
Strong mixing & $\F_{cf}$ & No & \multicolumn{1}{c|}{No}\\
\hline
\end{longtable}
We understand this table in the following way.
For example, a system is scattering if and only if it is $\Delta^*(\F_{ps})$-transitive
if and only if it is $\F_{ps}$-scattering if and only if it is $\{\Z\}$-scattering
if and only if it is weakly disjoint from every minimal system
if and only if it is weakly disjoint from every M-system.
\subsection{A new way to classify transitive systems}
In this paper we propose a new way to classify transitive systems.
Before going on, we recall some basic notions.
\medskip
For $x\in X$, denote the orbit of $x$ by $Orb(x,T)=\{x,Tx, T^2x, \ldots\}$.
Let $\omega(x,T)$ be the $\omega$-limit set of $x$, \ie, $\omega(x,T)$ is the limit set of $Orb(x,T)$.
A point $x\in X$ is called a {\em recurrent point} if $x\in\omega(x,T)$; a {\em transitive point}
if $\omega(x,T)=X$. It is easy to see that a system $(X,T)$ is transitive if and only if
the set of all transitive points, denoted by $Trans(X,T)$, is a dense $G_\delta$ subset of $X$ and
$(X,T)$ is minimal if and only if $Trans(X,T)=X$.
\medskip
Let $(X,T)$ be a TDS, for $x\in X$ and an opene subset $U$ of $X$, set
\[N(x,U)=\{n\in\Z:\, T^nx\in U\}. \]
We call $N(x,U)$ {the \em entering time set} of $x$ into $U$.
Then it is easy to see that a point $x\in X$ is a transitive point if and only if
for every opene subset $U$ of $X$ the entering time set $N(x,U)$ is infinite.
This suggests that we can use the entering time set of a point into an opene subset
to characterize transitive points.
Let $\F$ be a family, a point $x\in X$ is called an {\em $\F$-transitive point}
if for every opene subset $U$ of $X$ the entering time set $N(x, U)\in\F$.
A system $(X,T)$ is called {\em $\F$-point transitive} if there exists some $\F$-transitive point.
In this paper, we aim to classify transitive systems by $\F$-point transitivity.
We summarize our results in the following table:
\begin{table}[!hbt]
\begin{tabular}{|c|c||c|c|}
\hline
\multirow{2}{*}{Transitive properties} & $\F$-point & \multirow{2}{*}{Transitive properties} &$\F$-point \\
&transitivity& &transitivity\\
\hline
Transitive system & $b\F_{ip}$ & Weakly mixing system & ?\\ \hline
E-system & $\F_{pubd}$ & Weakly mixing E-system & $\F_D$ \\ \hline
M-system & $\F_{ps}$ & Weakly mixing M-system & $\F_{cen}$\\ \hline
Transitive system with & \multirow{2}{*}{$b\F_{wt}$}
& \multirow{2}{*}{HY-system} & \multirow{2}{*}{$\F_{wt}$}\\
dense small periodic sets&&& \\ \hline
\end{tabular}
\end{table}
This paper is organized as follows.
In Section 2, we discuss the connection between families and topological dynamics and
show some basic properties of $\F$-point transitivity.
In Sections 3 and 4, we prove the main results which are declared in the above table.
It is shown that every weakly mixing system is $\F_{ip}$-point transitive, while
we construct an $\F_{ip}$-point transitive system which is not weakly mixing.
In the final section as applications, we discuss disjointness and weak disjointness.
We show that every transitive system with dense small periodic sets is disjoint from
every totally minimal system and a system is $\Delta^*(\F_{wt})$-transitive if and only if it is
weakly disjoint from every P-system.
\subsection*{Acknowledgement}
This work was supported in part by
the National Natural Science Foundation of China (Nos.\@ 11001071, 11071231, 10871186).
The author wish to thank Prof.\@ Xiangdong Ye
for the careful reading and helpful suggestions.
The author also thanks the referee for his/her helpful
suggestions concerning this paper.
\section{Preliminary}
The idea of using families to
describe dynamical properties goes back at least to Gottschalk and Hedlund \cite{GH55}.
It was developed further by Furstenberg \cite{F81}.
For a systematic study and recent results, see \cite{A97}, \cite{G04}, \cite{HY04} and \cite{HY05}.
Let $(X,T)$ be a TDS and $\F$ be a family,
a point $x\in X$ is called an {\em $\F$-recurrent point}
if for every neighborhood $U$ of $x$ the entering time set $N(x, U)\in\F$.
Denote the set of all $\F$-recurrent points by $Rec_{\F}(X,T)$.
It is well known that the following lemmas hold
(see, e.g., \cite[Theorem 1.15, Theorem 1.17, Theorem 2.17]{F81})
\begin{lem} Let $(X, T)$ be a TDS and $x\in X$. Then
\begin{enumerate}
\item $x$ is a minimal point if and only if it is an $\mathcal F_{s}$-recurrent point.
\item $x$ is a recurrent point if and only if it is an $\mathcal F_{ip}$-recurrent point.
\end{enumerate}
\end{lem}
Recall that a TDS $(X,T)$ is called {\em $\mathcal F$-transitive}
if for every two opene subsets $U,V$ of $X$ the hitting time set $N(U,V) \in\mathcal F$;
{\em $\mathcal F$-mixing} if $(X\times X, T\times T)$ is $\mathcal F$-transitive.
\begin{lem}[\cite{F67,A97}] Let $(X, T)$ be a TDS and $\mathcal F$ be a family. Then
\begin{enumerate}
\item $(X, T )$ is weakly mixing if and only if it is $\mathcal F_{t}$-transitive.
\item $(X, T )$ is strongly mixing if and only if it is $\mathcal F_{cf}$-transitive.
\item $(X,T)$ is $\mathcal F$-mixing if and only if it is $\mathcal F$-transitive and weakly mixing.
\end{enumerate}
\end{lem}
Recall that a TDS $(X,T)$ is called {\em $\F$-center}
if for every opene subset $U$ of $X$ the hitting time set $N(U,U)\in\F$.
It is well known that a system $(X,T)$ is transitive if and only if there exists some transitive point.
It is interesting that how to characterize transitive systems by transitive points via a family.
Let $(X,T)$ be a TDS and $\mathcal F$ be a family,
a point $x\in X$ is called an {\em $\mathcal F$-transitive point}
if for every opene subset $U$ of $X$ the entering time set $N(x, U)\in\mathcal F$.
Denote the set of all $\mathcal F$-transitive points by $Trans_{\mathcal F}(X,T)$.
The system $(X,T)$ is called {\em $\F$-point transitive} if there exists some $\F$-transitive point.
Though the terminology ``$\F$-point transitivity'' is first introduced in this paper,
the idea has appeared in several literatures, such as \cite{HPY07,HY05}.
We state their results in our way as
\begin{thm} \label{thm:E-M-system}
Let $(X,T)$ be a TDS. Then
\begin{enumerate}
\item $(X,T)$ is an E-system if and only if it is $\F_{pubd}$-point transitive if and only if
$Trans_{\F_{pubd}}(X,T)=Trans(X,T)\neq\emptyset$ (\cite{HPY07}).
\item $(X,T)$ is an M-system if and only if it is $\F_{ps}$-point transitive if and only if
$Trans_{\F_{ps}}(X,T)=Trans(X,T)\neq\emptyset$ (\cite{HY05}).
\end{enumerate}
\end{thm}
The following remark shows some basic facts about $\F$-point transitivity.
\begin{rem}
(1). It is easy to see that $x\in X$
is an $\mathcal F$-transitive point if and only if for every $F\in \kappa \mathcal F$
one has $\{T^nx: n\in F\}$ is dense in X.
(2). If $x\in X$ is an $\F$-transitive point, then so is $Tx$.
Thus, if $(X,T)$ is $\F$-point transitive then $Trans_{\F}(X,T)$ is dense in $X$.
(3). It should be noticed that $\F$-transitivity may differ greatly from $\F$-point transitivity.
For example, $(X,T)$ is $\F_t$-transitive if and only if it is weakly mixing,
but if $Trans_{\F_t}(X,T)\neq\emptyset$ then $X$ must be a singleton.
\end{rem}
Similarly to $\F$-center, we can define $\F$-point center.
A system $(X,T)$ is called {\em $\F$-point center}
if for every opene subset $U$ of $X$ there exists $x\in U$
such that the entering time set $N(x,U)\in\F$.
\begin{lem}\label{lem:F-point-center}
Let $(X,T)$ be a TDS and $\F$ be a family. If $(X,T)$ is transitive and $\F$-point center, then
$Trans_{b\F}(X,T)=Trans(X,T)$.
\end{lem}
\begin{proof}
Let $x$ be a transitive point and $U$ be an opene subset of $X$.
Since $(X,T)$ is $\F$-point center, there exists $y\in U$ such that $N(y,U)\in \F$.
For every finite subset $W$ of $N(y,U)$, by the continuity of $T$, there exists $m\in\mathbb Z_+$
such that $m+W\subset N(x,U)$. Then $N(x,U)\in b\F$ and $x$ is a $b\F$-transitive point.
\end{proof}
\begin{rem}
(1). Since every recurrent point is $\F_{ip}$-recurrent, every transitive system
is $\F_{ip}$-point center. Thus, a system $(X,T)$ is transitive if and only if
it is $b\F_{ip}$-point transitive.
(2). If $(X,T)$ is $\F$-point transitive, then it is $\F$-point center,
so by Lemma \ref{lem:F-point-center} $Trans_{b\F}(X,T)=Trans(X,T)$.
In particular, if $b\F=\F$, then $(X,T)$ is $\F$-point transitive if and only if
$Trans_{\F}(X,T)=Trans(X,T)\neq\emptyset$.
\end{rem}
\section{Weakly mixing systems}
In this section, we consider weakly mixing systems. We should use the following useful lemma:
\begin{lem}[Ulam \cite{A04}]
Let $X$ be a compact metric space without isolated points and
$R$ be a dense $G_\delta$ subset of $X\times X$.
Then there exists a dense $G_\delta$ subset $Y$ of $X$
such that for every $x\in Y$, $R(x)=\{y\in X:\, (x,y)\in R\}$ is a dense $G_\delta$ subset of $X$.
\end{lem}
Recall that a sequence $F$ in $\Z$ is called a {\em Poincar\'e sequence} if for any
measure-preserving system $(X, \mathcal B, \mu, T)$ and $A \in\mathcal B$ with $\mu(A) > 0$
there exists $0\neq n\in F$ such that $\mu(A \cap T^{-n}A)>0$.
Let $\F_{Poin}$ denote the family of all Poincar\'e sequences.
It is well known that $\F_{Poin}=\Delta^*(\F_{pubd})$ \cite{F81,W00}.
\begin{thm}\label{thm:weak-mixing}
Let $(X,T)$ be a TDS. Consider the following conditions:
\begin{enumerate}
\item $(X,T)$ is weakly mixing.
\item $Trans_{\F_{ip}}(X,T)$ is residual in $X$.
\item $(X,T)$ is $\F_{ip}$-point transitive.
\item $(X,T)$ is $\F_{Poin}$-transitive.
\end{enumerate}
Then (1)$\Rightarrow$(2)$\Rightarrow$(3)$\Rightarrow$(4).
In addition, if $(X,T)$ is an E-system, then (4)$\Rightarrow$(1).
\end{thm}
\begin{proof}
(1)$\Rightarrow$(2) Let $R=Trans(X\times X, T\times T)$,
then $R$ is a dense $G_\delta$ subset of $X\times X$.
Since $(X,T)$ is transitive, $X$ has no isolated points.
By Ulam Lemma, there exists a dense $G_\delta$ subset $Y$ of $X$
such that for every $x\in Y$, $R(x)$ is a dense $G_\delta$ subset of $X$.
Then it suffices to show that $Y\subset Trans_{\F_{ip}}(X,T)$.
Let $x\in Y$ and $y\in R(x)$, we have $(y,y)\in \overline{Orb((x,y), T\times T)}$,
then by the following claim, $x$ is an $\F_{ip}$-transitive point.
\smallskip
{\bf Claim}: If $(y,y)\in \overline{Orb((x,y), T\times T)}$,
then for every neighborhood $U$ of $y$ the entering time set $N(x,U)$ is an IP set.
{\bf Proof of the Claim}:
For every neighborhood $U$ of $y$, let $U_1=U$
then there exists $p_1\in\mathbb N$ such that
$T^{p_1}x\in U_1$ and $T^{p_1}y\in U_1$. Let $U_2=U_1 \bigcap T^{-p_1}U_1$,
then $U_2$ is a neighborhood of $y$, so there exists $p_2\in\mathbb N$ such that
\[T^{p_2}x\in U_2\textrm{ and } T^{p_2}y\in U_2. \]
Then for every $m\in FS\{ p_i\}_{i=1}^2$
\[T^mx\in U \textrm{ and } T^my\in U.\]
We continue inductively. Assume $p_1,p_2,\ldots,p_n$ have been found such that
for every $m\in FS\{ p_i\}_{i=1}^n$
\[T^mx\in U \textrm{ and } T^my\in U.\]
Let $U_{n+1}=U \bigcap (\bigcap_{m\in FS\{ p_i\}_{i=1}^n} T^{-m}U)$, then $U_{n+1}$ is a
neighborhood of $y$, so there exists $p_{m+1}\in\mathbb N$ such that
\[T^{p_{m+1}}x\in U_{m+1}\textrm{ and } T^{p_{m+1}}y\in U_{m+1}. \]
Then for every $m\in FS\{ p_i\}_{i=1}^{n+1}$
\[T^mx\in U \textrm{ and } T^my\in U.\]
Thus, $FS\{p_i\}_{i=1}^\infty \subset N(x,U)$.
(2)$\Rightarrow$(3) is obvious.
(3)$\Rightarrow$(4) follows from the fact that every IP set is a Poincar\'e sequence (\cite[p74]{F81}).
In addition, if $(X,T)$ is an E-system, we show that (4)$\Rightarrow$(1).
It is sufficient to show that for every opene subsets $U_1, U_2, V$ of $X$,
$N(U_1, U_2)\cap N(V,V)\neq\emptyset$.
Since $(X,T)$ is an E-system, there exists an invariant measure $\mu$ with full support.
Then $(X,\mathcal B_X,\mu, T)$ is a measure dynamical system and $\mu(V)>0$.
Since $N(U_1,U_2)\in \F_{Poin}$, by the definition of Poincar\'e sequence one has
$N(U_1,U_2)\cap N(V,V)\neq\emptyset$.
\end{proof}
\begin{cor}
Let $(X,T)$ be a minimal system. Then $(X,T)$ is weakly mixing if and only if it is $\F_{ip}$-point transitive.
\end{cor}
In \cite{HY02}, the authors constructed an $\F_{Poin}$-transitive system which is not weakly mixing,
we show that:
\begin{prop}\label{prop:Fip-not-wm}
There exists an $\F_{ip}$-point transitive system which is not weakly mixing.
\end{prop}
\begin{proof}
Since the construction is somewhat long and complicated, we leave it to the appendix.
\end{proof}
\subsection{Weakly mixing M-system} In this subsection, we characterize weakly mixing M-systems.
To this end, we need the concept of the central set which was first introduced in \cite{F81}.
Let $(X,T)$ be a TDS, a pair $(x,y)\in X\times X$ is called {\em proximal} if there exists a
sequence $\{n_i\}_{i=1}^\infty$ in $\N$ such that $\lim_{i\to\infty}T^{n_i}x=\lim_{i\to\infty}T^{n_i}y$.
A subset $F$ of $\Z$ is called a {\em central set}, if there exists a system
$(X,T)$, $x\in X$, a minimal point $y\in X$ and a neighborhood of $U$ of $y$
such that $(x,y)$ is proximal and $N(x,U)\subset F$.
Let $\F_{cen}$ denote the family of all central sets.
\begin{lem}\cite[Proposition 8.10]{F81}
$\F_{cen}\subset \F_{ip}\bigcap \F_{ps}$.
\end{lem}
\begin{thm}[Akin-Kolyada \cite{AK03}]
Let $(X,T)$ be a TDS. If $(X,T)$ is weakly mixing, then for every $x\in X$
the proximal cell $Prox(x)=\{y\in X: (x,y)\textrm{ is proximal}\}$
is a dense $G_\delta$ subset of $X$.
\end{thm}
\begin{thm}
Let $(X,T)$ be a TDS. Then the following conditions are equivalent:
\begin{enumerate}
\item $(X,T)$ is a weakly mixing M-system.
\item $Trans_{\F_{cen}}(X,T)$ is residual in $X$.
\item $(X,T)$ is $\F_{cen}$-point transitive.
\end{enumerate}
\end{thm}
\begin{proof}
(1)$\Rightarrow$(2). Let $\{y_n\}_{n=1}^\infty$ be a sequence of minimal points which is dense in $X$.
By Akin-Kolyada Theorem, $\bigcap_{n=1}^\infty Prox(y_n)$ is a dense $G_\delta$ subset of $X$.
Then it suffices to show that $\bigcap_{n=1}^\infty Prox(y_n)\subset Trans_{\F_{cen}}(X,T) $.
Let $x\in \bigcap_{n=1}^\infty Prox(y_n)$ and $U$ be an opene subset of $X$.
There exists $n\in \mathbb N$ such that $y_n\in U$.
Since $(x,y)$ is proximal and $y$ is a minimal point,
by the definition of central set, we have $N(x, U)\in \F_{cen}$.
Thus, $x$ is a $\F_{cen}$-transitive point.
(2)$\Rightarrow$(3) is obvious.
(3)$\Rightarrow$(1).
By $\F_{cen}\subset \F_{ps}$ and Theorem \ref{thm:E-M-system}, $(X,T)$ is an M-system.
Therefore, by Theorem \ref{thm:weak-mixing} and $\F_{cen}\subset \F_{ip}$, $(X,T)$ is weakly mixing.
\end{proof}
\subsection{Weakly mixing E-system} In this subsection, we characterize weakly mixing E-systems.
To this end, we need the concept of the D-set which was first introduced in \cite{BD08}.
A subset $F$ of $\Z$ is called a {\em D-set} if there exists a system
$(X,T)$, $x\in X$, an $\F_{pubd}$-recurrent point $y\in X$ and a neighborhood of $U$ of $y$
such that $(y,y)\in \overline{Orb((x,y),T\times T)}$ and $N((x,y), U\times U)\subset F$.
Let $\F_D$ denote the family of all D-sets.
\begin{lem}[\cite{BD08}]
$\F_D\subset \F_{ip}\bigcap \F_{pubd}$.
\end{lem}
\begin{thm}
Let $(X,T)$ be a TDS. Then the following conditions are equivalent:
\begin{enumerate}
\item $(X,T)$ is a weakly mixing E-system.
\item $Trans_{\F_{D}}(X,T)$ is residual in $X$.
\item $(X,T)$ is $\F_{D}$-point transitive.
\end{enumerate}
\end{thm}
\begin{proof}
(1)$\Rightarrow$(2).
It is easy to see that $(X\times X,T\times T)$ is also an E-system.
Let $R=Trans_{pubd}(X\allowbreak\times X, T\times T)$, then
by Theorem \ref{thm:E-M-system} $R$ is a dense $G_\delta$ subset of $X\times X$.
By Ulam Lemma, there exists a dense $G_\delta$ subset $Y$ of $X$
such that for every $x\in Y$, $R(x)$ is a dense $G_\delta$ subset of $X$.
Then it suffices to show that $Y\subset Trans_{\F_{D}}(X,T)$.
Let $x\in Y$ and $U$ be an opene subset of $X$. Choose $y\in R(x)\cap U$,
then $y$ is $\F_{pubd}$-recurrent.
Since $(x,y)$ is a transitive point in $X\times X$,
$(y,y)\in \overline{Orb((x,y),T\times T)}$.
Clearly, $N((x,y), U\times U)\subset N(x,U)$,
so by the definition of D-set one has $N(x, U)\in \F_D$.
Thus, $x$ is an $\F_{D}$-transitive point.
(2)$\Rightarrow$(3) is obvious.
(3)$\Rightarrow$(1). By $\F_D\subset \F_{pubd}$ and Theorem \ref{thm:E-M-system}, $(X,T)$ is an E-system.
Therefore, by Theorem \ref{thm:weak-mixing} and $\F_D\subset \F_{ip}$, $(X,T)$ is weakly mixing.
\end{proof}
\begin{rem} (1). There exists a weakly mixing system which is not an E-system (\cite{HZ02,HLY09}).
(2). There exists a weakly mixing E-system which is not an M-system (\cite{BD08}).
\end{rem}
\section{Systems with dense small periodic sets}
In this section, we characterize transitive systems with dense small periodic sets.
Let $(X,T)$ be a TDS, we call $(X,T)$ has {\em dense small periodic sets} (\cite{HY05}) if
for every open subset $U$ of $X$ there exists a closed subset $Y$ of $U$ and $k \in\mathbb N$ such
that $Y$ is invariant for $T^k$ (\ie, $T^kY\subset Y$).
Clearly, every P-system has dense small periodic sets.
If $(X,T)$ is transitive and has dense small periodic sets,
then it is an M-system.
To characterize the system with dense small periodic set,
we need a new kind of subsets of $\Z$. A subset $F$ of $\mathbb Z_+$ is called {\em weakly thick}
if there exists some $k\in\mathbb N$ such that
$\{n\in\mathbb Z_+: kn\in F\}$ is thick.
Let $\mathcal F_{wt}$ denote the family of all weakly thick sets.
\begin{lem}\label{lem:dsps}
Let $(X,T)$ be a TDS. Then the following conditions are equivalent:
\begin{enumerate}
\item $(X,T)$ has dense small periodic sets.
\item For every opene subset $U$ of $X$, there exists $k\in\mathbb N$
such that $U$ contains a minimal subsystem of $(X,T^k)$.
\item It is $b\F_{wt}$-point center.
\end{enumerate}
\end{lem}
\begin{proof}
(1)$\Rightarrow$(2) follows from the fact that every system constants a minimal subsystem.
(2)$\Rightarrow$(3). Let $U$ be an opene subset of $X$, then
there exists $k\in\mathbb N$ such that $U$ contains a minimal subsystem $Y$ of $(X,T^k)$.
Choose a point $y\in Y\subset U$. Then $k\N\subset N(y,U)$ and $(X,T)$ is $b\F_{wt}$-point center
since $k\N\in b\F_{wt}$.
(3)$\Rightarrow$(1). Let $U$ be an opene subset of $X$.
Choose an opene subset $V$ of $X$ such that $\overline{V}\subset U$.
Since $(X,T)$ is $b\F_{wt}$-point center, there exists $x\in V$ such that $N(x, V)\in b\F_{wt}$.
Let $F=N(x,V)$, then there exist two sequences $\{a_i\}_{i=1}^\infty$, $\{n_i\}_{i=1}^\infty$ in $\Z$
and $k\in \N$ such that
$\bigcup_{i=1}^\infty(a_i+ k[n_i, n_i+i])\subset F$.
Without lose of generality, assume that
$\lim_{i\to \infty}T^{a_i+kn_i}x \allowbreak=y\in \overline{\{T^n x: n\in F\}}\subset \overline{V}$.
Let $Y=\overline{Orb(y,T^k)}$. Clearly, $T^k Y\subset Y$.
Then it suffices to show that $Y\subset U$.
For every $m\in\N$, if $i>m$, then $a_i+k(n_i+m)\in F$, so
\[T^{km}y=\lim_{i\to\infty}T^{a_i+k(n_i+m)}x\in \overline{\{T^n x: n\in F\}}\subset \overline{V}. \]
Thus, $Y=\overline{\{T^{km}y: m\in\N\}}\subset\overline{V}\subset U$.
\end{proof}
\begin{thm}\label{thm:trans-periodic-set}
Let $(X,T)$ be a TDS. Then the following conditions are equivalent:
\begin{enumerate}
\item $(X,T)$ is transitive and has dense small periodic sets.
\item $Trans_{b\F_{wt}}(X,T)=Trans(X,T)\neq\emptyset$.
\item $(X,T)$ is $b\F_{wt}$-point transitive.
\end{enumerate}
\end{thm}
\begin{proof}
(1)$\Rightarrow$(2) follows from Lemma \ref{lem:F-point-center} and Lemma \ref{lem:dsps}.
(2)$\Rightarrow$(3) is obvious.
(3)$\Rightarrow$(1) also follows from Lemma \ref{lem:dsps}.
\end{proof}
Denote $\F_{rs}=\{F\subset\Z:\, \exists k\in\N, \textrm{ s.t. } k\N\subset F\}$.
Let $(X,T)$ be a TDS, a point $x\in X$ is called a {\em quasi-periodic point}
if it is $\F_{rs}$-recurrent.
\begin{cor}
Let $(X,T)$ be an infinite minimal system. Then the following conditions are equivalent:
\begin{enumerate}
\item $(X,T)$ has dense small periodic sets.
\item It is an almost one-to-one extension of some adding machine system.
\item It has some quasi-periodic point.
\item It is $b\F_{wt}$-point transitive.
\end{enumerate}
\end{cor}
\begin{proof}
(1)$\Leftrightarrow$(2)$\Leftrightarrow$(3) were proved in \cite{HLY09}
and (1)$\Leftrightarrow$(4) follows from Theorem \ref{thm:trans-periodic-set}.
\end{proof}
Recall that a system $(X,T )$ is called {\em totally transitive} if for every $k\in \N$, $(X,T^k)$ is transitive.
In \cite{HY05} Huang and Ye showed that a system which is totally transitive and
has dense small periodic sets is disjoint from every minimal system.
We call such a system {\em HY-system} for abbreviation.
It is not hard to see that an HY-system is a weakly mixing M-system (\cite{HY05}).
\begin{thm}\label{thm:HY-system}
Let $(X,T)$ be a TDS, then the following conditions are equivalent:
\begin{enumerate}
\item $(X,T)$ is an HY-system.
\item $Trans_{\mathcal F_{wt}}(X,T)=Trans(X,T)\neq\emptyset$.
\item $(X,T)$ is $\mathcal F_{wt}$-point transitive.
\item $(X,T^k)$ is an HY-system for every $k\in\N$.
\item $(X^k, T^{(k)})$ is an HY-system for every $k\in\N$,
where $X^k=X\times X\times \cdots\times X$($k$ times) and
$T^{(k)}=T\times T\times \cdots\times T$($k$ times).
\end{enumerate}
\end{thm}
\begin{proof}
(1)$\Rightarrow$(2). We need to show every transitive point is an $\F_{wt}$-transitive point.
Let $x$ be a transitive point and $U$ be an opene subset of $X$.
Since $(X,T )$ has dense small periodic sets,
there exists a closed subset $Y$ of $U$ and $k\in\mathbb N$
such that $T^k Y\subset Y$. Since $(X,T)$ is totally transitive,
$x$ is also a transitive point for the system $(X,T^k)$.
By the continuity of $T$, it is easy to see that $\{n\in\mathbb Z_+: (T^k)^nx\in U\}$ is thick.
Thus, $N(x,U)=\{n\in\Z: T^n x\in U\}$ is weakly thick, so $x$ is an $\F_{wt}$-transitive point.
(2)$\Rightarrow$(3) is obvious.
(3)$\Rightarrow$(4). We first prove the following Claim.
\smallskip
{\bf Claim}: If $F$ is weakly thick, then for every $k\in \mathbb N$,
$F_k=\{m\in\mathbb Z_+: mk\in F\}$ is also weakly thick.
{\bf Proof of the Claim}:
If $F$ is weakly thick, then there exists $r\in\mathbb N$ such that $F_r=\{m\in\mathbb Z_+: rm\in F\}$ is thick.
Then for every $k\in\N$,
\[\{m\in\mathbb Z_+: rm\in F_k\}=\{m\in\mathbb Z_+: krm\in F\}=\{m\in \Z: mk\in F_r\}\]
is also thick. Thus, $F_k$ is weakly thick.
To show that $(X,T^k)$ is an HY-system for every $k\in\N$,
it suffices to show that $(X,T^k)$ is transitive and has dense small periodic sets for every $k\in\N$.
Now fix $k\in\N$. Let $x\in X$ be an $\F_{wt}$-transitive point in $(X,T)$, then by the Claim,
$x$ also is an $\F_{wt}$-transitive point in $(X,T^k)$. Thus, $(X,T^k)$ is transitive
and by Lemma \ref{lem:dsps} $(X,T^k)$ has dense small periodic sets.
(4)$\Rightarrow$(1) and (5)$\Rightarrow$(1) are obvious.
(1)$\Rightarrow$(5). Since every HY-system is weakly mixing, $(X^k, T^{(k)})$ is totally transitive for every $k\in\N$.
Then it suffices to show that $(X^k, T^{(k)})$ has dense small periodic sets for every $k\in\N$.
Now fix $k\in\N$. Let $W$ be an opene subset of $X^k$, then there exist opene subsets $U_1,U_2,\ldots,U_k$ of $X$
such that $U_1\times U_2\times \cdots \times U_k\subset W$. Since $(X,T)$ has dense small periodic sets,
there exists closed subsets $Y_1,Y_2,\ldots, Y_k$ of $U_1, U_2,\ldots, U_k$ and $n_1, n_2,\ldots, n_k\in\N$
such that $T^{n_i}Y_i\subset Y_i$ for $i=1,2,\ldots,k$. Let $Y=Y_1\times Y_2\times\cdots\times Y_k$ and $n=n_1n_2\cdots n_k$.
Then $Y\subset W$ and $(T^{(k)})^{n}Y\subset Y$. Thus, $(X^k, T^{(k)})$ has dense small periodic sets.
\end{proof}
\begin{rem}
(1). There exists a weakly mixing M-system which is not an HY-system (\cite{HY05}).
(2). There exists an HY-system without periodic points (\cite{HY05}).
\end{rem}
\section{Applications}
In this section as applications, we discuss disjointness and weak disjointness.
\subsection{Disjointness}
The notion of disjointness of two TDSs was introduced by Furstenberg his seminal
paper \cite{F67}. Let $(X, T)$ and $(Y, S)$ be two TDSs. We call $J\subset X\times Y$ is a {\em joining} of
$X$ and $Y$ if $J$ is a non-empty closed invariant set, and is projected onto $X$ and $Y$ respectively.
If each joining is equal to $X\times Y$ then we call $(X, T)$ and $(Y, S)$
are {\em disjoint}, denoted by $(X, T) \bot (Y, S)$ or $X\bot Y$. Note that if $(X, T)\bot (Y, S)$
then one of them is minimal, and if $(X, T)$ is minimal then the set of recurrent
points of $(Y, S)$ is dense (\cite{HY05}).
\begin{lem} \label{lem:disjoint}
Let $\F$ be a family.
If $(X,T)$ is $\F$-point transitive and $(Y,S)$ is minimal with $Rec_{\kappa\F}(Y,S)=Y$,
then $(X, T) \bot (Y, S)$.
\end{lem}
\begin{proof}
Let $J$ be a joining of $X$ and $Y$, and $x\in Trans_{\F}(X,T)$.
There exists some $y\in Y$ such that $(x,y)\in J$.
For every opene subset $U$ of $X$ and neighborhood $V$ of $y$,
we have $N(x, U)\in \F$ and $N(y, V)\in \kappa\F$,
then $\overline{Orb((x,y), T\times S)}\cap U\times V\neq\emptyset$.
Therefore, $X\times \{y\}\subset \overline{Orb((x,y), T\times S)}\subset J$.
By $J$ is $T\times S$-invariant and $Y$ is minimal, we have $J=X\times Y$.
Thus, $(X, T) \bot (Y, S)$.
\end{proof}
Recall that a system is called {\em distal} if there is no proper proximal pairs.
\begin{thm}\cite[Theorem 9.11]{F81}
A minimal system $(X,T)$ is distal if and only if $Rec_{\kappa\F_{ip}}(X,T)=X$
\end{thm}
\begin{cor}
Every $\F_{ip}$-point transitive system is disjoint from every minimal distal system.
\end{cor}
Recall that a system $(X,T)$ is called {\em totally minimal} if for every $k\in\N$, $(X,T^k)$ is minimal.
\begin{thm}
(1). Every HY-system is disjoint from every minimal system (\cite{HY05}).
(2). Every transitive system with dense small periodic sets
is disjoint from every totally minimal system.
\end{thm}
\begin{proof}
(1). By Theorem \ref{thm:HY-system}, every HY-system is $\F_{wt}$-point transitive.
Then by Lemma \ref{lem:disjoint} it suffices to show that for every minimal system $(X,T)$ we have
$Rec_{\kappa\F_{wt}}(X,T)=X$.
Let $(X,T)$ be a minimal system, $x\in X$ and $U$ be a neighborhood of $x$.
Then it suffices to show that for every $F\in\F_{wt}$, $N(x, U)\cap F\neq\emptyset$.
Let $F\in\F_{wt}$, there exists $k\in\N$
such that $\{n\in\N: kn\in F\}$ is thick. Since $x$ is a minimal point in $(X,T)$,
$x$ also is a minimal point $(X,T^k)$. Then $\{n\in\N: (T^k)^nx\in U\}$ is syndetic.
Therefore, $N(x,U)\cap F\neq\emptyset$.
(2). By Theorem \ref{thm:trans-periodic-set},
every transitive system with dense small periodic sets is $b\F_{wt}$-point transitive.
Then by Lemma \ref{lem:disjoint} it suffices to show that for every totally minimal system $(X,T)$
we have $Rec_{\kappa b\F_{wt}}(X,T)=X$. We first prove the following Claim.
\smallskip
{\bf Claim}: If $F\in b\F_{wt}$, there exists $q\in \Z$ such that $-q+F\in \F_{wt}$.
{\bf Proof of the Claim}:
Let $F\in b\F_{wt}$, then there exist two sequences $\{a_i\}_{i=1}^\infty$, $\{b_i\}_{i=1}^\infty$ in $\Z$
and $k\in\N$ such that $\bigcup_{i=1}^\infty(a_i + k[b_i, b_i+i])\subset F$.
Without lose of generality, assume that there exists $q\in [0,k)$ such that for every $i\in N$
$a_i\equiv q \pmod k$. Let $a_i=kc_i +q$ and $F'=\bigcup_{i=1}^\infty k[b_i+c_i, b_i+c_i+i]$,
then $F'$ is weakly thick and $q+F'\subset F$.
Thus, $-q+F$ is weakly thick.
Let $(X,T)$ be a totally minimal system, $x\in X$ and $U$ be a neighborhood of $x$.
Then it suffices to show that for every $F\in b\F_{wt}$, $N(x, U)\cap F\neq\emptyset$.
Let $F\in b\F_{wt}$, by the Claim there exists $q\in\Z$
such that $-q+F$ is weakly thick. Let $F'=-q+F$.
Since $(X,T)$ is totally minimal, $N(T^px,U)\cap F'\neq\emptyset$,
then $N(x,U)\cap F\neq\emptyset$ since $N(T^px,U)+p\subset N(x,U)$.
\end{proof}
\subsection{Weak disjointness}
Let $(X, T)$ and $(Y, S)$ be two TDSs, they are called {\em weakly disjoint} if $(X\times Y, T\times S)$ is
transitive, denoted by $(X, T) \curlywedge (Y, S)$ or $X\curlywedge Y$ (\cite{P72}).
It is easy to see that if $(X, T)$ and $(Y, S)$ are weakly disjoint, then both of them are transitive.
\begin{lem}\label{lem:weakly-disjoint}
Let $\F$ be a family.
If $(X,T)$ is $\F$-transitive and $(Y,S)$ is transitive and $\kappa\F$-center,
then $(X, T) \curlywedge (Y, S)$.
\end{lem}
\begin{proof}
Let $U_1, U_2$ be two opene subsets of $X$ and $V_1, V_2$ be two opene subsets of $Y$.
Since $(Y,S)$ is transitive, then there exists $n\in\mathbb N$ such that $V_1\cap S^{-n}V_2\neq\emptyset$.
Let $V=V_1\cap S^{-n}V_2$. Then it is easy to see that
\[ n+N(U_1, T^{-n}U_2)\cap N(V,V)\subset N(U_1\times V_1, U_2\times U_2).\]
Since $(X,T)$ is $\F$-transitive and $(Y,S)$ is $\kappa\F$-center,
$N(U_1, T^{-n}U_2)\cap N(V,V)\neq\emptyset$.
Then $N(U_1\times V_1, U_2\times U_2)\neq\emptyset$.
Thus, $(X\times Y, T\times S)$ is transitive, \ie, $(X, T)\curlywedge (Y, S)$.
\end{proof}
\begin{lem}\label{lem:wt-P}
Let $F\subset \Z$ be weakly thick, then there exists a P-system $(X,T)$,
a transitive point $x\in X$ and an opene subset $U$ of $X$ such that
$N(x,U)\subset F$.
\end{lem}
\begin{proof}
Let $F\subset \Z$ be a weakly thick set, there exists $k\in\N$ and
a sequence $\{a_n\}_{n=1}^\infty$ of $\Z$ such that
$\bigcup_{n=1}^\infty(k[a_n, a_n+n])\subset F$.
Without lose of generality, assume that $a_{n+1}>a_n+2n$.
Let $F_0=\bigcup_{n=1}^\infty k[a_n,a_n+n]$.
We will construct $x^{(n)}=\mathbf{1}_{A_n}\in \{0,1\}^{\Z}$ such that
$A_n\subset F_0$ and $x=\lim x^{(n)}=\mathbf{1}_A$.
Moreover, $X=\overline{Orb(x,T)}$ is a P-system and
$N(x,U)=A\subset F_0$, where $U=\{y\in X:\,y(0)=1\}$.
To obtain $x^{(n)}$ we construct a finite word $B_n$ such that
$x^{(n)}$ begins with $B_n$ and $B_n$ appears in $x^{(n)}$ weakly thick,
and in the next step let $B_{n+1}$ begin with $B_n$.
Let $\{P_i\}_{i=0}^\infty$ is a partition of $\N$ and each $P_i$ is infinite.
\smallskip
{\bf Step 1:} Construct $x^{(1)}$.
Let $B_1=\mathbf{1}_{F_0}[0, ka_1+k-1]$ and $r_1=|B_1|$ be the length of $B_1$.
Put $x^{(1)}[0, r_1-1]=B_1$.
For every $p\in P_1$, if $kp\geq r_1$, let $l$ be the integer part of $\frac{kp}{r_1}$,
put $x^{(1)}[ka_p+jr_1, ka_p+(j+1)r_1-1]=B_1$ for $j=0,1,\ldots, l-1$ and
$x^{(1)}(i)=0$ for other undefined position $i$.
Let $A_1\subset \Z$ such that $\mathbf 1_{A_1}=x^{(1)}$, then $A_1\subset F_0$.
\smallskip
{\bf Step 2:} Construct $x^{(2)}$.
Let $B_2=x^{(1)}[0, ka_2-1]$ and $r_2=|B_2|$ be the length of $B_2$.
Put $x^{(2)}[0, r_2-1]=B_2$.
For every $p\in P_2$, if $kp\geq r_2$, let $l$ be the integer part of $\frac{kp}{r_2}$,
put $x^{(2)}[ka_p+jr_2, ka_p+(j+1)r_2-1]=B_2$ for $j=0,1,\ldots, l-1$ and
$x^{(2)}(i)=x^{(1)}(i)$ for other undefined position $i$.
Let $A_2\subset \Z$ such that $\mathbf 1_{A_2}=x^{(2)}$, then $A_1\subset A_2\subset F_0$.
\smallskip
{\bf Step 3:} If $x^{(n)}$ has been constructed, now construct $x^{(n+1)}$.
Let $B_{n+1}=x^{(n)}[0, ka_{n+1}-1]$ and $r_{n+1}=|B_{n+1}|$ be the length of $B_{n+1}$.
Put $x^{(n+1)}[0, r_{n+1}-1]=B_{n+1}$.
For every $p\in P_{n+1}$, if $kp\geq r_{n+1}$,
let $l$ be the integer part of $\frac{kp}{r_{n+1}}$,
put $x^{(n+1)}[ka_p+jr_{n+1}, ka_p+(j+1)r_{n+1}-1]=B_{n+1}$ for $j=0,1,\ldots, l-1$ and
$x^{(n+1)}(i)=x^{(n)}(i)$ for other undefined position $i$.
Let $A_{n+1}\subset \Z$ such that $\mathbf 1_{A_{n+1}}=x^{(n+1)}$,
then $A_n\subset A_{n+1}\subset F_0$.
In such a way, let $x=\lim x^{(n)}=\mathbf{1}_A$ and $X=\overline{Orb(x,T)}$,
then $x$ is a recurrent point and $N(x,U)=A\subset F_0$.
It is easy to see that $X$ has dense periodic points.
This completes the proof.
\end{proof}
\begin{thm}
Let $(X,T)$ be a TDS, then the following conditions are equivalent:
\begin{enumerate}
\item $(X,T)$ is $\Delta^*(\F_{wt})$-transitive.
\item $(X,T)$ is weakly disjoint from every transitive and $\Delta(\F_{wt})$-center system.
\item $(X,T)$ is weakly disjoint from every $b\F_{wt}$-point transitive system.
\item $(X,T)$ is weakly disjoint from every P-system.
\end{enumerate}
\end{thm}
\begin{proof}
(1)$\Rightarrow$(2) follows from Lemma \ref{lem:weakly-disjoint}.
(2)$\Rightarrow$(3). Let $(Y,S)$ be a $b\F_{wt}$-point transitive system.
It suffices to show that $(Y,S)$ is $\Delta(\F_{wt})$-center.
Let $U$ be an opene subset of $Y$, there exists a $b\F_{wt}$-transitive point $y\in U$.
It is easy to see that $N(U,U)=N(y,U)-N(y,U)$.
Then $N(U,U)\in \Delta(\F_{wt})$ since $N(y,U)\in b\F_{wt}$.
(3)$\Rightarrow$(4) follows from the fact that
every P-system is $b\F_{wt}$-point transitive.
(4)$\Rightarrow$(1). Let $U,V$ be two opene subsets of $X$.
For every $F\in \F_{wt}$, by Lemma \ref{lem:wt-P}
there exists a P-system $(Y,S)$, a transitive point $x\in X$ and an opene subset $W$ of $X$
such that $N(x,W)\subset F$. Then $N(W,W)\subset (F-F)\cup\{0\}$.
Since $(X,T)$ is weakly disjoint from $(Y,S)$, $N(U,V)\cap N(W,W)$ is infinite,
then $ N(U,V)\cap (F-F)\neq\emptyset$.
Thus, $N(U,V)\in \Delta^*(\F_{wt})$.
\end{proof}
\section*{Appendix A}
In this appendix, as announced before, we construct an $\F_{ip}$-point transitive system which is not weakly mixing.
\begin{proof}[Proof of Proposition \ref{prop:Fip-not-wm}]
Let $\Sigma=\{0,1\}^{\mathbb Z_+}$ and $T$ be the shift map.
We will construct a sequence $W_1\subset W_2\subset W_3\cdots\subset \Z$
and $W=\bigcup_{n=1}^\infty W_n$.
Let $x=\mathbf 1_W$ and $X=\overline{Orb(x,T)}$,
we want to show that $(X,T)$ is $\F_{ip}$-point transitive but not weakly mixing.
Let $p_0=0$ and $p^{(i)}_{j,0}=0$ for all $i,j\in\N$. Let $W_0=\{p_0\}=A_0$. Choose $p_1>2$ and set
\[W_1=W_0\cup \{p_1\}.\]
Arrange $W_2$ as
\[W_1=A_0\cup A_1,\]
where $A_1=\{p_1\}$. Let $k_1=1$,
then $\min A_i>3\max A_j +2$ for $0\leq j<i\leq k_1$.
Choose $p_2$ and $p^{(1)}_{1,1}$ are positive integers to be defined later and set
\[W_2=W_1\bigcup\left(W_1+p_2\right)
\bigcup\left(W_1+p^{(1)}_{1,1}-1\right).\]
Arrange $W_2$ as
\[W_2=W_1\bigcup A_2 \bigcup A_3.\]
Let $k_2=3$, choose appropriate $p_2$ and $p^{(1)}_{1,1}$ such that
$\min A_i>3\max A_j +2$ for $0\leq j<i\leq k_2$.
Choose $p_3$, $p^{(1)}_{1,2}$, $p^{(2)}_{1,1}$ and $p^{(2)}_{2,1}$
are positive integers to be defined later and set
\begin{eqnarray*}
W_3=W_2&\bigcup&\left(W_2+p_3\right)
\bigcup \left(W_1+p^{(1)}_{1,2}-1+FS\{p^{(1)}_{1,i}\}_{i=0}^1\right)\\
&\bigcup&\left(W_2+p^{(2)}_{1,1}-1\right)
\bigcup \left(W_2+p^{(2)}_{2,1}-2\right).
\end{eqnarray*}
Arrange $W_3$ as
\[W_3=W_2\bigcup A_{4}\bigcup A_{5}\bigcup A_{6}\bigcup A_{7}.\]
Let $k_3=7$,
choose appropriate $p_3$, $p^{(1)}_{1,2}$, $p^{(2)}_{1,1}$ and $p^{(2)}_{2,1}$
such that
$\min A_i>3\max A_j +2$ for $0\leq j<i\leq k_3$.
Assume that $W_n$ has been constructed, now construct $W_{n+1}$,
choose $p_{n+1}$, $p^{(1)}_{1,n}$, $p^{(2)}_{1,n-1}$, $p^{(2)}_{2,n-1}$,
$\ldots$, $p^{(n)}_{n,1}$
are positive integers to be defined later and set
\begin{eqnarray*}
W_{n+1}=&W_n&\bigcup\left(W_n+p_{n+1}\right)\bigcup
\left(W_1+p^{(1)}_{1,n}-1+FS\{p^{(1)}_{1,i}\}_{i=0}^{n-1}\right)\\
&&\bigcup\left(\bigcup_{j=1}^2\left(W_2+p^{(2)}_{j,n-1}-j+FS\{p^{(2)}_{j,i}\}_{i=0}^{n-2}\right)\right)\bigcup\cdots\\
&&\bigcup\left(\bigcup_{j=1}^{n-1} \left(W_{n-1}+p^{(n-1)}_{j,2}-j+FS\{p^{(n-1)}_{j,i}\}_{i=0}^1\right)\right)\\
&&\bigcup\left(\bigcup_{j=1}^{n} \left(W_n+p^{(n)}_{j,1}-j\right)\right).
\end{eqnarray*}
Arrange $W_{n+1}$ as
\[ W_{n+1}=W_{n}\bigcup A_{k_{n}+1}\bigcup A_{k_{n}+2}\bigcup
\cdots\bigcup A_{k_{n+1}},\]
where $A_s$ is $W_n+p_{n+1}$ or in the form of $W_r+p^{(r)}_{j,t}-j+FS\{p^{(r)}_{j,i}\}_{i=0}^{t-1}$.
Choose appropriate $p_{n+1}$, $p^{(1)}_{1,n}$, $p^{(2)}_{1,n-1}$,
$p^{(2)}_{2,n-1}$,
$\ldots$, $p^{(n)}_{n,1}$
such that
$\min A_i>3\max A_j +2$ for $0\leq j<i\leq k_{n+1}$.
Let $W=\bigcup_{n=0}^\infty W_n$. For every $n\in\N$, let
$P^{(n)}_0=FS\{p_j\}_{j=n+1}^\infty$,
$P^{(n)}_i=FS\{p^{(n)}_{i,j}\}_{j=1}^\infty$ for $i=1,\ldots,n$.
Then for every $n\in \N$, $W_n+P^{(n)}_i-i\subset W$ for $i=0,1,\ldots, n$.
Let $x=\mathbf 1_W$ and $X=\overline{Orb(x,T)}$.
First, we show that $(X,T)$ is $\F_{ip}$-point transitive.
It suffices to show that $x$ is an $\F_{ip}$-transitive point.
Let $r_n=\max W_n$ and
$[W_n]=\{y\in X: $ for $i\in [0, r_n]$, $y(i)=1$ if and only if $i\in W_n\}$.
Then $\{[W_n]: n\in \N\}$ is a neighborhood base of $x$. It is easy to see that
\[N(x, [W_n])\supset \bigcup_{i=0}^n (P^{(n)}_i-i).\]
For every opene subset $U$ of $X$, there exists $k\in \Z$ such that $T^kx\in U$.
By the continuity of $T$, there exists $n>k$ such that $T^k([W_n])\subset U$,
so
\[P^{(n)}_k \subset k+N(x, [W_n])\subset N(x,U).\]
Thus, $x$ is an $\F_{ip}$-transitive point.
To see that $(X,T)$ is not weakly mixing, it suffices to show that $N([1], [1])=W-W$ is not thick.
In fact, we show that for every $n\geq 1$, if $a\neq b\in W_n-W_n$ then $|a-b|>2$.
{\bf Claim 1}: If $a\in A_i-A_i$ for some $i\in [k_n+1, k_{n+1}]$ is not zero, then
there exist $1\leq i_1< i_2 \leq k_n$ such that $a\in A_{i_2}-A_{i_1}$.
We prove this claim by induction of $n$. It is easy to see that
the result holds for $n=1, 2$. Now assume that the result holds for $1,2,\ldots,n-1$.
Let $a\in A_i-A_i$ for some $i\in [k_n+1, k_{n+1}]$.
By the construction, every element in $A_i$ has the same part in
$\{$ $p_{n+1}$, $p^{(1)}_{1,n}-1$, $p^{(2)}_{1,n-1}-1$,
$p^{(2)}_{2,n-1}-2$, $\ldots$, $p^{(n)}_{n,1}-n$$\}$, then
$a\in W_{n-1}-W_{n-1}$, by the assume the proof is complete.
{\bf Claim 2}: For every $n\geq 1$, if $a\neq b\in W_n-W_n$ then $|a-b|>2$.
Again we prove this claim by induction of $n$. It is easy to see that
the result holds for $n=1, 2$.
Now assume that the result holds for $1,2,\ldots,n-1$.
Let $a\neq b\in W_n-W_n$. There exist $s_1, s_2, t_1$ and $t_2\in [0, k_n]$
such that $a\in A_{s_1}-A_{s_2}$ and $b\in A_{t_1}-A_{t_2}$.
By the Claim 1, we can assume that $s_1>s_2$ and $t_1> t_2$.
\begin{enumerate}
\item Case $s_1>t_1$. Then $a-b\geq \min A_{s_1}-3\max A_{s_1-1}>2$.
\item Case $s_1=t_1$.
\begin{enumerate}
\item $s_2=t_2$. Then $a-b\in (A_{s_1}-A_{s_2})- (A_{t_1}-A_{t_2})
=(A_{s_1}-A_{t_1})-(A_{s_2}-A_{t_2})$.
By the Claim 1, we have $a-b=c-d$ for some $c,d\in W_{n-1}-W_{n-1}$,
so $a-b>2$ by induction.
\item $s_2>t_2$. Then $a-b\in (A_{s_1}-A_{s_2})- (A_{t_1}-A_{t_2})
=(A_{s_1}-A_{t_1})-(A_{s_2}-A_{t_2})$. By the Claim 1, we have
$A_{s_1}-A_{t_1}\subset W_{n-1}-W_{n-1}$.
If $s_2> k_{n-1}$, this turn to the case (1), so $|a-b|>2$.
If $s_2\leq k_{n-1}$, then $|a-b|>2$ by induction.
\end{enumerate}
\end{enumerate}
Hence, $(X,T)$ is as required.
\end{proof}
|
\section{The notorious linearity of the evolution equations}
This article is motivated by discussions of the origin of
quantum mechanics, as presented, for example, in Refs. \cite{tHooft10,Elze09a}.
These suggest that quantum mechanics could possibly {\it emerge} by coarse-graining and
sufficiently far from the Planck scale. Thus, the quantum phenomena would
arise from dynamics beneath.
Numerous indications exist which support this view. Nevertheless, the general principles
governing such hypothetical dynamics are under debate
\cite{Elze09b,Elze08,tHooft07,tHooft06a,tHooft06b,Elze05,Blasone05,Vitiello01,Smolin,Adler,Wetterich08,Isidro08} and, perhaps, have not even been touched upon, so far.
In this situation, it may be worth while to focus on aspects of
quantum and classical mechanics which bring them as closely together as possible,
yet sharpen remaining differences. The {\it linearity} of the respective evolution
equations must certainly be counted among such aspects.
Indeed, it is remarkable that dynamical evolution in physics, to a large extent, can be encoded
in one or the other of three linear equations:
\begin{eqnarray}\label{Liouville}
\partial_t\rho &=&\{ H,\rho\}
\;\;, \\ [1ex] \label{Schroedinger}
i\partial_t\Psi &=&\hat H\Psi
\;\;, \\ [1ex] \label{vonNeumann}
i\partial_t\hat\rho &=&[\hat H,\hat\rho ]
\;\;, \end{eqnarray}
which, respectively, are the {\it Liouville equation} (\ref{Liouville}), governing the
evolution of a classical phase space density $\rho$ in terms of the
Poisson bracket $\{\;,\;\}$ and the Hamilton function $H$ of the object under study,
the {\it Schr\"odinger equation} (\ref{Schroedinger}) for the quantum mechanical wave function(al)
in terms of the appropriate Hamilton operator $\hat H$, and the {\it von\,Neumann equation}
(\ref{vonNeumann}) for the density operator $\hat\rho$ in terms of the commutator
$[\;,\;]$ with $\hat H$.
We emphasize that all three equations pertain to (or can be interpreted as)
ensemble theories. They have the generic structure:
\begin{equation}\label{linform}
i\partial_t\;\mbox{``state''}=\hat{\cal L}\;\mbox{``state''}
\;\;, \end{equation}
where the ``state'' appears only {\it linearly} in all cases. This first-order in time differential equation, of course,
assumes widely different forms depending on the forces or dynamics under study.
That is, the ``Liouville superoperator'' $\hat{\cal L}$ on the right-hand side
allows to incorporate physics ranging from classical
non-relativistic single particles all the way to relativistic quantum fields,
employing a phase space ensemble of initial conditions in classical
statistical mechanics and the functional Schr\"odinger picture for interacting quantum fields.
Since the Schr\"odinger equation describes special cases (pure states)
of situations (pure and mixed states) covered by the von\,Neumann equation,
we consider Eqs.\,(\ref{Liouville}) and (\ref{vonNeumann}), eventually in
the form of Eq.\,(\ref{linform}), in the following.
In passing, we recall that {\it nonlinear} modifications of quantum mechanics
have been proposed and severely constrained
in various ways, see, for example, Refs.\,\cite{Iwo,Kibble78,Kibble80,Weinberg}.
In distinction, modifications of the
linear Eq.\,(\ref{vonNeumann}), in particular, have been derived
for ``open systems'' or quantum mechanical objects interacting
with an environment (such as a heat bath of oscillators, etc.).
-- Here, one is clearly {\it not} interested in modifying quantum mechanics
but in applying it to complex situations. --
The generic form of the resulting {\it consistent} dynamics,
replacing Eq.\,(\ref{vonNeumann}) while preserving the defining properties of
the density operator, is known as the {\it Lindblad equation} \cite{L1,L2}.
Its generator provides an important example of a {\it completely positive map}.
Some aspects of this will be discussed in more detail in Section~4,
in order to contrast it with modifications of the linear dynamics introduced there.
A hypothetical fundamental coupling of quantum mechanical and classical degrees of freedom
-- {\it ``hybrid dynamics''} -- instead,
presents a departure from quantum mechanics. While this does not touch the linearity,
in the sense of Eq.\,(\ref{linform}), consistency of the resulting theory
has to be carefully examined.
We remark that there is also a practical interest in certain forms of hybrid dynamics.
The Born-Oppenheimer approximation, for example, is based on
a separation of interacting slow and fast degrees
of freedom of a compound object. The former are treated as approximately classical
while the latter as of quantum mechanical nature.~\footnote{Mean field theory,
based on the expansion of quantum mechanical variables
into a classical part plus quantum fluctuations, leads to another
approximation scheme and another form of hybrid dynamics. This is discussed
more generally for macroscopic quantum phenomena in Ref.\,\cite{Huetal}.}
It must be emphasized, however, that in these cases,
hybrid dynamics is considered as relevant {\it approximate description} of
an intrinsically quantum mechanical object. This has recently been employed,
for example, in a re-derivation of geometric forces and Berry's phase \cite{ZhangWu06}.
Such considerations are and will become increasingly important for
precise manipulations of
quantum mechanical objects by apparently, for all practical purposes classical means,
especially in (sub-)nanotechnological devices.
However, this may be also relevant for studies of
the {\it measurement problem}: Namely, quantum
theory, endowed with the Copenhagen interpretation, {\it assumes} a
coupling of the quantum mechanical object to the classical measuring (or manipulating)
apparatus. However, it remains
profoundly silent about {\it how} this is realized and how the ensuing state reduction
(``collapse of the wave function'') proceeds. --
Truly hybrid dynamics, which does not arise from an approximation, augmented by effects of environment induced decoherence and anharmonic forces, must have something to say about
measurement situations, if it can be formulated consistently.
Presently, we are motivated by interest in the
{\it back-reaction} effect of quantum fluctuations on classical degrees of freedom,
in particular if they are physically distinct. We recall here especially
discussions of the ``semiclassical'' Einstein equation coupling the classical metric
of spacetime to the expectation value of the energy-momentum tensor of quantized matter
fields. Can this be made into a consistent hybrid theory leaving gravity unquantized?
This has recently been re-examined in Ref.\,\cite{Diosi10}; earlier related work includes
Refs.\,\cite{Diosi84,DiosiRev05,Penrose98,Adler03,Mavromatosetal92,Pullinetal08,Hu09,Diosi09}.
Concerning the origin of quantum mechanics from a coarse-grained
deterministic dynamics \cite{tHooft10,Elze09a}, to which we
alluded above, the back-reaction problem can be more provocatively stated
as the problem of the interplay of fluctuations among underlying deterministic and
emergent quantum mechanical degrees of freedom. Or, in short:
{\it ``Can quantum mechanics be seeded?''}
In this paper, we present a few ingredients which may be helpful in further
studies of these questions. --
In the following Section~2, we begin with a brief recapitulation of
our path integral formulation of classical Hamiltonian dynamics.
Some simple observations following from this lead us to
take up the subject of hybrid dynamics in Section~3.
In Section~4, we continue the discussion from a different angle,
namely with some remarks on problems encountered when trying to embed
classical or quantum mechanics in a more general linear dynamics.
\section{A path integral for classical Hamiltonian dynamics}
Our aim here is to solve the Liouville equation (\ref{Liouville}) with the help of a suitable
propagator which, in turn, is represented as a path integral. We summarize essential
steps, while more details of the derivation can be found in Ref.\,\cite{EGV10}.
The following considerations are rather independent of the number of degrees of freedom and
apply to matrix or Grassmann valued variables as well; field theories require a classical
functional formalism, which has been considered elsewhere~\cite{Elze05,Elze07}.
We consider a one-dimensional system, for simplicity.
\subsection{The quantum-like version of the Liouville equation}
We assume conservative forces acting on the classical object under study
and that Hamilton's equations are determined by the generic Hamiltonian function:
\begin{equation}\label{HamiltonianF}
H(x,p):=\frac{1}{2}p^2+V(x)
\;\;, \end{equation}
where $x$ and $p$ denote generalized coordinate and momentum, respectively, and where
$V(x)$ stands for an external potential (a mass parameter will be inserted later).
An ensemble of such objects can be described by
a probability distribution function $\rho$ depending on the $x,p$-coordinates of
phase space and time.
Such a distribution evolves according to the {\it Liouville equation}:
\begin{equation}\label{LiouvilleEq}
-\partial_t\rho =\frac{\partial H}{\partial p}\cdot\frac{\partial \rho}{\partial_x}
-\frac{\partial H}{\partial x}\cdot\frac{\partial \rho}{\partial_p}
=\big\{p\partial_x-V'(x)\partial_p\big\}\rho
\;\;, \end{equation}
with $V'(x):=\mbox{d}V(x)/\mbox{d}x$. -- The relative minus sign
in the Poisson bracket, or between terms here,
reflects the symplectic phase space symmetry. It will give rise to a
commutator structure, which reminds one of quantum mechanics, as we shall see momentarily.
Following a Fourier transformation, $\rho (x,p;t)=\int\mbox{d}y\;\mbox{e}^{-ipy}\rho (x,y;t)$,
and a transformation of the effectively {\it doubled number of spacelike coordinates},
\begin{equation}\label{coordtrans}
Q:=x+y/2\;\;,\;\;\;q:=x-y/2
\;\;, \end{equation}
the Liouville equation becomes:
\begin{eqnarray}\label{Schroed}
i\partial_t\rho &=&\big\{ \hat H_Q-\hat H_q+{\cal E}(Q,q)\big\}\rho
\;\;, \\ [1ex] \label{HX}
\hat H_\chi &:=&-\frac{1}{2}\partial_\chi ^{\;2}+V(\chi )\;\;,
\;\;\;\mbox{for}\;\;\chi =Q,q
\;\;, \\ [1ex] \label{I}
{\cal E}(Q,q)&:=&(Q-q)V'(\frac{Q+q}{2})
-V(Q)+V(q)\;=\;-{\cal E}(q,Q)
\;\;. \end{eqnarray}
Thus, it bears strong resemblance to the {\it von\,Neumann equation}, cf.
Eq.\,(\ref{vonNeumann}),
considering $\rho (Q,q;t)$ as matrix elements of a density operator $\hat \rho (t)$.
We automatically recover the Hamiltonian operator $\hat H$
related to the Hamiltonian function, Eq.\,(\ref{HamiltonianF}),
as in quantum theory. Furthermore, reality and normalization of the phase space
probability distribution $\rho (x,p;t)$ translate into hermiticity and trace
normalization of the density operator $\hat \rho (t)$ \cite{EGV10}.
However, an essential dynamical feature here consists
in the interaction ${\cal E}$ between
{\it bra-} and {\it ket-}states. Thus, generally,
the Hilbert space and its dual are coupled by
a genuine {\it superoperator}, a concept to be defined
in the following
subsection.~\footnote{The interaction ${\cal E}$ is antisymmetric under
$Q\leftrightarrow q$. It follows that the complete
Liouville superoperator on the right-hand side of Eq.\,(\ref{Schroed}), to be compared
with Eq.\,(\ref{linform}), has a symmetric spectrum
with respect to zero and, in general, will not be bounded below.
Related observations were discussed, for example, in
Refs.~\cite{tHooft06a,Elze05,Blasone05,Vitiello01}.}
It is remarkable that the interaction between
bra- and ket-states vanishes under certain
circumstances:
\begin{equation}\label{Ezero}
{\cal E}\equiv 0\;\;\Longleftrightarrow\;\;
\mbox{potential}\;V(x)\;\mbox{is constant, linear, or harmonic}
\;, \end{equation}
rendering a Liouville superoperator of quantum mechanical form,
i.e., as in
Eq.\,(\ref{vonNeumann}).~\footnote{Analogously, the vanishing of ${\cal E}$ in
a field theory amounts to having massive or massless free fields, with or without
external sources, and with or without bilinear couplings. In these cases, anharmonic
forces or interactions are absent.}
This has been discussed in Ref.\,\cite{Elze09a} under the perspective of having
quantum phenomena emerge due to discrete spacetime structure.
In the following, we will study in more detail
the classical dynamics described by Eq.\,(\ref{Schroed}), or by
appropriate generalizations, and pay particular attention to the presence of the
superoperator ${\cal E}$, when comparing with the von\,Neumann equation
and its solution by a propagator.
\subsection{The superspace}
The dynamics of density operators in the general form of Eq.\,(\ref{linform}) can
be conveniently rewritten by introducing the concept of {\it superspace}, also called
{\it Liouville space} \cite{EGV10,superspace}.~\footnote{No relation with the case of
supersymmetry is implied.}
Let $\hat{H}$ denote a Hamiltonian operator, as in quantum
theory, and let $\{ |j\rangle\}$, $j=1,\dots ,N$, present a complete orthonormal
set of basis states,
assuming here that the relevant Hilbert space is $N$-dimensional.
Then, the matrix elements of the von\,Neumann equation (\ref{vonNeumann}) read:
\begin{equation}\label{vN}
i \partial_t \rho_{jk} = [(\hat{H}\hat{\rho})_{jk}- (\hat{\rho}\hat{H})_{jk}]\;\;, \;\;\;
j,k = 1,2,\dots, N
\;\;, \end{equation}
with a density matrix $\rho$ of $N^2$ elements. Or, written as in Eq.\,(\ref{linform}):
\begin{equation}\label{vN1}
i \partial_t \rho_{jk} = \sum_{l,m} {\cal L}_{jk,lm} \rho_{lm}\equiv
\big (\hat{\cal L}\hat\rho\big )_{jk}
\;\;, \end{equation}
where the {\it Liouville superoperator} $\hat {\cal L}$ is now defined by its matrix elements:
\begin{equation}\label{LvN}
{\cal L}_{jk,lm}:= H_{jl} \delta_{mk} - \delta_{jl}H_{mk}
\;\;. \end{equation}
This simple rewriting may suggest to introduce a space in which the
density operator is
a vector. This is the role of the Liouville space (or superspace).
The dynamics of density operators
can then be described in parallel for classical and quantum mechanics. In general, they will
differ, of course, by the precise form of the superoperator, examples of which we have seen
here and in the previous subsection.
Given the Hilbert space, as above, the density operator can be expanded as:
\begin{equation}
\hat{\rho}= \sum_{j,k} \rho_{jk} |j \rangle \langle k|
\;\;. \end{equation}
The family of $N^2$ operators $|j\rangle \langle k|$, with $j,k=1,\cdots,N$, can
be interpreted as a complete set of matrices, or vectors, such that the density operator
becomes:
\begin{equation}
|\rho \rangle= \sum_{j,k} \rho_{jk} |jk \rangle \rangle
\;\;, \end{equation}
where the ``ket'' $|jk \rangle \rangle$ denotes the Liouville space {\it vector} representing the
Hilbert space {\it operator} $|j\rangle \langle k|$. Similarly, we introduce a ``bra''-vector
$ \langle \langle jk|$ as the Hermitian conjugate of $|jk \rangle \rangle$.
Consequently, any operator $\hat{A}$ is represented by a vector, denoted by
$|A \rangle \rangle $, and can be expanded as:
\begin{equation} \label{opex}
| A \rangle \rangle= \sum_{j,k} A_{jk} |jk \rangle \rangle
\;\;, \end{equation}
where $A_{jk}$ are the usual matrix elements $\langle j| \hat{A}| k \rangle$. --
Furthermore, with a bra-vector $ \langle \langle B |$ representing $\hat{B}^{\dag}$,
the scalar product of two operators is defined by:
\begin{equation}
\langle \langle B| A \rangle \rangle :=\mbox{Tr} (\hat{B}^{\dag}\hat{A})
\;\;. \end{equation}
This implies the orthonormality condition:
\begin{equation}\label{orthogonality}
\langle \langle jk|mn \rangle \rangle =\mbox{Tr}(|k\rangle \langle j| m \rangle \langle n|)=
\delta_{kn} \delta_{jm}
\;\;, \end{equation}
in analogy to and based on $\langle j|k\rangle = \delta_{jk}$. -- Finally, consider the scalar product:
\begin{equation}
\langle \langle jk| A \rangle \rangle =\mbox{Tr}(|k \rangle \langle j| \hat{A})=
\sum_{l} \langle l|k \rangle \langle j |\hat{A}| l \rangle= \langle j| \hat{A}|k \rangle \equiv A_{jk}
\;\;. \end{equation}
Upon substitution in Eq.\,(\ref{opex}), this yields:
\begin{equation}
|A \rangle \rangle = \sum_{j,k} |jk \rangle \rangle \langle \langle jk | A \rangle \rangle
\;\;. \end{equation}
This is consistent with the following completeness relation in Liouville space:
\begin{equation}\label{completeness}
\sum_{j,k} |jk \rangle \rangle \langle \langle jk | = {\mathbf 1}
\;\;. \end{equation}
It is easy to see now that the Liouville
space is a linear space and that the density operator $\hat{\rho}$, in particular, is a
vector in this space. -- To conclude these formal considerations, we define a linear operator
in terms of its matrix elements:
\begin{equation}
\hat{\cal F}:= \sum_{j,k,m,n} |jk \rangle \rangle \langle \langle jk |
\hat{\cal F}|mn \rangle \rangle \langle \langle mn| \equiv
\sum_{j,k,m,n}{\cal F}_{jk,mn}|jk \rangle \rangle \langle \langle mn|
\;\;, \end{equation}
an example of which is given in Eq.\,(\ref{LvN}).
The importance of Liouville space for classical {\it and} quantum dynamics resides in that
Liouville and von\,Neumann equations, both, assume the form of Eq.\,(\ref{linform}),
incorporating an appropriate superoperator $\hat {\cal L}$, cf.
Eqs.\,(\ref{Schroed})--(\ref{I}) and (\ref{vN})--({\ref{LvN}), respectively.
Thus, we find a close formal similarity between the structure of these equations
and the Schr\"odinger equation (\ref{Schroedinger}).
Therefore, it is plausible that formal derivations or techniques concerning the solution of
the Schr\"odinger equation can be transferred to the cases of Liouville or
von\,Neumann equations with the help of Liouville space notions. This regards
perturbation theory as much as nonperturbative methods, the path integral approach
in particular, to which we turn next.
\subsection{The Liouville propagator and path integral}
The technical ingredients of the Feynman path integral approach,
for the derivation of quantum mechanical propagators in particular, are very
well known \cite{Schulman,Kleinert}. We will make use of these ideas, in order
to derive a path integral for the propagator of density matrices based on
the Liouville space formulation of the preceding subsection.
Our derivation relies on the close formal similarity between the classical
Liouville equation and the von\,Neumann equation on one hand side and the
Schr\"odinger equation on the other, in the appropriate representation that
we discussed.~\footnote{In this subsection, we reinstate $\hbar$ explicitly.}
In particular, the formal solutions of the classical Liouville equation
and of the quantum mechanical von\,Neumann equation, both, can be written in the form:
\begin{equation}
|\rho(t)\rangle\rangle =\mbox{e}^{-i\hat {\cal L}(t-t_0)/\hbar}|\rho(t_0)\rangle\rangle
\;\;, \end{equation}
where $\hat {\cal L}$ is the relevant Liouville superoperator. Here, we have:
\begin{equation}
\langle\langle Q,q|\hat {\cal L}|Q',q'\rangle\rangle =
\delta (Q-Q')\delta (q-q')\big (\hat H(Q)-\hat H(q)+{\cal E}(Q,q))
\;\;, \end{equation}
where $\hat H$ denotes the appropriate Hamilton operator in coordinate representation, as
indicated, which alone is relevant for the von\,Neumann equation, while ${\cal E}$
represents the additionally present superoperator term for classical dynamics, cf.
Subsection\,2.1.
In order to solve the problem of time evolution in the present case, we need to know
the (super)matrix elements entering the propagation equation:
\begin{equation}\label{rhoprop}
\langle\langle Q,q|\rho(t)\rangle\rangle =\int\mbox{d}Q'\mbox{d}q'\;
\langle\langle Q,q|\mbox{e}^{-i\hat {\cal L}(t-t_0)/\hbar}|Q',q'\rangle\rangle\langle\langle Q',q'|\rho(t_0)\rangle\rangle
\;\;, \end{equation}
in analogy to the case of a state vector evolving according the Schr\"odinger equation.
Not surprisingly, we can now follow the usual steps \cite{Schulman,Kleinert}, in order
to construct the path integral representation of our propagator \cite{EGV10}.
In this derivation, one has to pay attention to a suitable generalization of the
Trotter product formula. This works out in a straightforward way; the relevant
definitions and details of the proof are given in Ref.\,\cite{FabThesis}.
Rewriting the Eq.\,(\ref{rhoprop}) as:
\begin{equation}\label{rhoprop1}
\rho (Q,q;t) = \int\mbox{d}Q'\mbox{d}q'\;{\cal G}(Q,q;t|Q',q';t_0)\rho (Q',q';t_0)
\;\;, \end{equation}
our interest is to know the {\it superpropagator} ${\cal G}$.
Following the recipe to arrive at a path integral representation,
we make use here of suitably inserted complete sets of superspace vectors, such as:
\begin{equation}\label{supercomplete}
\int\mbox{d}Q\mbox{d}q\; |Q,q\rangle\rangle\langle\langle Q,q|={\mathbf 1}
\;\;, \end{equation}
and, correspondingly, for momentum space, cf. Eq.\,(\ref{completeness}).
Using the plane wave relation between coordinate and momentum eigenfunctions, we also employ:
\begin{equation}\label{planewaves}
\langle\langle P,p|=\frac{1}{2\pi\hbar }\int\mbox{d}Q\mbox{d}q\;
\exp\big (-\frac{i}{\hbar}(PQ-pq)\big )\langle\langle Q,q|
\;\;. \end{equation}
Furthermore, the orthogonality relation
$\langle\langle Q,q|Q',q'\rangle\rangle =\delta (Q-Q')\delta(q-q')$, cf.
Eq.\,(\ref{orthogonality}), implies:
\begin{equation}\label{transf}
\langle\langle P,p|Q,q\rangle\rangle =\frac{1}{2\pi\hbar }
\exp\big (-\frac{i}{\hbar}(PQ-pq)\big )
\;\;. \end{equation}
Then, with all following steps of the derivation in parallel with the usual ones, it is
straightforward to obtain the {\it Liouville path integral}~\cite{EGV10,FabThesis}:
\begin{equation}\label{superpropagator}
{\cal G}(Q_f,q_f;t|Q_i,q_i;t_0)=\int {\cal D}Q{\cal D}q
\;\exp\big (\frac{i}{\hbar}{\cal S}[\dot Q,Q;\dot q,q]\big )
\;\;, \end{equation}
with the boundary conditions $Q(t_i)=Q_i$, $q(t_i)=q_i$, $Q(t_f)=Q_f$, and $q(t_f)=q_f$,
otherwise unrestricted paths, and with the {\it superaction} ${\cal S}$ defined by:
\begin{equation}
\label{superaction1}
{\cal S}:=\int_{t_0}^t\mbox{d}\tau\;\Big ({\textstyle \frac{m}{2}}\dot Q^2-V(Q)
-\big ({\textstyle \frac{m}{2}}\dot q^2-V(q)\big )
-{\cal E}(Q,q)\Big )
\;\;, \end{equation}
for a particle of mass $m$.
We recall that ${\cal E}\equiv 0$ corresponds to
evolution according to the von\,Neumann equation, whereas ${\cal E}\neq 0$ represents
the {\it only} modification due to classical dynamics, in accordance with the Liouville
equation, cf. Eqs.\,(\ref{Schroed})--(\ref{I}).
However simple this result may seem, the Eqs.\,(\ref{superpropagator})--(\ref{superaction1})
describe time evolution of the full density matrix. The
particular new feature is that formally {\it classical dynamics is treated on the same
footing as quantum mechanics}, differing only in the action entering the phase in the
integrand of the path integral.
These considerations are limited neither by one dimension nor by
single-particle physics, but can be extended all the way to relativistic
field theories. This offers new calculational tools, new approximation methods
in particular, and may be of interest for applications in classical statistical mechanics.
In the following, however, we turn to the quantum-classical divide.
\subsection{Simple properties of the Liouville path integral}
First of all, it seems worth while to record a few simple properties of the
Liouvile path integral.
We observe that in the absence of forces, $V\equiv 0$, the classical and quantum
mechanical propagators {\it coincide}, according to
Eqs.\,(\ref{superpropagator})--(\ref{superaction1}). Which implies that classical
and quantum mechanical behaviour can differ at most in the initial states that
are being propagated, in this case. -- This holds true even for $V\neq 0$,
if ${\cal E}\equiv 0$, i.e., for harmonic forces, cf. (\ref{Ezero}). -- Thus,
the cherished ``textbook effect'' of the {\it spreading of
wave packets} is not a peculiar (kinematical) quantum effect, but rather an effect
of the particular states considered!
A simple calculation, taking into account our above transformations from
$x,p$- to $Q,q$-coordinates (and back), shows that a {\it classical point particle} with
initial phase space distribution $\propto\delta (x-x_0)\delta (p-p_0)$ is
propagated to the distribution $\propto\delta (x-x_0-p_0t/m))\delta (p-p_0)$,
i.e. along a straight line path, as expected. -- Instead, a free
Gaussian wave packet in $Q,q$-space spreads in the way described in textbooks on quantum mechanics.
Finally, we can convince ourselves that the classical dynamics is properly
represented by the Liouville propagator, in general, by undoing the coordinate
transformations (\ref{coordtrans}) directly in the path integral representation of
the superpropagator:
\begin{equation}\label{coordtransInverse}
y:=Q-q\;\;,\;\;\;x:=\frac{Q+q}{2}
\;\;, \end{equation}
a transformation with unit Jacobian. Following a partial integration of
the superaction, the path integral over the $y$-coordinate simply yields a
functional $\delta$-``function'' and results in:
\begin{equation}\label{superpropagator1}
{\cal G}(x_f,y_f;t|x_i,y_i;t_0)=\int {\cal D}x\;
\mbox{exp}(im\dot x_fy_f)\;\delta [m\ddot x+V'(x)]\;\mbox{exp}(-im\dot x_iy_i)
\;\;, \end{equation}
with the boundary conditions $x(t_{i,f})=x_{i,f}$, similarly for the
velocities, and where $y_{i,f}$
denote the initial and final values of the $y$-variable, respectively.
The phase factors stem from the partial integration of the superaction.
Thus, we find the expected result that only paths following solutions of the
{\it classical} equations of motion contribute to the propagator. The role of the
phase factors is easily understood by recalling Eq.\,(\ref{rhoprop1}):
After the inverse coordinate transformations (\ref{coordtransInverse}), one
of the two ordinary integrations there becomes here, including the relevant (initial) phase
factor:
\begin{equation}\label{phaseintegral}
\int\mbox{d}y_i\;\mbox{e}^{-im\dot x_iy_i}\rho (x_i,y_i;t_0)=
\rho (x_i,m\dot x_i\equiv p_i;t_0)
\;\;, \end{equation}
i.e., this incorporates the inverse Fourier transformation, back to the momentum variable
of phase space. The second (final) phase factor, then, is necessary for
the propagator to fulfil the important semi-group property \cite{Schulman}.
\section{Hybrid dynamics for two interacting objects}
We have seen that the path integral for the propagator of the classical Liouville equation
shows significant similarity with the propagator of the quantum mechanical
von\,Neumann equation. This suggests a new perspective on {\it hybrid dynamics} --
the hypothetical direct coupling of quantum and classical degrees of freedom that we discussed
in Section\,1.
In particular, we are interested here in the propagator
for the density matrix of a bi-partite system, composed of a classical and a quantum mechanical
particle of masses $m,m'$ in external potentials $V,v$, respectively.
Correspondingly, the relevant coordinates
will be denoted by $Q,q$ and $Q',q'$, respectively. Our previous
considerations lead us to propose the following
{\it hybrid superpropagator} (cf. Eq.\,(\ref{superpropagator})):
\begin{equation}\label{hsuperpropagator}
{\cal G}_{\mbox{h}}(Q_f,q_f;Q'_f,q'_f;t|Q_i,q_i;Q'_i,q'_i;t_0)=
\int {\cal D}Q{\cal D}q{\cal D}Q'{\cal D}q'
\;\exp\big (\frac{i}{\hbar}{\cal S}[\dot Q,Q;\dot q,q;\dot Q',Q';\dot q',q']\big )
\;\;, \end{equation}
where the superaction naturally consists of three contributions,
${\cal S}\equiv {\cal S}_{\mbox{cl}}+{\cal S}_{\mbox{qm}}+{\cal S}_{\mbox{h}}$:
\begin{eqnarray}\label{Scl}
{\cal S}_{\mbox{cl}}&:=&\int_{t_0}^t\mbox{d}\tau\;\Big (
{\textstyle \frac{m}{2}}\big (\dot Q^2-\dot q^2\big )-(Q-q)V'(\frac{Q+q}{2})
\Big )
\;\;, \\ [1ex] \label{Sqm}
{\cal S}_{\mbox{qm}}&:=&\int_{t_0}^t\mbox{d}\tau\;\Big (
{\textstyle \frac{m}{2}}\big (\dot Q'^2-\dot q'^2\big )
-\big (v(Q')-v(q')\big )
\Big )
\;\;, \\ [1ex] \label{Sh}
{\cal S}_{\mbox{h}}&:=&-\int_{t_0}^t\mbox{d}\tau\; {\cal V}(Q,q;Q',q')
\;\;. \end{eqnarray}
Several remarks are in order here. The contribution ${\cal S}_{\mbox{cl}}$ is the
classical superaction arrived at in Eqs.\,(\ref{superpropagator})--(\ref{superaction1});
we just inserted ${\cal E}$, Eq.\,(\ref{I}), explicitly and collected terms. Correspondingly,
the contribution ${\cal S}_{\mbox{qm}}$ is the action for the propagator of the
von\,Neumann equation, i.e., it describes the quantum mechanical particle in the usual way.
Thus, in the absence
of the interaction term ${\cal S}_{\mbox{h}}$, the hybrid superpropagator
${\cal G}_{\mbox{h}}$ factorizes into the corresponding ones for Liouville and
von\,Neumann equations. This would consistently describe a composite system
of two independent, classical and quantum mechanical particles.
The interaction term ${\cal S}_{\mbox{h}}$ introduces the coupling responsible for
{\it hybrid dynamics}. -- For our present purposes, we restrict this
coupling by two consistency requirements:~\footnote{The question of
consistency of the map induced by the superpropagator which evolves the quantum-classical
hybrid density will be discussed in more detail elsewhere.} \\ ({\bf A})
${\cal S}_{\mbox{cl}}+{\cal S}_{\mbox{h}}$ describes a classical particle
interacting with another particle (coordinates $Q',q'$), as if the latter were
{\it classical}; \\ ({\bf B})
${\cal S}_{\mbox{qm}}+{\cal S}_{\mbox{h}}$describes a quantum mechanical particle
interacting with another particle (coordinates $Q,q$), as if the latter were
{\it quantum mechanical}.
In view of
the general form of the external potential terms in Eqs.\,(\ref{Scl}) and (\ref{Sqm}),
these conditions can only be fulfilled, if ${\cal V}(Q,q;Q',q')$ is {\it harmonic},
i.e., is a polynomial of degree less than or equal to two in all variables. Thus,
the hybrid coupling is of a form that admits a classical or quantum mechanical
interpretation, depending on whether it is viewed from the classical or quantum mechanical
subsystem.
An example is provided by a distance dependent oscillator potential,
$\lambda^{-1}{\cal V}:=(x-x')^2$. Following the derivation of the classical superpropagator,
this becomes, in the above coordinates:
\begin{eqnarray}
\lambda^{-1}{\cal V}(Q,q;Q',q')&=&
\left ((Q-q)\partial_{(Q+q)/2}+(Q'-q')\partial_{(Q'+q')/2}\right )
\left ((Q+q)/2-(Q'+q')/2\right )^2
\nonumber \\ [1ex] \label{CLQM}
&=&(Q-q)(Q+q)+(Q'-q')(Q'+q')^2-2(QQ'-qq')
\;\;, \end{eqnarray}
i.e., it can be seen as coupling between two classical particles, fulfilling condition ({\bf A}).
On the other hand, the separable oscillator terms here also
equal $Q^2-q^2+Q'^2-q'^2$, i.e., are of quantum mechanical form, in accordance with
(\ref{Ezero}), and the bilinear coupling can as well be seen as quantum mechanical, fulfilling condition
({\bf B}).
Interesting consequences of these conditions, determining a {\it harmonic hybrid interaction},
are summarized in the following Table:
\\
\begin{center}
\begin{tabular}{|lll|l|}
\hline
$V$&$v$&${\cal V}$&resulting dynamics \\
\hline
h&h&h&CL or QM \\
anh&h&h&CL \\
h&anh&h&QM \\
anh&anh&h&{\bf ???} \\
\hline
\end{tabular}
\end{center}
{\small \vskip 0.2cm Indicated are the
nature of the potentials acting on the classical and quantum mechanical
particles and their interaction, $V$, $v$, and ${\cal V}$, respectively -- h: harmonic; anh: anharmonic -- and the character of the
resulting hybrid dynamics -- CL: classical; QM: quantum mechanical.}
\vskip 0.25cm
For example, if
the classical and quantum mechanical particles are governed by a harmonic potential and
an anharmonic potential, respectively, then the resulting dynamics of the
composite system can be considered as that of two bilinearly coupled quantum mechanical
particles, one of which moves in a harmonic potential. -- The case of both potentials
being anharmonic is most interesting. However, no general statements can be made in
this case without further study.
The quantum mechanical or classical behaviour of the
composite system might very well depend on {\it where} (concerning its variables) one
is looking at this object!~\footnote{These findings might help with
the problem posed in Section~1: {\it ``Can quantum mechanics be seeded?''}}
\subsection{Intra- and interspace entanglement}
There is a {\it qualitative difference between CL and QM}, contained in
the full path integral for the superpropagator,
Eqs.\,(\ref{hsuperpropagator})--(\ref{Sh}).
Following standard arguments which lead back from a path integral
to an equivalent equation of motion for the (hybrid) density operator $\hat\rho$,
several related observations may be interesting.
The QM evolution is
generated by a commutator of the Hamiltonian with the density operator. Generally, this
superposes and, for multi-partite systems, in particular, entangles underlying
bra- and ket-states separately, $\propto H_{ij}\rho_{jk}-\rho_{ij}H_{jk}$
(using a convenient notation with discrete indices). For a bi-partite system,
it is revealing to write the relevant interaction terms explicitly:
\begin{equation}\label{QMentangle}
[\hat H_{int},\hat\rho ]=\hat H_1\hat\rho_1\otimes\hat H_2\hat\rho_2-\hat\rho_1\hat H_1
\otimes\hat\rho_2\hat H_2
\;\;, \end{equation}
for an interaction $\hat H_{int}:=\hat H_1\otimes\hat H_2$, with the factors acting on
subsystems ``1'' and ``2'', respectively, and where $\hat\rho =\hat\rho_1\otimes\hat\rho_2$,
for a separable initial state.
This has been called {\it dynamically assisted entanglement generation}, see, for example, Refs.\,~\cite{Jacquod,JacquodRev,Hornberger}.
It may come as a surprise that the CL evolution does this just as well,
due to the structure of the superoperator. For polynomial interactions,
in particular, the superoperator
{\it always} contains a contribution proportional to the usual QM terms.
However, the CL evolution, generally, produces additional correlations in $\hat\rho$,
due to terms contained in ${\cal L}_{ij;kl}\rho_{kl}$
which {\it entangle bra- and ket-states}. --
In comparison with Eq.\,(\ref{QMentangle}), such terms can have the unfamiliar structure:
\begin{equation}\label{CLentangle}
\hat H'_1\hat\rho_1\otimes\hat\rho_2\hat H'_2-\hat\rho_1\hat H'_1\otimes\hat H'_2\rho_2
\;\;, \end{equation}
which differs decidedly from a commutator. -- This leads us to distinguish {\it intra-}
(i.e., within given tensor product Hilbert space of subsystems ``1'' and ``2'') and
{\it inter-space entanglement} (i.e., between said Hilbert space and its dual).
Consider, for example, the anharmonic potential
$V(x_1-x_2):=\lambda (x_1-x_2)^4$ for a bi-partite system consisting of particles ``1''
and ``2''. Similarly as before, this leads here to the interaction:
\begin{equation}\label{2particle}
{\cal V}(Q_1,q_1;Q_2,q_2)=
\frac{1}{2}\lambda\big (Q_1-q_1-(Q_2-q_2)\big )\big (Q_1+q_1-(Q_2+q_2)\big )^3
\;\;, \end{equation}
in terms of, by now, familiar variables, taking into account subsystems ``1'' and ``2'' with
$Q$'s and $q$'s refering to bra- and ket-states, respectively. Besides separable terms,
$\propto (Q_a-q_a)(Q_a+q_a)^3,\; a=1,2$, there are terms which mix (and entangle)
variables of both subsystems, as usual in QM. However, there are also additional
terms that refer to Hilbert space and its dual simultaneously (and entangle corresponding
states), for example, $\propto Q_aQ_bq_b^2,\; b\neq a$.
In retrospect, somehow, such difference between CL and QM
evolution had to be expected:
instead with superstates $|Q,q\rangle\rangle$, we could have
worked with superstates $|x,p\rangle\rangle$, relating to coordinates and momenta of
the classical theory. There, coordinates and momenta end up tightly correlated, due to
Hamilton's equations, and produce inter-space entanglement in an interacting bi-partite
system.
Thus, the confrontation of CL with QM, as in our side-by-side study,
is quite revealing. In particular, we speculate that this opens new views on
generating entanglement in multipartite systems, perhaps, by evolving through
quasiclassical stages or by making use of decohered intermediary
states.~\footnote{Previous considerations of the semiclassical regime, such as in
Refs.~\cite{Jacquod,JacquodRev}, were motivated as suitable approximations of the
quantum mechanical evolution, in particular, for studies of the different decoherence
properties between classically regular and chaotic systems. Our results seem
to show that crossing the quantum-classical divide may offer an additional resource
for entanglement generation and related ``truly quantum'' phenomena. This might be related to
a common ``underlying reality'' of CL and QM physics, assumed to consist, for example,
only in statistical correlations in Refs.\,\cite{Zwitters,Khrennikov}.}
Concerning the quantum-classical divide, the present
analysis shows that there is an appealing, if not puzzling formal similarity between CL and
QM. However, this demonstrates one more time that what has been discussed in various ways
as CL limit of QM -- and which is similarly relevant for ``emergent QM'' -- deserves further
study.
While our work has been concerned mainly with the evolution of CL or QM objects,
we recall that V.I.\,Man'ko and collaborators have pointed out that classical
states may differ widely from what could be obtained as the ``$\hbar\rightarrow 0$''
limit of quantum mechanical ones. They show that all states can
be classified by their `tomograms' as {\it either} CL {\it or} QM, CL {\it and} QM,
and {\it neither} CL {\it nor} QM~\cite{Manko}.
The classical limit can be considered a limit
``{\it F}or{\it A}ll{\it P}ractical{\it P}urposes'', gradually approached with decoherence
as an essential but insufficient ingredient or, formally, following the mnemonic
``$\hbar\rightarrow 0$'' rule.
However, in order to truly bridge the qualitative difference between intra- and inter-space
entanglement that we find, and explain the ``Man'ko classes of states'',
some unknown dynamics seems missing.~\footnote{A simple attractor model,
motivated by assumptions about effects of fundamental spacetime discreteness~\cite{Elze09a},
has been discussed in Ref.~\cite{Elze09b}.}
The problems discussed here lead us to the question:
``Does the $\hbar\rightarrow 0$ deformation of QM provide the only {\it interesting
linear dynamics} besides QM itself?''
\section{How special is quantum mechanics?}
We are interested here once more in the structure of the linear evolution equations
that we have discussed side by side, namely the classical Liouville equation and
the von\,Neumann equation of quantum mechanics. However, we consider this with respect
to a hypothetical most {\it general linear dynamics} that can be represented in the
generic form of Eq.\,(\ref{linform}):
\begin{equation}\label{linform1}
\partial_t\rho_{ij}=\frac{1}{i}\sum_{k,l}{\cal L}_{ij,kl}\rho_{kl}
\;\;, \end{equation}
where the indices ($i,j,k,l=1,\dots ,N$) refer to a discrete, finite dimensional
Hilbert space, which we assume for simplicity. Thus, generally, there are $N^2\times N^2$
complex coefficients to be specified. Imposing the {\it constraint} that the `density
matrix' is and remains Hermitian, which requires that
${\cal L}_{ij,kl}=-{\cal L}_{ji,lk}^\ast$, the set of coefficients
can be specified by $N^4$ {\it real parameters}.
This should be compared with the von\,Neumann equation -- which we discussed, in order to
introduce the concept of a superoperator, cf. Eqs.\,(\ref{vN})--(\ref{LvN}) in Section~2.2.
In terms of a Hermitian Hamiltonian, $\hat H$, the matrix elements of which are
specified by $N^2$ {\it real parameters}, the corresponding superoperator has been obtained as:
${\cal L}_{ij,kl}\equiv H_{ik}\delta_{lj}-\delta_{ik}H_{lj}$.
Therefore, we can write a symbolic relation, concerning
the number of real parameters which determine the respective dynamical
equation:
\begin{equation}\label{mnemonic}
``\; \mbox{QM}\sim\sqrt{\mbox{GL}}\; "
\;\;, \end{equation}
i.e., the number of real parameters entering the quantum mechanical evolution law (QM)
scales with the Hilbert space dimension like the square root of the corresponding
number for the most general linear dynamics (GL) which preserves hermiticity.
By choosing the eigenstates of the Hamiltonian as basis of the Hilbert space,
the number of relevant parameters in QM could be further reduced to $N$, the number of real
eigenvalues of $\hat H$.
We may wonder
whether the commutator in Eq.\,(\ref{vN}), $[\hat H,.]$, presents
the {\it minimal structure} preserving hermiticity, normalization, and positivity
of the density matrix. -- Conversely, can
the different numbers of real parameters between a more general linear evolution
law and QM be attributed to an {\it attractor mechanism} or {\it information loss} if and when
QM is emergent \cite{tHooft10,Elze09a}?
As a first step to investigate these issues, we consider here whether QM admits a more general
linear evolution. -- We recall the following result for the QM
of {\it open systems}, see Ref.\,\cite{Diosi} and references there:
\begin{itemize}
\item If and only if the Liouville superoperator $\hat{\cal L}$ entering the right-hand side
of Eq.\,(\ref{linform1}) generates a
{\it Hermitian, trace preserving}, {\it completely positive map},
$\hat{\cal M}(t)\hat\rho (0):=\exp (-i\hat{\cal L}t)\hat\rho (0)$,
then it can be written in canonical {\it Kraus form}:
\begin{equation}\label{KrausForm}
\hat{\cal M}(t)\hat\rho (0) =\sum_k\hat M_k(t)\hat\rho (0)\hat M_k^{\;\dagger}(t)
\;\;, \end{equation}
with $\sum_k\hat M_k^{\;\dagger}\hat M_k={\mathbf 1}_{N\times N}$.
\item Such a map $\hat{\cal M}$ applied to a
Hermitian, positive semidefinite, normalized density operator yields another one
and linear maps beween density operators can be written in Kraus form.
\end{itemize}
An evolution equation corresponding to such a map, in general, contains the
von\,Neumann equation as a special case. --
In fact, the solution of the von\,Neumann equation (\ref{vN}) is
provided by a unitary transformation,
$\hat\rho (t)=\hat U(t)\hat\rho (0)\hat U(t)^\dagger$,
which presents the simplest case with only one unitary Kraus operator,
$\hat M(t)\equiv \hat U(t):=\exp (-i\hat Ht)$.
Therefore, any generalization
which maintains the defining properties of density operators, generally,
will need additional Kraus operators specified by additonal parameters. --
In this sense, the commutator defining the von\,Neumann equation presents
the {\it minimal structure} indeed.
Furthermore,
it has been shown that the dynamics generated by $\hat {\cal M}$ of the Kraus form,
in general,
is equivalent to a non-unitary reduced dynamics of a unitary dynamics on a bigger
(tensor product) Hilbert space ``$S$ystem$\,\otimes E$nvironment'' \cite{Diosi}:
\begin{equation}\label{KrausDyn}
\hat {\cal M}\hat\rho_S =
\mbox{Tr}_E\left (\hat U_{big}\hat\rho_S\otimes\hat\rho_E\hat U_{big}^\dagger\right )
\;\;, \end{equation}
with $\hat U_{big}\hat U_{big}^\dagger =\hat U_{big}^\dagger\hat U_{big}=1$.
Hence, we learn:
\begin{itemize}
\item Hermitian, trace preserving, completely positive maps, while generalizing
the von\,Neumann dynamics,
are {\it not} general enough to ``leave QM''. They describe the QM of {\it open systems}.
\end{itemize}
However, this kind of maps does {\it not}
exhaust the larger set of maps generated by all possible Liouville superoperators.
We conclude that
one cannot ``leave QM'' by invoking {\it more general linear dynamics than that
generated by trace preserving, completely positive maps} without affecting properties and
interpretation of the states represented by density matrices.}~\footnote{It will be
interesting to see how (any form of) hybrid dynamics, cf. Section~3., fares in
this respect.}
It is useful to recall here (the motivation behind)
the assumptions made concerning the
properties and interpretation of density operators. --
A density operator $\hat\rho$ is required to be {\bf i)} {\it Hermitian},
i.e., to have real eigenvalues, in order to qualify as an observable. This
is needed, in turn, if one requires $\hat\rho$ to be {\bf ii)} {\it positive-semidefinite}
and {\bf iii)} {\it normalized}. All three properties, together, are necessary
assumptions for the standard probability interpretation of the eigenvalues of a
density matrix, according to the Born rule.
A recurrent theme, when comparing quantum with classical states, for example,
with the help of the Wigner (function) transform of the density matrix, is the
appearance of negative eigenvalues or {\it ``negative probabilities''}.
Numerous attempts have been made to give a satisfactory physical interpretation
and mathematically consistent definition
to these, see Refs.\,\cite{Khrennikov,Mueckenheim,Burgin} and further references there.
Instead of entering this discussion, we look at the arguably simplest
example of {\it general linear dynamics} in the following. This is obtained by giving
up the requirement {\bf ii)} above, i.e., by abandoning positivity, and
by considering a two-dimensional Hilbert space.
\subsection{General linear $2\times 2$ dynamics: a model}
A convenient parametrization of the most general Liouville
superoperator $\hat{\cal L}$ for a two-dimensional state space
can be written with the help of the Pauli matrices $\vec\sigma$.
Thus, the (super)matrix elements ${\cal L}_{ij,kl}$, $i,j,k,l=1,2$, are defined by:
\begin{equation}\label{Ltwobytwo}
{\cal L}_{ij,kl}:=i{\cal G}_{\mu\nu}
\langle i|a_\mu+\vec b_\mu\cdot\vec\sigma |k\rangle
\langle l|a_\nu+\vec b_\nu\cdot\vec\sigma |j\rangle
\;\;, \end{equation}
where a summation over repeated indices is understood; in this notation
we have $(\hat{\cal L}\hat\rho )_{ij}={\cal L}_{ij,kl}\rho_{kl}$. Hermiticity requires
${\cal G}_{\mu\nu}={\cal G}_{\nu\mu}^\ast$. Generally, three vectors
$\vec b_{\mu =0,1,2}\in {\mathbf R^3}$ are needed. Furthermore, we define:
\begin{equation}\label{G}
{\cal G}:=\left ( \begin{array}{ccc}
1 & 0 & 0 \\
0 & g_{11} & g_{12} \\
0 & g_{12}^\ast & g_{22}
\end{array} \right )
\end{equation}
i.e., where $g$ is a Hermitian $2\times 2$ matrix. This Ansatz saturates the expected
number of $N^4=2^4=16$ real parameters, since $a_\mu$, $\vec b_\mu$, and $g$ contribute
3, 9, and 4 parameters, respectively. In order that the map generated by $\hat {\cal L}$
be trace preserving, we must have $0\stackrel{!}{=}i\partial_t\mbox{Tr}\hat\rho
=\mbox{Tr}\hat {\cal L}\hat\rho$, for all $\hat\rho$, which leads to
${\cal L}_{ii,kl}\stackrel{!}{=}0$. This yields four real constraints,
\begin{eqnarray}\label{C1}
a\cdot {\cal G}\cdot a+\vec b\cdot {\cal G}\cdot\vec b&\stackrel{!}{=}&0
\;\;, \\ [1ex] \label{C2}
2a\cdot\mbox{Re}{\cal G}\cdot\vec b+ig_{ij}\vec b_i\times\vec b_j&\stackrel{!}{=}&0
\;\;, \end{eqnarray}
which reduce the number of available real parameters to twelve.
Next, we conveniently choose the vectors $\vec b_{0,1,2}$ to form a right-handed
orthogonal system; with respect to suitable coordinates, this sets six vector
components to zero. Thus, we are left with six real parameters, three of which
could characterize the QM evolution of a two-state object, while three pertain
to the generalization we are concerned with.
Finally, we assume for simplicity
that the `QM part' of $\hat {\cal L}$ is diagonal,
corresponding to a spin-1/2 particle in a constant external magnetic field
parallel to the quantization axis, for
example.
Setting one of the remaining parameters to zero, we obtain a simplified model
with altogether three real parameters,
$\alpha ,\beta ,\gamma\in\mathbf{R}$:
\begin{eqnarray}
{\cal L}_{ij,kl}&:=&\alpha\Big\{\langle i|\sigma_z|k\rangle\delta_{lj}
-\delta_{ik}\langle l|\sigma_z|j\rangle\Big\}
\nonumber \\ [1ex]
&\;&+\beta\gamma\Big\{\langle i|\sigma_z|k\rangle\langle l|\sigma_x|j\rangle
-\langle i|\sigma_x|k\rangle\langle l|\sigma_z|j\rangle\Big\}
+i\beta\gamma\Big\{\langle i|\sigma_y|k\rangle\delta_{lj}
+\delta_{ik}\langle l|\sigma_y|j\rangle\Big\}
\nonumber \\ [1ex] \label{Lsimp}
&\;&+i\beta^2\delta_{ik}\delta_{lj}
+i\gamma^2\langle i|\sigma_y|k\rangle\langle l|\sigma_y|j\rangle
-i(\beta^2+\gamma^2)\langle i|\sigma_z|k\rangle\langle l|\sigma_z|j\rangle
\;\;, \end{eqnarray}
where QM terms $\propto\alpha\sigma_z$ feature in the first line.
All other contributions on the right-hand side have no counterpart in
QM; this will become obvious by explicitly solving the model.
The first two terms in the second line and the two last ones in the third couple
bra- and ket-states,
which we discussed as a consequence of classical evolution in
Sections~2.1 and 3.1. It is easy to verify that
the generator $-i\hat {\cal L}$ is Hermitian and trace preserving, as it should.
Taking hermiticity and trace normalization of the density matrix into account by:
\begin{equation}\label{rhoparam}
\hat\rho\equiv
\left ( \begin{array}{cc}
\rho_{11} & \rho_{12} \\
\rho_{12}^\ast & 1-\rho_{11}
\end{array} \right )
\;\;, \end{equation}
we write the resulting evolution equation explicitly:
\begin{eqnarray}\label{GLevol}
&\;&i\partial_t\hat\rho\; =\;\hat{\cal L}\hat\rho \; =\;
\\ [1ex] \nonumber
&\;&\left ( \begin{array}{cc}
i\gamma^2(1-2\rho_{11}) & 2\beta\gamma +(2\alpha +i[2\beta^2 +\gamma^2])\rho_{12}
-i\gamma^2\rho_{12}^\ast \\
-2\beta\gamma -(2\alpha -i[2\beta^2 +\gamma^2])\rho_{12}^\ast
-i\gamma^2\rho_{12} & -i\gamma^2(1-2\rho_{11})
\end{array} \right )
. \end{eqnarray}
Note that for $\beta,\gamma\rightarrow 0$, we recover the von\;Neumann equation, e.g., for
a spin-1/2 particle in a constant magnetic field. In this case, besides the Hamiltonian,
$\hat H:=\alpha\sigma_z$, also the
total spin-squared $S^2:=\vec\sigma^2/4\propto\mathbf{1}$ is a constant of motion.
\subsection{Constants of motion}
We remark that if a constant operator,
$\hat C$, obeys $\mbox{Tr}(\hat C\hat\rho (0))=\mbox{Tr}(\hat C\hat\rho (t))$,
for all solutions $\hat\rho$ of the general linear evolution equation (\ref{linform1}),
then $C_{ij}{\cal L}_{ji,kl}=0$, which generalizes the vanishing of the
commutator $[\hat C,\hat H]_{lk}$ in QM, for constants of motion. This is fulfilled
for any $\hat C\equiv c\mathbf{1}$, with constant $c$, due to the preservation of the trace
normalization of the density matrix, incorporated by $\hat{\cal L}_{ii,kl}\stackrel{!}{=}0$.
However, this raises also the interesting question, whether there exist conserved superoperators,
$\hat Q$, which can be defined by the vanishing supercommutator:
\begin{equation}\label{superC}
[\hat Q,\hat{\cal L}]_{ij,mn}:=Q_{ij,kl}{\cal L}_{kl,mn}-{\cal L}_{ij,kl}Q_{kl,mn}
\stackrel{!}{=}0
\;\;. \end{equation}
In the present case, this amounts to a $4\times 4$ matrix equation, to be studied.
\subsection{Solution of the $2\times 2$ model}
We now turn to the explicit solution of Eq.\,(\ref{GLevol}), which can be represented
as follows. The matrix elements are:
\begin{eqnarray}\label{rho11}
\rho_{11}(t)=1-\rho_{22}(t)&=&\frac{1}{2}
+\Big (\rho_{11}(0)-\frac{1}{2}\Big )\mbox{e}^{-2\gamma^2t}
\;\;, \\ [1ex] \label{rho12}
\rho_{12}(t)=\rho_{21}^\ast (t)&=&
r_+\mbox{e}^{\Omega_+t}+r_-\mbox{e}^{\Omega_-t}+r_c
\;\;, \end{eqnarray}
where $r_{\pm}$ are determined by initial conditions.~\footnote{Interestingly, the
underlying equation for the off-diagonal matrix elements,
for $\gamma\neq 0$ in particular, is of second order;
it decouples into two first order equations in the limit $\gamma\rightarrow 0$,
independently of the value of $\beta$.}
The remaining constants
are defined by:
\begin{eqnarray}\label{Rc}
r_c&:=&-\beta\gamma\frac{\alpha -i\beta^2}{\alpha^2+\beta^2(\beta^2+\gamma^2)}
\;\;, \\ [1ex] \label{Omega}
\Omega_\pm&:=&2\beta^2+\gamma^2
\pm 2i\alpha (1-\gamma^4/4\alpha^2)^{1/2}
\;\;, \end{eqnarray}
and we may choose $\alpha\geq 0$.
Diagonalizing the resulting traceless Hermitian density matrix,
we obtain the eigenvalues:
\begin{equation}\label{eigenval}
\rho_\pm (t)=\frac{1}{2}
\pm\Big ((\rho_{11}(0)-1/2)^2\mbox{e}^{-4\gamma^2t}+|\rho_{12}(t)|^2\Big )^{1/2}
\;\;, \end{equation}
with $|\rho_{12}|^2$ given by Eqs.\,(\ref{rho12})--(\ref{Omega}).
Several features of the solutions of our model seem unexpected. --
To begin with, for $\beta,\gamma\rightarrow 0$, we obtain:
\begin{eqnarray}\label{rho11s}
\rho_{11}(t)=1-\rho_{22}(t)&=&\rho_{11}(0)
\;\;, \\ [1ex] \label{rho12s}
\rho_{12}(t)=\rho_{21}^\ast (t)&=&
r_+\mbox{e}^{2i\alpha t}+r_-\mbox{e}^{-2i\alpha t}
\;\;, \end{eqnarray}
which amounts to the usual QM solution only, if we choose $r_+\equiv 0$.
From a QM perspective, the {\it larger set of solutions}, parametrized by a
larger set of initial conditions, is quite surprising. It leads to
time dependent `probabilities', if we try to interpret the eigenvalues
of $\hat\rho$ in the usual way:
\begin{equation}\label{probabs}
\rho_\pm (t)=\frac{1}{2}
\pm\Big ((\rho_{11}(0)-1/2)^2
+|r_+|^2+|r_-|^2+2\mbox{Re}(r_+r_-^\ast\mbox{e}^{4i\alpha t})\Big )^{1/2}
\;\;. \end{equation}
While these `probabilities' are real and normalized, by
construction of our model, they might temporarily fall outside
the interval $[0,1]$, depending on the parameters $r_\pm$.
This remark
applies also for the general case, with $\beta,\gamma\neq 0$, where amplitudes
grow $\propto\exp (2\beta^2+\gamma^2)t$, see Eqs.\,(\ref{rho12}),(\ref{Omega}).
The effect of such anomalous `probabilities' is clearly seen in
the following mean square deviations sensitive to fluctuations.
With $\langle\hat O\rangle :=\mbox{Tr}(\hat O\hat\rho )$,
$\sigma_{x,y,z}^2=\mathbf{1}$, and
$\langle\sigma_x\rangle =2\mbox{Re}(\rho_{12})$, $\langle\sigma_y\rangle =-2\mbox{Im}(\rho_{12})$,
$\langle\sigma_z\rangle =2\rho_{11}-1$, we obtain:
\begin{eqnarray}\label{x}
\Delta_x^2:=
\langle\sigma_x^2\rangle -\langle\sigma_x\rangle^2&=&1-\Big (2\mbox{Re}(\rho_{12})\Big )^2
\;\;, \\ [1ex] \label{y}
\Delta_y^2:=
\langle\sigma_y^2\rangle -\langle\sigma_y\rangle^2&=&1-\Big (2\mbox{Im}(\rho_{12})\Big )^2
\;\;, \\ [1ex]\label{z}
\Delta_z^2:=
\langle\sigma_z^2\rangle -\langle\sigma_z\rangle^2&=&4\rho_{11}(1-\rho_{11})
\;\;. \end{eqnarray}
While $\Delta_z^2\stackrel{t\rightarrow\infty}{\longrightarrow}1$, for $\gamma\neq 0$,
which would correspond to a completely mixed state in QM, $\Delta_{x,y}^2$ can have
oscillatory contributions with growing amplitude
$\propto\exp\Omega_\pm t$,
see Eqs.\,(\ref{rho11})--(\ref{Omega}). -- Also note that the would-be
QM energy expectation varies in time,
$\alpha\langle\sigma_z\rangle =\alpha (2\rho_{11}(0)-1)\exp (-\gamma^2t)$, approaching
the stationary mixed-state value zero.
To summarize: {\it Modifying a QM two-state model
by applying a general linear perturbation, however small --
which preserves hermiticity and trace normalization of the
density matrix --
opens the possibility of a larger state space, reflected in a
doubling of the number of degrees of
freedom}.~\footnote{I.e., a doubling of the initial conditions for the
off-diagonal matrix elements here; it will be interesting to see, whether this extends
to all degrees of freedom in a more general model.}
We emphasize that the QM evolution becomes exponentially unstable when
such perturbation is introduced.
Generically, this spoils the standard interpretation of the eigenvalues
of the density matrix as probabilities and can
result in interesting oscillatory effects, as we have seen.~\footnote{One might
speculate on the relevance for flavour oscillations.}
It is an interesting question,
whether there is a ``classical'' formulation of general linear dynamics,
suggested by a doubled number of certain degrees freedom and ``anomalous probabilities''
that seem invariably to appear -- and which remind of similar phenomena
encountered when relating quantum and classical mechanics, e.g., via the
Wigner function.
\section{Conclusions}
In this article we have summarized our earlier derivation of a {\it path integral}
for classical Hamiltonian systems based on an ensemble description \cite{EGV10,FabThesis}.
This leads us to point out the correlation
properties of classical dynamics in parallel to the quantum mechanical
ones and to identify characteristic similarities and differences. In particular,
it seems useful to distinguish {\it intra- and interspace entanglement}.
The former has been held characteristic of quantum mechanics and a feature of
superpositions of tensor product states.
The latter concerns classical mechanics only; it correlates
Hilbert space states and their duals. However, surprisingly, for anharmonic
potentials or interactions, classical mechanics additionally shows intraspace entanglement,
as in quantum mechanics.
As a first application, we propose a new formulation of {\it hybrid dynamics},
i.e., based on a hypothetical direct coupling between quantum and classical objects.
This may be of practical as well as foundational interest.
Finally, we study a generalization of quantum evolution,
{\it general linear dynamics}, where the evolution is generated
by a superoperator that preserves hermiticity and trace normalization of
density matrices. We argue that one cannot ``leave QM'' without giving up
one of the three defining properties of density matrices, to be Hermitian,
normalized, and positive-semidefinite.
In the most simple case of a two-state system, we solve such dynamics explicitly.
We show that the corresponding von\,Neumann dynamics becomes
{\it exponentially unstable} under the influence of a general linear perturbation.
Most interestingly, it leads to the appearance of ``anomalous probabilities''
and an enlargement of the state space, possibly pointing towards a
sort of prequantum dynamics. -- We intend to study more complex
objects consisting of such two-state systems as
building blocks which interact. This may be useful in
trying to understand quantum mechanics as an emergent phenomenon \cite{tHooft10}.
\ack{We thank M.J.~Everitt, F.~Finster, A.~Khrennikov, V.I.~Man'ko, and T.~Padmanabhan
for discussions.}
\section*{References}
|
\section{Introduction}
One usually thinks of Mathematics as a precise discipline,
often confusing mathematical rigor with logical formality.
In fact, most mathematics is simply too informal to be directly handled
by the logical and algebraic means offered by interactive or automated
theorem provers.
The typical mathematical discourse systematically exploits symbol
overloading and notational abuses that can hardly be understood by
automatic devices without substantial help from the user side,
that is one of the reasons why formal encoding is so expensive
and frustrating.
The crucial point is that the intrinsic ambiguity of the mathematical
vernacular can only be resolved by a sufficiently
contextual interpretation, requiring not just the knowledge of its
specific notation and conventions, but nontrivial skills in the given
mathematical discipline. A large part of this background knowledge is
expressed in form of equalities and isomorphisms, allowing a mathematician
to freely move between different incarnations (intensions) of the same entity
without even mentioning the transformation. Providing ITP-systems with
similar capabilities seems to be a major way to improve their intelligence,
and to ease the communication between the user and the machine. In the
present paper, we discuss our experience of integration of a superposition
calculus within the Matita interactive prover, providing in particular
a very flexible, ``smart'' application tactic, and a simple, innovative
approach to automation.
The need for a stronger integration between fully automatic
(resolution) provers and interactive ones, and more generally for
a stronger automation support in proof assistants is a major
challenge (see e.g. \cite{harrison-book})
and many efforts have been already done in this
direction: for instance, KIV has been integrated with the tableau
prover $3T^AP$ \cite{AB98}; HOL has been integrated with various
first order provers, such as
Gandalf \cite{Hurd99} and Metis \cite{metis};
Coq has been integrated with Bliksem \cite{BHN02};
Isabelle was first
integrated with a purpose-built prover \cite{blast} and more recently with
Vampire \cite{MQP06}.
We share most of the principles guiding these
efforts, and in particular the need to refer to a large library of
known lemmas, and the goal to deliver a checkable proof, in conformity
with the trusted kernel philosophy (sometimes referred to as De Bruijn
principle) inspiring most interactive provers.
However, there are two
different uses of automation that have different requirements
and possibly deserve different approaches and solutions.
The first one (small scale automation)
is to relieve the user from the burden to fill in
relatively trivial steps, by automatically completing the missing
gaps. This kind of automation must be {\em fast}, {\em robust}
and sufficiently {\em predictable}, in the sense that the automation
procedure should not miss simple solutions when they exist.
In general, in this case, the user is not interested to read
back the proof and there is no point in trying to transform it
in a human readable format.
The second use of automation (large scale automation)
is to really help the user in the process
of {\em devising} the proof. In spite of all the progresses in this
field, a fully automatic approach still looks highly problematic: a
more promising approach seems to be that of improving the
cooperation between human and machine, and in particular of
making a better profit of the machine's combinatorial capabilities.
At present, ``interaction'' in ITP systems is essentially restricted to
a master-slave command execution loop, that frustrates the computational
power of machines. A better repartition of the work could consist in
leaving to the user the most intelligent tasks, such as identifying
the key lemmas, proof principles and intermediate results of interest
for the proof, assigning to the system the burden of composing them
into a coherent proof (eventually exploiting a huge library of known
results). The user has hence the responsibility to cut the search
space, while the machine is supposed to automatically and
systematically explore it (this was also our guiding idea behind
the design of the automation driver of Matita version 0.5.x
\cite{auto-driver}). Another recent system close to this conception is
$\Omega$mega \cite{omega}: in this case the user drives the
search by means of proof plans, invoking external reasoners
to fullfill them.
For large scale automation, we could bear
to run time consuming jobs, possibly working for hours in the background.
The result is completely unpredictable, and probably unstable.
The system should produce a proof in a format as readable as possible,
since the user is surely interested to inspect it himself, apart from
pasting it into the proof script (if proof searching is expensive
we probably wish to avoid running it over and over
every time we recompile the script, or at least to have
the possibility to choose between these two alternatives).
The heuristic, unstable nature of complex automation procedures, combined
with the verbose nature of fully formal proofs, naturally suggest
the idea of producing ``proof traces'' as a compact, readable and
reproducible output of automation devices.
This is the point where the two kinds of automation recombine together,
since the execution of proof traces precisely requires
small scale automation capabilities. Our point is that a good part
of these capabilities are fulfilled by reasoning up to equalities,
providing the {\em connective glue} that constitutes most
of that background knowledge tacitly, almost unconsciously used in
the typical mathematical reasoning. Superposition provides a
natural support for reasoning {\em modulo} a congruence on
propositions, implementing ideas similar to \cite{modulo},
and providing a flexible and powerful tool for small scale
automation.
One of the components of the Matita interactive theorem prover
is a state of the art, first order, untyped superposition algorithm,
able to compete with the best tools currently
available: in particular, our system scored in fourth position
in the unit equality division at the 22nd CADE ATP System Competition,
beating a glorious
system such as Otter, and being awarded as the best
new entrant tool \cite{Sutcliffe09}.
Note in particular that
Matita is entirely written in a functional language (OCaml), while most
ATP system (with the relevant exception of Metis, that was however
beaten by Matita) are written in imperative code.
In this paper, we shall provide a theoretical and
architectural description - as self contained as possible -
of this tool (Section \ref{sec:superposition});
hence we shall discuss its integration
with Matita (Section \ref{sec:integration}), and show some
of its applications, mostly aimed to improve the intelligence
of commands (smart application,
Section \ref{sec:applyS}) and the overall automation of the system
(Section \ref{sec:auto}).
\section{Superposition}\label{sec:superposition}
Techniques for equational reasoning are a key component in many
automated theorem provers and interactive proof and verification
systems \cite{BG94,paramodulation,equality-handbook}.
The main deductive mechanism is a {\em completion}
technique \cite{Knuth-Bendix} attempting to transform a given set of
equations into a confluent rewriting system so that any two terms
are equal if and only if they have identical normal forms.
Not every equational theory can be presented as such a confluent
rewriting system, but you may progressively approximate it
by means of a refutationally complete method called
{\em ordered completion}.
The deductive inference rule used in completion procedures is
{\em superposition}
which consists of first unifying one side of one equation with a subterm
of another, and hence rewriting it with the other side;
the selection of the two terms to be unified is guided by a given
term ordering, which imposes certain restrictions on inferences, with the
major benefit to prune the search space. All results in this section
are known, and we only report them for the sake of completeness.
\subsection{Preliminaries}
Let $\mathcal{F}$ bet a countable alphabet of functional symbols, and
$\mathcal{V}$ a countable alphabet of variables.
We denote with $\mathcal{T}(\mathcal{F},\mathcal{V})$ the set of terms
over $F$ with variables in $V$. A term
$t\in \mathcal{T}(\mathcal{F},\mathcal{V})$ is either a 0-arity element of
$\mathcal{F}$ (constant), an element of $\mathcal{V}$ (variable),
or an expression of the form $f(t_1, \dots, t_n)$ where $f$ is a element of
$\mathcal{F}$ of arity $n$ and $t_1, \dots, t_n$ are terms.
Let $s$ and $r$ be two terms.
$s|_p$ denotes the subterm of $s$ at position $p$ and $s[r]_p$
denotes the term $s$ where the subterm at position $p$ has been
replaced by $r$.
A substitution is a mapping from variables to terms.
Two terms $s$ and $t$ are unifiable if there exists a substitution
$\sigma$ such that $s\sigma = t\sigma$. Moreover, in the previous case,
$\sigma$ is called a most general unifier (mgu) of $s$ and $t$ if
for all substitution $\theta$ such that
$s\theta = t\theta$, there exists a substitution $\tau$ which satisfies
$\theta = \tau \circ \sigma$.
A literal is either an abstract predicate (represented by a term),
or an equality between two terms. A clause $\Gamma \vdash \Delta$
is a pair of multisets of literals: the negative literals $\Gamma$,
and the positive
ones $\Delta$. If $\Gamma = \emptyset$ (resp. $\Delta = \emptyset$),
the clause is said to be positive (resp. negative).
A Horn clause is a clause with at most one positive literal.
A unit clause is a clause
composed of a single literal. A unit equality is a unit clause
where the literal is an equality.
\subsection{Term orderings and Inference rules}
A strict ordering $\prec$ over $\mathcal{T}(\mathcal{F},\mathcal{V})$
is a transitive and irreflexive (possibly
partial) binary relation. An ordering is {\em stable} under substitution
if $s \prec t$ implies $s\sigma \prec t\sigma$ for all terms $t, s$ and
substitutions $\sigma$. A well founded monotonic ordering stable under
substitution is called {\em reduction ordering} (see \cite{Dershowitz82}).
The intuition behind the use of reduction orderings for
limiting the combinatorial explosion of the number of equations during
inference, is to only rewrite big terms to smaller ones.
\begin{description}
\item[superposition left] This rule defines backward reasoning steps. The
equational fact $l = r$ is combined with the goal $t_1 = t_2$ obtaining a
the goal $(t_1[r]_p = t_2)\sigma$.
\begin{displaymath}
\frac{
\vdash l = r \quad\quad t_1 = t_2 \vdash
}{
(t_1[r]_p = t_2 \vdash)\sigma
}
\end{displaymath}
if $\sigma = mgu(l, {t_1}|_p)$, $t_1|_p$ is not a variable, $l\sigma
\not\preceq r\sigma$ and $t_1\sigma \not\preceq t_2\sigma$;\\
\item[superposition right] This rule defines forward reasoning steps. The two
equational facts $l = r$ and $t_1 = t_2$ obtaining a new fact
$(t_1[r]_p = t_2)\sigma$.
\begin{displaymath}
\frac{
\vdash l = r \quad\quad \vdash t_1 = t_2
}{
(\vdash t_1[r]_p = t_2)\sigma
}
\end{displaymath}
if $\sigma = mgu(l, {t_1}|_p)$, $t_1|_p$ is not a variable, $l\sigma
\not\preceq r\sigma$ and $t_1\sigma \not\preceq t_2\sigma$;\\
\item[equality resolution] This is the rule that ends the proof search.
\begin{displaymath}
\frac{
t_1 = t_2 \vdash
}{
\Box
}
\end{displaymath}
if there exists $\sigma = mgu(t_1, t_2)$.
\end{description}
\subsection{Simplification rules}
For efficiency reasons, the calculus must be integrated with a few
additional optimization rules, the most important one being
demodulation (\cite{demodulation}):
\begin{description}
\item[subsumption] This rule allows to identify and drop clauses that happen to
be the instance of more general ones, and are thus superflous.
\begin{displaymath}
S \cup \{C, D\} \Rightarrow S \cup \{C\}
\end{displaymath}
if $C$ \emph{subsumes} $D$, i.e. if there exists a substitution $\sigma$
such that $C\sigma \equiv D$.\smallskip
\item[tautology elimination] This rule eliminates equational facts that
are provable with the \textbf{equality resolution} rule and are thus superflous.
\begin{displaymath}
S \cup \{\vdash t = t\} \Rightarrow S.
\end{displaymath}
\item[demodulation] This rule aims at reducing the size of the clauses involved
in the proof search, speeding up all operations whose complexity is determined
by the size of the terms involved. The intuitive idea is to consider clauses
modulo know equational facts and record only their smaller representative.
\vspace*{-.1cm}
\begin{displaymath}
S \cup \{\vdash l = r, C\} \Rightarrow S \cup
\{\vdash l = r, C[r\sigma]_p\},
\end{displaymath}
if $l\sigma \equiv C|_p$ and $l\sigma \succ r\sigma$.
\end{description}
\subsection{The main algorithm}
\label{sec:main_algorithm}
To avoid combining the same clauses twice, it is convenient to keep
them in two sets, that are traditionally called {\em active} and
{\em passive}. The general invariant is that clauses in the active
sets have been already composed together in all possible ways.
A step consists in selecting some clauses from the passive set,
add them to the active set, compose them with the current active set
- and thus with themselves - (inference),
and finally add the newly generated clauses to the passive set
(possibly after a simplification).
A natural strategy would consist in selecting the whole passive set
at each iteration, realizing a sort of breadthfirst strategy.
The advantage of this strategy is that it is very predictable, and
hence particularly easy to debug. Unfortunately the number of new
equations generated at each step grows extremely fast, practically
preventing to iterate the main loop more than a few steps.
To avoid this problem, the opposite solution is usually adopted,
consisting in selecting just {\em one} passive equation at each step.
The equation is selected according to suitable heuristics (size,
goal similarity, and so on), usually comprising some fairness
criterion to ensure completeness (we must ensure that any passive
equation will be selected, soon or later).
This approach
is called the \emph{given-clause algorithm} (figure \ref{mainloop}),
\begin{figure}[htp]
\begin{center}
\includegraphics[width=0.50\textwidth]{main_loop.pdf}
\caption{given-clause algorithm\label{mainloop}}
\end{center}
\end{figure}
and it is the procedure used (with some variations)
by all modern theorem provers (see e.g. \cite{vampire_annals}).
The advantage of this method
is that the passive set grows much slower, allowing a more focused
and deeper inspection of the search space. The drawback is that
the algorithm becomes extremely sensitive to the selection heuristic,
leading to more unpredictable behaviour.
\smallskip
In order to get a high performance tool, the given clause algorithm has
to be tuned and optimized in several ways. The critical areas are:
\begin{itemize}
\item Data structures and code optimization
\item Orderings used to orientate rewriting rules
\item Selection strategy
\item Demodulation
\end{itemize}
We are currently using relatively
simple data structures (discrimination
trees \cite{McCune}) for term indexing,
but we plan to exploit in the near future more specific data structures
(such as substitution \cite{Graf95} or context trees \cite{codedcontexttrees}).
On complex problems (e.g. problems in the TPTP library with rating
greater then $0.30$) the choice of a good ordering for inference rules
is of critical importance.
As we already mentioned, we have implemented several orderings,
comprising standard Knuth-Bendix (KBO), non recursive Knuth-Bendix
(NRKBO), lexicographic path ordering (LPO) and recursive path ordering (RPO).
The best suited ordering heavily depends on the kind of problem,
and is hard to predict\footnote{Our approach to the CADE ATP System Competition
was to run in parallel different processes with different orderings.}.
Luckily, on simpler problems (of the kind required for small
scale automation) the given-clause algorithm is less sensitive
to the term-ordering, and any of them usually produce a solution in a
reasonable amount of time.
The selection strategy currently implemented by Matita is a based on
combination of age and weight. The weight is a positive integer that
provides an estimation of the ``complexity'' of the
clause, and is tightly related to the number of occurrences of symbols
in it.
Another important issue for performance is demodulation: the given clause
algorithm spends most of its time (up to 80\%) in simplification, hence
any improvement in this part of the code has a deep impact on performance.
However, while reductions strategies, sharing issues and abstract machines
have been extensively investigated for lambda calculus (and in general
for left linear systems) less is known for general first
order rewriting systems.
In particular, while an innermost (eager)
reduction strategy seem to work generally better than an outermost
one (especially when combined with lexicographic path ordering), one
could easily create examples showing an opposite behaviour (even
supposing to always reduce needed redexes).
Although we did not want to focus too much on developing specific
heuristics, two widespread techniques, not yet implemented, would
still be of great interest. The
first one is Limited Resource Strategy \cite{LRS}, which basically allows
the procedure to skip some inference steps if the resulting clauses
are unlikely to be processed, because of a lack of time or memory.
The other promising technique is indexing
modulo associativity and commutativity \cite{assoccommut}, which is often
heavily used when working on algebraic structures.
\section{Integrating superposition with Matita}
\label{sec:integration}
\subsection{Library management}
A simple possibility for integrating superposition with Matita
is simply to solve goals assuming as initial passive set all
equational facts in the library (plus the equations in the
local context). The main drawback of this approach is that
passive equations would be selected slowly, and in a quite
{\em repetitive way} every time a new problem is met. In fact, superposition
right, as any forward operation, only concern {\em facts}, and
apart from the local hypothesis, most of these facts are known
in advance. This suggest that, in ITP systems, forward operations
should be processed, as much as possible, off line;
but then we have to face the
dual problem, namely to avoid an unnecessary proliferation of
results, polluting the library (and the memory) with trivialities.
The compromise adopted in Matita was suggested
by the observation that, in a given-clause algorithm,
{\em selection} is a conspicuous operation
requiring an intelligent choice; but all theorems in the library
are indeed already a ``selected'' subset (otherwise, there would be
no point to record them). In other words, the idea is
to use the unit equalities in the library not as initial passive set,
but as the {\em active} one. This means that every time a new
equality is added to the library it also goes through
one cycle of the given-clause algorithm as if it was the newly
selected passive equation: it is composed (after simplification)
with all existing active equations (that is, up to simplifications,
all previously proved equalities), and the newly created equations
are added to the passive list\footnote{This approach is particularly
important in view of the fact that, typically, the passive set is not
even used for demodulation.} This way, we have a natural,
simple but traceable syntax to drive the saturation process: it is
enough to explicitly list in the library the selected equation.
At the same time, this approach reduces the verbosity of the library, since
trivial results generated by superposition in the passive list
may be used without the need to declare (and name) them explicitly.
\subsection{Input/Output}
The communication between Matita and the superposition tool is
not precise. As we already said, our superposition algorithm is
first order and untyped; instead of attempting a complex encoding
of the Calculus of Inductive Constructions (CIC) in first order logic
(that is the approach adopted e.g. in \cite{MP08}), we prefer to
use a naive, but efficient translation, possibly losing information.
We shall then try to automatically reconstruct the missing information
during proof reconstruction, exploiting the sophisticated inference
capability of the Matita {\em refiner} \cite{hints}.
As a consequence, automation is a best effort service:
not only it may obviously fail to produce a proof, but sometimes
it could produce an argument that the system is not able to refine correctly
(independently from the fact if the delivered proof was ``correct''
or less).
Although there is no particular problem to implement a typed
superposition algorithm, or even embedding types as first order
terms (in more or less naive ways, according to the way
we wish to take convertibility into account), for performance
reasons we decided to work with completely untyped terms.
In particular, equations $r =_T s$ of the calculus of constructions
are translated to first order equations by merely following the
applicative structure of $r$ and $s$, and translating
any other subterm into an opaque constant. The type $T$ of the equation
is recorded, but we are not supposed to be able to compute types
for subterms.
Since all equations are combined together via superposition rules, there is
a (modest) risk of producing ``ill-typed'' terms.
Consider for instance the superposition left rule (the reasoning is similar
for the other rules)
\begin{displaymath}
\frac{
\vdash l = r \quad\quad t_1 = t_2 \vdash
}{
(t_1[r]_p = t_2 \vdash)\sigma
}
\end{displaymath}
where $\sigma = mgu(l, {t_1}|_p)$ and $l\sigma \not\preceq r\sigma$.
The risk is that $t_1|_p$ has
a different type from $l$, hence resulting into an illegal rewriting
step. Note however that $l$ and $r$ are usually rigid terms, whose
type is uniquely determined by the outermost symbol. Moreover,
$t_1|_p$ cannot be a variable, hence they must share this outermost
symbol. If $l$ is not rigid, it is usually a variable $x$ and if
$x \in r$ (like e.g. in $x=x+0$) we have (in most orderings)
$l \preceq r$ that again rules out rewriting in the wrong direction.
This leads us to the following notion of {\em admissibility}.
We say that an applicative term $f(x_1,\dots,x_n)$ is {\em implicitly
typed} if its type is uniquely determined by the type of $f$.
We say that an equation $l = r$ is admissible if both $l$ and $r$
are implicitly typed, or $l \preceq r$ and $r$ is implicitly typed.
If an equation is not admissible, we forbid to take it into account
for superposition.
\subsection{(Re)construction of the proof term}\label{sec:proofs}
Reading back a superposition proof inside any interactive prover
is a relatively simple operation (just requiring rewriting),
and one of the reasons for sticking to this fragment.
In the superposition module, each proof step in encoded as a
tuple
\begin{verbatim}
Step of rule * int * int * direction * position * substitution
\end{verbatim}
where \verb+rule+ is the kind of rule which has been applied,
the two integers are the two $id's$ of the composing equations
(referring to a ``bag'' of unit clauses),
\verb+direction+ is the direction the second equation is applied
to the first one, \verb+position+ is a path inside the rewritten term and
finally \verb+substitution+ is the mgu required for the rewriting step.
\begin{figure}[htp]
\begin{center}
\includegraphics[width=0.5\textwidth]{readback.pdf}
\caption{given-clause algorithm\label{readback}}
\end{center}
\end{figure}
The proof has the shape depicted in Figure \ref{readback}, where all
superposition left steps are on the rightmost spine leading from the
Goal to the empty clause. Superpostion right steps are
forward rewriting operations and their translation is straightforward;
on the other side, superposition left steps must be reverted in order to build
a direct proof of the (suitably instantiated) goal from its refutation.
\noindent
Formally, let's call \verb+eq_ind+ the higher order
rewriting step
\[
eq\_ind: \forall A: \textsc{Type}. \forall x:A. \forall P: A \to
\textsc{Prop}.P~x \to \forall y: A. x = y \to P~y
\]
Let us consider a superposition right step
\[
\frac{
\vdash l =_A r \quad\quad \vdash t =_B s
}{
\vdash t[r]_p\sigma =_B s\sigma
}
\]
If $h:l =_A r$ and $k:t =_B s$ then
\[eq\_ind ~A~l\sigma~ (\lambda x:A.t[x]_p =_B s)\sigma~k\sigma~
r\sigma~h\sigma : t[r]_p\sigma =_B s\sigma \]
Conversely, given a superposition left step
\[
\frac{
\vdash l =_A r \quad\quad \alpha:t =_B s\vdash
}{
t[r]_p\sigma =_B s\sigma \vdash
}
\]
if $h:r =_A l$ and $k: t[r]_p\sigma =_B s\sigma$ then
\[eq\_ind ~A~r\sigma~ (\lambda x:A.t[x]_p =_B s)\sigma~
k~l\sigma~h\sigma : t\sigma =_B s\sigma
\]
To generate a CIC proof term, clauses are topologically sorted
w.r.t. their dependendices (to respect scoping),
their free variables are explicitly quantified, and nested let-in
patterns are used to build the proof.
A delicate point of the translation is closing each clause
w.r.t. its free variables, since we should infer a type for them.
The simplest solution is to generate so called ``implicit'' arguments
leaving to the Matita {\em refiner}~\cite{hints}, the burden of guessing them.
For instance, superposing $plusC: x + y = y \underline{+} x$ with
$plusA: x + (y + z) \stackrel{\leftarrow}{=} (x + y) + z$ at the underlined
position and in the given direction
gives rise to the following piece of code, where question marks stands
for implicit arguments:
\begin{lstlisting}
let clause_59:
$\forall x183: ?.$
$\forall x184: ?.$
$\forall x185: ?.$
$x183 + (x184 + x185)) = x184 + (x185 + x183)$
$:= \lambda x183:?.$
$\lambda x184:?.$
$\lambda x185:?.$
eq_ind$nat\; ((x184 + x185) + x183)$
$(\lambda x:nat. x183 + (x184 + x185) = x)$
(plusC $x183 (x184 + x185)) (x184 + (x185 + x183)$)
(plusA $x184 x185 x183$) in
...
\end{lstlisting}
\section{Smart applications}
\label{sec:applyS}
The first interesting application of superposition (apart its
use for solving equational goals), is the implementation of a more
flexible application tactic. As a matter of fact, one of the most
annoying aspects of formal development is the need of transforming
notions to match
a given, existing result. As we already said, most of these
transformations are completely transparent to the typical mathematical
discourse, and we would like to obtain a similar behaviour in interactive
provers.
Given a goal $G$ and a theorem t: $\Gamma \to A$, the goal is to try
to match $A$ with $G$ up to the available equational knowledge base, in
order to apply $t$. We call it, the {\em smart application} of $t$ to
$G$.
We use superposition in the most direct way, exploiting on
one side the higher-order features of CIC, and on the other
the fact that the translation to first order terms does
not make any difference between predicates and functions:
we simply generate a
goal $A = G$ and pass it to the superposition tool (actually,
it was precisely this kind of operation that motivated our original
interest in superposition). If a proof is found, $G$ is transformed
into $A$ by rewriting and $t$ is then
normally applied.
Superposition, addressing a typically undecidable problem,
can easily diverge, while we would like to have a reasonably
fast answer to the smart application invocation, as for any other
tactic of the system. We could simply
add a timeout, but we prefer to take a different, more predictable
approach. As we already said, the overall idea is that superposition
right steps - realising the {\em saturation} of the equational
theory - should be thought of as off line operations. Hence, at run
time, we should conceptually work as if we had a {\em
confluent} rewriting system, and the only operation worth to do
is {\em narrowing} (that is, left superposition steps). Narrowing
too can be undecidable, hence we fix a given number of narrowing
operations to apply to each goal (where the new goal instances generated at
each step are treated in parallel). The number of narrowing steps
can be fixed by the user, but a really small number is usually
enough to solve the problem if a solution exists.
\begin{example}
\label{example:smart}
Suppose we wish to prove that the successor function
is $\le$-reflecting,
namely
\[(*)\hspace{.5cm}\forall n,m. S n \le S m \to n \le m\]
Suppose we already proved that the predecessor function is monotonic:
\[monotonic\_pred: \forall n,m. n \le m \to pred\; n \le pred\; m\]
We would like to merely ``apply'' the latter to prove the former.
Unfortunately, this would not work, since there is no way to match
$pred\; X \le pred\; Y$ versus $n \le m$, unless {\em narrowing} the
former. By superposing twice with the equation
$\forall n. pred (S n) = n$ we
immediately solve our matching problem via the substitution
$\{X := S n, Y := S m\}$. Hence, the smart application of
$monotonic\_pred$ to the goal $n \le m$ succeeds, opening the new
goal $S n \le S m$ that is the assumption in $(*)$.
\end{example}
\begin{example}
\label{example:smart8}
Let us use the notation $A[B/i]$ to express the substitution of
$B$ for the $i-th$ free variable in $A$. The substitution lemma
says that for all $k,i$
\[A[B/i][C/i+k] = A[C/S(k+i)][B[C/k]/i]\]
(where $S$ is the successor function).
The idea is to prove the substitution lemma by structural induction
over $A$. Suppose now $A$ is a binder, e.g. a lambda term $FUN(M)$
where $M$ is the body of the function.
The definition of substitution tells us that
\[FUN(M)[B/i] = FUN(M[B/i+1])\]
Hence, after normalization and elimination of congruent terms,
we are left to prove\footnote{We added some artificial
parenthesis to the terms to emphasize the left associativity of
plus.}
\[(M[B/i+1][C/(k+i)+1] = Fun(M[C/S((k+i)+1)][B[C/k]/i+1]\]
under the inductive hypothesis
\[Hind: \forall j.M[B/i][C/k+j] = M[C/S(k+j)][B[C/k]/j]\]
It is evident that it is enough to instantiate $j$ with $i+1$
but in order to unify $(k+i)+1$ with $k+?_j$ we have to use the
associativity law for the sum! Hence the smart application of
$Hind$ succeeds where the normal application would fail.
\end{example}
\section{The auto tactic}
\label{sec:auto}
By itself, smart application is less interesting than expected.
The point is that, compared to the effort of {\em finding} the
``right'' theorem $t$ in the library, the work of transforming the goal
to match the conclusion is a boring, but minor task.
What is really interesting, instead, is the possibility to combine
smart application with a goal-oriented proof searching
technique, to achieve a cheap, simple but surprisingly effective
management of equality.
According to our philosophy, forward operations in ITP systems
should be performed off line, and explicitly or implicitly
recorded in the library (if a forward step is {\em really} useful
in some context, it is likely to be useful in other, similar
contexts as well, hence it is a very good candidate to explicitly
appear in the library). For this reason, the Matita automation
tool is backward-based (backward operations act on the goal, that
is only known at run time), essentially trying to build a proof
by a repetitive application of tactics. The proof we are looking for is not
in normal form: in fact, the most relevant tactic is application,
and the automation tool is supposed to explore the library for
all known results matching the current goal. In this respect,
automation resembles a prolog-like program, and we use a traditional
depth-first strategy (with bounded, user configurable depth)
to explore the proof space.
The main optimizations\footnote{All these optimizations destroy the so called
procedural interpretation of logic programs, and received
very little attention by the logic programming community.}
implemented are the following:
\begin{description}
\item[goal clustering] a {\em cluster} in a set of (conjunctive)
goals $\Delta = g_1,\dots,g_n$ is a minimal subset closed w.r.t. its
free variables: any variable appearing in a goal of the cluster
can only appear in other goals of the same cluster. Clusters
obviously form a partition of the original set;
their interest is that the processing of different clusters
can be separated by {\em green cuts};
\item[loop detection] if a goal $\Delta$ generates
another goal $\Delta'$ subsumed by the former, the proof branch can
be pruned\footnote{At present this is only implemented in case $\Delta'$
is a single literal.}(if we find a proof for $\Delta'$ it works
for $\Delta$ as well - recall that variables in goals are
existentially quantified).
\end{description}
We also plan to implement a failure cache (indexed by the failure
depth); instead, the advantage of caching successes
looks much more questionable (either we pre-compute the
whole success set, requiring a different proof searching
strategy, or we easily end-up duplicating solutions).
Smart application can be easily integrated in our automated
proof searching tactic. Per se, due to the severe constraints
imposed on superposition, smart application is not much slower than
normal application. The real problem is the brutal explosion in the
number of candidates. With normal application, using good data
structures for indexing the universe of known results (we use
discrimination trees \cite{McCune}), we are able to retrieve, for each goal,
a relatively small number of candidates. In the case of smart
application, any theorem predicating something ``similar'' to the goal
is a potential candidate. Our notion of similarity is particularly
weak: we look for any theorem whose conclusion shares with the goal
(possibly up to reduction) the same top predicate.
Note however that what really
matters from the complexity point of view is not the number
of candidates which are tried, but the number of them whose
application {\em succeeds}, giving rise to new branches in the
the search tree. Luckily, in general, smart application does not
sensibly enlarge the number of applicable theorems, and the overall
complexity remains feasible, especially for small depths (3/4).
\subsection{Proof traces}
Since most of the time is spent in searching the right theorems
composing the proof, a natural idea is to let the automation
tactic return a trace of the proof consisting of all
library results used to build the proof.
We omit the local assumptions, all equations used by superposition,
and to further reduce the verbosity of the trace,
we also omit all library {\em facts} (i.e. all results with no
hypothesis, hence appearing in leaf position inside the proof).
The resulting set is passed as an optional ``by'' argument to
the auto tactic.
If the argument is present, the automation tactic would use
the set passed as an argument as candidates for smart applications,
apart from at depth $0$, where facts in the whole library would be
taken into account. Local assumptions are always tried, too.
Using these simple proof traces automation becomes extremely fast,
and almost comparable to a fully expanded proof script.
\subsubsection{Example}
This is a relatively complex example borrowed from the Matita standard
library (in particular, in a contribution regarding lifting
and substitution in DeBrujin notation). The goal to prove is
$k \le n-1$ under the assumption $H: j + k < n$, where $j,k$ and $n$
are natural numbers. The relation $n < m$ is definitionally equivalent
to $S n \le m$ where $S$ is the successor function. Note that
the successor function is extensionally equal to (but does not
coincide with) the operation of adding 1, in the same way as the
predecessor function is extensionally equal to (but does not coincide with)
the operation of subtracting 1. Another delicate point is that
the minus operation $x - y$ returns $0$ when $y > x$, so
$S(x - 1) = x$ only if $x > 0$. \\
The solution automatically found
by Matita is depicted in Figure \ref{example_glue} (the picture
is better understood reading it from the bottom to the top):
\begin{figure}[htp]
\begin{center}
\includegraphics[width=0.5\textwidth]{example_glue.pdf}
\caption{Glueing together lemmas via rewriting\label{example_glue}}
\end{center}
\end{figure}
it first applies the monotonicity of the
predecessor function, passing from $k \le n-1$ to $S k \le n$; then
it applies the lemma
\[le\_plus\_to\_le: \forall n,m,a: n + a \le m \to n \le m\]
obtaining the goal $k + Z \le n$, and finally applies the
hypothesis $H$, instantiating $Z$ with $S\;j$. In order to do these
passages, the system exploits the following equivalences: $k = pred (S\;k)$, $n - 1 = pred\;n$, $S k \le n = k < n$,
$S (j + k) = (S\;j) + k = k + (S \;j)$; they are
the {\em logical glue} permitting to compose together the relevant
applicative steps (\verb+le_plus_to_le+ and \verb+monotonic_pred+,
forming the {\em trace} of the proof).
\subsubsection{Some statistics}
We are currently porting the old matita library (containing almost
no automation) to the new Matita system.
The following table compares the two
libraries on a fragment of about one hundred elementary arithmetical theorems.
The ``size'' is the dimension of the gzipped file in bytes.
\[
\begin{array}{c|c|c|c|c|}
& \mbox{lines} & \mbox{size} & \mbox{size} & \mbox{compilation} \\
& & \mbox{(whole)}& \mbox{(proofs)} & \mbox{time}\\\hline
\mbox{no auto} & 1139 & 5753 & 3433 (60\%) & 4.6s \\\hline
\mbox{with auto} & 627 & 3788 & 2027 (53\%) & 50.4s \\\hline
\mbox{traces} & 627 & 3982 & 2163 (54\%) & 5.3s \\\hline
\end{array}
\]
The sensible increase of compilation time in presence of automation
was somehow expected: the leading idea of the paper is that we are
ready to pay some extra execution time if this can reduce the
encoding effort on the user side (provided it does not sensibly
slow-down the system reactiveness to user commands in interactive
sessions: note that the average execution time per theorem is about
0.35 seconds). Traces seem to provide a natural balance between performance
and verbosity.
\section{Conclusions}
\label{sec:conclusions}
In this paper we introduced a general methodology to address
the complex problem of automation in interactive provers. The main
principles underlying our approach are the following:
\begin{enumerate}
\item there is an important distinction to be made between {\em small scale
automation}, mostly meant to reduce the verbosity of the proof script
(resolution of trivial steps, verification of side conditions,
automatic inference of missing information, etc.),
and {\em large scale automation} (problem solving): the problems and
requirements in the two cases are different, eventually
deserving different approaches and solutions;
\item a major component of small scale automation is the capability
to ``reason'' (apply logical rules and theorems) up to equalities,
covering most part of the background knowledge tacitly used in
the typical mathematical reasoning as an underlying connective
glue between logical steps (see Figure\ref{example_glue});
\item large scale automation must return a human readable and
system executable proof trace; the trace must be simple, hence
its execution will eventually require small scale automation
capabilites (independently of the choice of implementing or not
large scale automation {\em on top} of small scale automation).
\end{enumerate}
The paper also describes the current state of the implementation of
this program inside the Matita interactive theorem prover.
In particular we presented the architectural design of the
superposition tool supporting equational reasoning, and its
{\em not so trivial} integration inside the Matita Interactive Theorem
Prover.
The tool is already highly performant (we scored in fourth position
in the unit equality division at the 22nd CADE ATP System Competition),
but many improvements can still be done for efficiency.
In particular, more specialised data structures for indexes would
hopefully give us a chance to scale up with the current best ATPs.
Another interesting research direction is to extend the management
of equality to setoid rewriting \cite{setoidrewriting}.
Indeed, the current version of the superposition tool merely works
with an intensional equality, and it would be
interesting to try to figure out how to handle more general
binary relations. The main problem is proof reconstruction, but
again it looks possible to exploit the sophisticated capabilities
of the Matita refiner \cite{hints} to automatically check the legality of
the rewriting operation (i.e. the monotonicity of the context
inside which rewriting has to be performed).
While we are, at present, reasonably happy of the small scale
automation capabilities of Matita, much work is left about
large scale automation. Our current approach tries to build
large scale automation {\em on top} of small scale automation
(e.g. substituting application by its smart version); this
approach is natural, especially in view of the generation of
proof traces but, as we already observed, it is not the only
possibility compatible with our methodology and alternative
solutions (requiring a tighter integration of small scale
automation {\em techniques} inside large scale functionalities)
are worth to be explored.
Finally, let us remark that proof traces are per se an interesting
object worthty of further investigation (and, possibly, standardization),
in order to optimize the trade-off between efficiency and verbosity,
or to improve interoperability bewteen different systems.
\bibliographystyle{plain}
|
\section{Introduction and overview}\label{notation}
\subsection{Main result}
Let $K$ be a field that is finitely generated over a finite field $\kappa$ of characteristic~$p$. Let $K^{\text{sep}}$ be a fixed separable closure of $K$,
and let $\overline{\kappa}$ be the algebraic closure of $\kappa$ in $K^\text{sep}$. Let $G_K:=\Gal(K^{\text{sep}}/K)$ denote the absolute Galois group and $G_K^{\text{geom}}:=\Gal(K^{\text{sep}}/K \overline{\kappa})$ the geometric Galois group of~$K$.
Let $F$ be a finitely generated field of transcendence degree $1$ over~$\mathbb{F}_p$. Let $A$ be the ring of elements of $F$ which are regular outside a fixed place $\infty$ of~$F$. Let $\phi: A \rightarrow K\{\tau\}$ be a Drinfeld $A$-module of rank $r$ over $K$ of special characteristic~$\pp_0$. For any prime $\pp\not=\pp_0$ of $A$ let $\rho_\pp: G_K \to \GL_r(A_\pp)$ denote the homomorphism describing the Galois action on the Tate module $T_\pp(\phi)$. We are interested in the image of the associated adelic Galois representation
$$\rho_{\text{ad}} := (\rho_\pp)_\pp:\
G_K \longrightarrow \prod_{\pp\not=\pp_0}\GL_r(A_\pp).$$
By Anderson \cite{Anderson_t_Motives}, \S4.2, it is known that the composite of $\rho_{\text{ad}}$ with the determinant map is the adelic Galois representation associated to some Drinfeld module of rank $1$ of the same characteristic~$\pp_0$.
Thus the image of $\rho_{\text{ad}}(G_K^{\text{geom}})$ under the determinant is finite: see Proposition \ref{finite} below. Consequently, the image of $\rho_{\text{ad}}(G_K)$ under the determinant is an extension of a finite group and a pro-cyclic group and therefore far from open. Also, the main problem in determining $\rho_{\text{ad}}(G_K)$ lies in determining $\rho_{\text{ad}}(G_K^{\text{geom}}) \cap \prod_{\pp\not=\pp_0}\SL_r(A_\pp)$.
Recall that two subgroups of a group are called commensurable if their intersection has finite index in both. We will show that $\rho_{\text{ad}}(G_K^{\text{geom}})$ is commensurable to an explicit subgroup of $\prod_{\pp\not=\pp_0}\SL_r(A_\pp)$ whose definition depends only on information on certain endomorphism rings associated to~$\phi$. We will also determine $\rho_{\text{ad}}(G_K)$ up to commensurability.
First, since the Galois representation commutes with the endomorphisms of $\phi$ over~$K$, the image of $\rho_{\text{ad}}$ must be contained in the centralizer of $\End_K(\phi)$ in $\prod_{\pp\not=\pp_0}\GL_r(A_\pp)$. Second, enlarging $K$ does not change the image of Galois up to commensurability, but may increase the endomorphism ring. Since all endomorphisms of $\phi$ over any extension of $K$ are defined over a finite separable extension, the relevant endomorphism ring is therefore $\End_{K^{\text{sep}}}(\phi)$.
For a Drinfeld module in generic characteristic it turns out that the image of $\rho_{\text{ad}}$ up to commensurability, which was determined in \cite{PR2}, indeed depends only on $\End_{K^{\text{sep}}}(\phi)$. But in special characteristic this cannot be so, due to a phenomenon described in \cite{PinII}. The problem is that the endomorphism ring of a Drinfeld module in special characteristic can be non-commutative. As a consequence, it is possible that for some integrally closed infinite subring $B\subset A$, the endomorphism ring of the Drinfeld $B$-module $\phi|B$ is larger than that of~$\phi$. The Galois representation associated to $\phi$ must then commute with the additional operators coming from endomorphisms of $\phi|B$, forcing the image of $\rho_{\text{ad}}$ to be smaller. But using the results of \cite{PinII} one can reduce the problem to the case where this phenomenon of growing endomorphism rings does not occur.
For the following results let $a_0$ be any element of $A$ that generates a positive power of~$\pp_0$. View $a_0$ as a scalar element of $\prod_{\pp\not=\pp_0}\GL_r(A_\pp)$ via the diagonal embedding $A\hookrightarrow \prod_{\pp\not=\pp_0} A_\pp$, and let $\smash{\overline{\langle a_0\rangle}}$ denote the pro-cyclic subgroup that is topologically generated by it.
In the simplest case, where the endomorphism ring of $\phi$ over $K^{\text{sep}}$ is $A$ and does not grow under restriction, our main result is the following:
\begin{thm} \label{main_theorem_1}
Let $\phi$ be a Drinfeld $A$-module of rank $r$ over a finitely generated field $K$ of special characteristic $\pp_0$. Assume that for every integrally closed infinite subring $B \subset A$ we have $\End_{K^{\text{sep}}}(\phi|B)=A$.
Then
\begin{enumerate}
\item[(a)] $\rho_{\text{ad}}( G_K^{\geom})$ is commensurable to $\prod_{\pp\not=\pp_0}\SL_r(A_\pp)$, and
\item[(b)] $\rho_{\text{ad}}( G_K)$ is commensurable to $\overline{\langle a_0\rangle} \cdot \prod_{\pp\not=\pp_0}\SL_r(A_\pp)$.
\end{enumerate}
\end{thm}
More generally, set $R:= \End_{K^{\text{sep}}}(\phi)$ and $F := \mathop{\rm Quot}(A)$. Assume for the moment that the center of $R$ is~$A$. Then $R \otimes_A F$ is a central division algebra over~$F$ of dimension $d^2$ for some $d$ dividing~$r$. For any prime $\pp\not=\pp_0$ of~$A$, the Tate module $T_\pp(\phi)$ is a module over $R_\pp := R\otimes_AA_\pp$, which is an order in a semisimple algebra over~$F_\pp$. Let $D_\pp$ denote the commutant of $R_\pp$ in $\End_{A_\pp}(T_\pp(\phi))$, which is an order in another semisimple algebra over $F_\pp$. Let $D^1_\pp$ denote the multiplicative group of elements of $D_\pp$ of reduced norm~$1$. This is isomorphic to $\SL_{r/d}(A_\pp)$ for almost all $\pp$ by Proposition \ref{labacu}, and equal to $\SL_r(A_\pp)$ for all $\pp$ if $R=A$.
In this situation a version of our main result is the following:
\begin{thm} \label{main_theorem}
Let $\phi$ be a Drinfeld $A$-module over a finitely generated field $K$ of special characteristic $\pp_0$. Assume that $R:=\End_{K^{\text{sep}}}(\phi)$ has center $A$ and that for every integrally closed infinite subring $B \subset A$ we have $\End_{K^{\text{sep}}}(\phi|B)=R$. Let $D^1_\pp$ and $\smash{\overline{\langle a_0\rangle}}$ denote the subgroups defined above. Then
\begin{enumerate}
\item[(a)] $\rho_{\text{ad}}( G_K^{\geom})$ is commensurable to $\prod_{\pp\not=\pp_0} D^1_\pp$, and
\item[(b)] $\rho_{\text{ad}}( G_K)$ is commensurable to $\overline{\langle a_0\rangle} \cdot \prod_{\pp\not=\pp_0} D^1_\pp$.
\end{enumerate}
\end{thm}
Theorem \ref{main_theorem} is the central result of this article; its special case $R=A$ is just Theorem \ref{main_theorem_1}. Sections \ref{LAG} to \ref{sect_surjective} are dedicated to proving Theorem \ref{main_theorem}. In Section \ref{generalcase} we deduce corresponding results without any assumptions on $\End_{K^{\text{sep}}}(\phi)$ that are somewhat more complicated to state.
\subsection{Outline of the proof}
In this outline we explain the key steps in the proof of Theorem \ref{main_theorem} in the case $R=A$; the general case follows the same principles. So we assume that for every integrally closed infinite subring $B \subset A$ we have $\End_{K^{\text{sep}}}(\phi|B)=\nobreak A$.
After replacing $K$ by a finite extension, we may assume that $\rho_{\text{ad}}( G_K^{\geom}) \subset \prod_{\pp\not=\pp_0}\SL_r(A_\pp)$. Let $\Gamma_\pp^{\geom}$ denote its image in $\SL_r(A_\pp)$ for any single prime $\pp\not=\pp_0$ of $A$, and let $\Delta_\pp^{\geom}$ denote its image in $\SL_r(k_\pp)$ over the residue field $k_\pp := A/\pp$.
A large part of the effort goes into proving that $\Delta_\pp^{\geom} = \SL_r(k_\pp)$ for almost all $\pp$. The key arithmetic ingredients for this are the absolute irreducibility of the residual representation from \cite{PinIII}, the Zariski density of $\Gamma^{\geom}_\pp$ in $\SL_{r,F_\pp}$ from \cite{PinI}, and the characterization of $k_\pp$ by the traces of Frobenius elements in the adjoint representation from \cite{PinII}.
In fact, the absolute irreducibility combined with a strong form of Jordan's theorem on finite subgroups of $\GL_r$ from \cite{L-P} shows that $\Delta_\pp^{\geom}$ is essentially a finite group of Lie type in characteristic $p:=\car(F)$. Let $H_\pp$ denote the ambient connected semisimple linear algebraic group over an algebraic closure $\bar k_\pp$ of $k_\pp$. If $H_\pp$ is a proper subgroup of $\SL_{r,\bar k_\pp}$, the eigenvalues of any element of $H_\pp$ must satisfy one of finitely many explicit multiplicative relations that depend only on~$r$. In this case we show that the eigenvalues of any Frobenius element in the residual representation satisfy a similar relation. If this happens for infinitely many $\pp$, the fact that the adelic Galois representation is a compatible system implies that the eigenvalues of Frobenius over any single $F_\pp$ satisfy the same kind of relation. But that is impossible, because $\Gamma^{\geom}_\pp$ is Zariski dense in $\SL_{r,F_\pp}$. Therefore $H_\pp$ is equal to $\SL_{r,\bar k_\pp}$ for almost all $\pp$.
This means that $\Delta_\pp^{\geom}$ is essentially the group of $k'_\pp$-rational points of a model of $\SL_{r,\bar k_\pp}$ over a subfield $k'_\pp \subset\bar k_\pp$. To identify this subfield we observe that the trace in the adjoint representation for any automorphism of the model is an element of $k'_\pp$. We show that this holds in particular for the images of Frobenius elements. But by \cite{PinII} the images of the traces of all Frobenius elements in the adjoint representation of $\SL_r$ generate $k_\pp$ for almost all $\pp$. It follows that $k_\pp\subset k'_\pp$ for almost all $\pp$, and then the inclusion $\Delta_\pp^{\geom} \subset \SL_r(k_\pp)$ must be an equality for cardinality reasons.
We also need to prove that the homomorphism $G_K^{\geom} \to \SL_r(k_{\pp_1}) \times \SL_r(k_{\pp_2})$ is surjective for any distinct $\pp_1,\pp_2$ outside some finite set of primes. This again relies on traces of Frobenius elements. Indeed, if the homomorphism is not surjective, the surjectivity to each factor and Goursat's lemma imply that its image is essentially the graph of an isomorphism $\SL_r(k_{\pp_1}) \stackrel{\sim}{\to} \SL_r(k_{\pp_2})$. This isomorphism must come from an isomorphism of algebraic groups over an isomorphism of the residue fields $\sigma: k_{\pp_1} \stackrel{\sim}{\to} k_{\pp_2}$. Using this we show that the traces of Frobenius in the adjoint representation of $\SL_r$ map to the subring $\mathop{\rm graph}(\sigma) \subset k_{\pp_1}\times k_{\pp_2}$. But that again contradicts the result from \cite{PinII} unless $\pp_1$ or $\pp_2$ is one of finitely many exceptional primes.
Next we prove that $\Gamma_\pp^{\geom} = \SL_r(A_\pp)$ for almost all $\pp$. For this we may already assume that $\Delta_\pp^{\geom} = \SL_r(k_\pp)$. That alone does not imply much, because $A_\pp$ is a local ring of equal characteristic, and so the Teichm\"uller lift of the residue field $k_\pp\hookrightarrow A_\pp$ induces a lift $\SL_r(k_\pp) \hookrightarrow \SL_r(A_\pp)$. But using successive approximation in $\SL_r(A_\pp)$ we reduce the problem to showing that $\Gamma_\pp^{\geom}$ surjects to $\SL_r(A/\pp^2)$. This in turn we can guarantee for almost all $\pp$ using traces of Frobenius elements again.
Indeed, suppose first that $(p,r)\not=(2,2)$. Then the result from \cite{PinII} implies that the images of the traces of all Frobenius elements in the adjoint representation of $\SL_r$ generate $A/\pp^2$ for almost all $\pp$. In particular these traces do not all lie in the Teichm\"uller lift $k_\pp\subset A/\pp^2$, and so the images of Frobenius elements in $\GL_r(A_\pp)$ cannot all lie in the lift of $\GL_r(k_\pp)$. The desired surjectivity $\Gamma_\pp^{\geom} \twoheadrightarrow \SL_r(A/\pp^2)$ follows from this using some group theory.
In the remaining case $p=r=2$ it may happen that the traces of Frobenius in the adjoint representation do not generate the field $F$, but the subfield of squares $F^2 := \{x^2\mid x\in F\}$, of which $F$ is an inseparable extension of degree $2$. This phenomenon stems from the fact that the adjoint representation of $\SL_2$ on $\psll_2$ in characteristic $2$ factors through the Frobenius ${\rm Frob}_2: x\mapsto x^2$. In that case, the result from \cite{PinII} implies that the images of the traces of all Frobenius elements in the adjoint representation of $\SL_r$ generate the subring $k_\pp\oplus \pp^2/\pp^3$ of $A/\pp^3$ for almost all~$\pp$, where $k_\pp$ denotes the canonical Teichm\"uller lift of the residue field $k_\pp$ of~$\pp$. By digging deeper into the structure of $\SL_2(A/\pp^3)$, and replacing $K$ by a finite extension at a crucial step in the argument, we can again show that $\Gamma_\pp^{\geom}$ surjects to $\SL_r(A/\pp^2)$.
Finally, using group theory alone the above results about $\SL_r(k_{\pp_1}) \times \SL_r(k_{\pp_2})$ and $\SL_r(A_\pp)$ imply that the homomorphism $G_K^{\geom} \to \prod_{\pp\not\in P_3}\SL_r(A_\pp)$ is surjective for some finite set of primes $P_3$. On the other hand, the homomorphism $G_K^{\geom} \to \prod_{\pp\in P_3}\SL_r(A_\pp)$ has open image by the main result of \cite{PinII}. While this does not directly imply that the image of the product homomorphism $G_K^{\geom} \to \prod_{\pp\not=\pp_0}\SL_r(A_\pp)$ is open, because the image of a product map may be smaller than the product of the images, some variant of the argument can be made to work, thereby finishing the proof of Theorem \ref{main_theorem}~(a).
Theorem \ref{main_theorem}~(b) is deduced from this as follows. Since $\rho_{\text{ad}}(G_K^{\geom})$ is already open in $\prod_{\pp\not=\pp_0}\SL_r(A_\pp)$, it suffices to show that $\det\rho_{\text{ad}}(G_K)$ is commensurable to $\smash{\overline{\langle a_0\rangle}}$ within $\prod_{\pp\not=\pp_0} A_\pp^\times$. As the determinant of $\rho_{\text{ad}}$ is the adelic Galois representation associated to some Drinfeld module of rank $1$ of the same characteristic~$\pp_0$, this reduces the problem to the case that $r=1$ and that $\phi$ is defined over a finite field, say over $\kappa$ itself. Then $\Frob_\kappa$ acts through multiplication by an element $a\in A$ which is a unit at all primes $\pp\not=\pp_0$ but not at~$\pp_0$.
It follows that $(a)=\pp_0^i$ for some positive integer~$i$. The same properties of $a_0$ show that $(a_0)=\pp_0^j$ for some positive integer~$j$. Together it follows that $(a^j) = \pp_0^{ij} = (a_0^i)$, and so $a^j/a_0^i$ is a unit in~$A^\times$. As the group of units is finite, we deduce that $a^{j\ell} = a_0^{i\ell}$ for some positive integer~$\ell$. In particular $\rho_{\text{ad}}(G_K) = \smash{\overline{\langle a\rangle}}$ is commensurable to $\smash{\overline{\langle a_0\rangle}}$, as desired. This finishes the proof of Theorem \ref{main_theorem}~(b).
\subsection{Structure of the article}
Section \ref{notation} is the present introduction and overview. Sections \ref{LAG} and \ref{approx} deal with subgroups of $\SL_n$ and $\GL_n$. They are independent of Drinfeld modules, of the rest of the article, and of each other.
Section \ref{LAG} deals with subgroups of $\SL_n$ and $\GL_n$ over a field and establishes suitable conditions for such a subgroup to be equal to $\SL_n$. It is based on some calculations in root systems, on known results on finite groups of Lie type, and on a strong form of Jordan's theorem from \cite{L-P}.
Section \ref{approx} deals with closed subgroups of $\SL_n$ and $\GL_n$ over a complete discrete valuation ring $R$ of equal characteristic $p$ with finite residue field, and establishes suitable conditions for such a subgroup to be equal to $\SL_n(R)$. The method uses successive approximation over the congruence filtration of $\SL_n(R)$, respectively of $\GL_n(R)$, whose subquotients are related to the adjoint representation. Curiously, the case $p=n=2$ presents special subtleties here, too, because the Lie bracket on $\sll_2$ in characteristic $2$ is not surjective.
In Section \ref{known_results} we list known results about Drinfeld modules in special characteristic or adapt them slightly to the situation at hand. This includes properties of endomorphism rings, Galois representations on Tate modules, characteristic polynomials of Frobenius, and bad reduction. We also create the setup in which the proof of Theorem \ref{main_theorem} takes place, and list the main arithmetic ingredients from \cite{PinI}, \cite{PinII}, and \cite{PinIII} with their immediate consequences.
Section \ref{sect_surjective} then contains (what remains of) the proof of Theorem \ref{main_theorem}, following the outline expained above.
In Section \ref{generalcase} we determine $\rho_{\text{ad}}(G_K^{\geom})$ and $\rho_{\text{ad}}(G_K)$ up to commensurability for arbitrary Drinfeld modules in special characteristic. The main ingredients for this are the special case of Theorem \ref{main_theorem} and some reduction steps from \cite{PinII}.
\medskip
This article is based on the doctoral thesis of the first author \cite{DevicThesis}; its results are roughly the same as the results there. We are grateful to Florian Pop for pointing out Theorem~\ref{Pop}.
\section{Subgroups of linear algebraic groups}\label{LAG}
\section{Subgroups of $\SL_n$ over a field}\label{LAG}
In a nutshell, the main goal of this section is to establish suitable conditions for subgroups of $\SL_n$ over a field to be equal to $\SL_n$. We first give conditions for root systems to be simple of type $A_\ell$, and then deal with the case of connected semisimple linear algebraic groups over a field. Based on this we treat the case of finite groups of Lie type, which must also take inner forms of $\SL_n$ into account. The main results are Theorems \ref{LAG_main}, \ref{finite_main}, and \ref{finite_main1}. We also recall a strong form of Jordan's theorem from \cite{L-P}.
\subsection{Root systems}\label{21root}
Let $\Phi$ be a non-trivial root system generating a euclidean vector space $E$. Let $W$ be the associated Weyl group, and let $S$ be a $W$-orbit in $E$. We are interested in the conditions:
\begin{enumerate}
\item[(a)] $S$ generates $E$ as a vector space.
\item[(b)] There are no distinct elements $\lambda_1,\ldots,\lambda_4 \in S$ such that $\lambda_1+ \lambda_2=\lambda_3+\lambda_4$.
\item[(c)] There are no distinct elements $\lambda_1,\ldots,\lambda_6 \in S$ such that $\lambda_1+ \lambda_2+\lambda_3=\lambda_4+\lambda_5+\lambda_6$.
\end{enumerate}
\begin{thm}\label{mainroot}
Assume (a) and (b). Then $\Phi$ is simple of type $A_\ell$ for some $\ell\geq 1$. Moreover, if
$$\Phi =\{\pm(e_i-e_j) \mid 0\leq i <j \leq \ell \} \subset E= \mathbb{R}^{\ell+1}/\diag(\mathbb{R})$$
in standard notation, and if $\ell\not=2$ or in addition (c) is satisfied, then
$$S=\{ c e_i \mid 0 \leq i \leq \ell\}$$
for some constant $c \neq 0$.
\end{thm}
The proof of this result extends over the rest of this subsection. Throughout we assume conditions (a) and (b). Note that (a) implies that $0\not\in S$.
\begin{lem}\label{A1A1}
Let $\lambda \in S$ and $\alpha_1$, $\alpha_2$ be two orthogonal roots in $\Phi.$ Then $\lambda\perp \alpha_1$ or $\lambda\perp \alpha_2.$
\end{lem}
\begin{proof}
Let $s_{\alpha_i}\in W$ denote the simple reflection associated to $\alpha_i$. The fact that $\alpha_1\perp\alpha_2$ implies that
\begin{eqnarray*}
s_{\alpha_i}(\lambda)
&=& \lambda - \frac{2(\lambda,\alpha_i)}{(\alpha_i,\alpha_i)}\cdot\alpha_i,
\qquad\hbox{and}\\
s_{\alpha_1}s_{\alpha_2}(\lambda)
&=& \lambda - \frac{2(\lambda,\alpha_1)}{(\alpha_1,\alpha_1)}\cdot\alpha_1
- \frac{2(\lambda,\alpha_2)}{(\alpha_2,\alpha_2)}\cdot\alpha_2,
\end{eqnarray*}
and hence
$$\lambda + s_{\alpha_1}s_{\alpha_2}(\lambda) =
s_{\alpha_1}(\lambda)+ s_{\alpha_2}(\lambda).$$
But if $\lambda$ is not orthogonal to $\alpha_1$ or $\alpha_2$, these are four distinct elements of $S$, contradicting condition (b).
\end{proof}
\begin{lem}\label{simple}
The root system $\Phi$ is simple.
\end{lem}
\begin{proof}
Assume that $\Phi=\Psi_1 +\Psi_2$ is decomposable and let $\lambda \in S$. Since $\Phi$ generates $E$, there exists an $\alpha \in \Phi$ which is not orthogonal to $\lambda$. Suppose without loss of generality that $\alpha \in \Psi_2$. Then, by Lemma \ref{A1A1}, the vector $\lambda$ is orthogonal to all roots that are orthogonal to $\alpha$; in particular $\lambda \perp \Psi_1$ . Then $w(\lambda) \perp \Psi_1$ for all $w \in W$ and therefore $S \perp \Psi_1.$ However, this contradicts condition (a).
\end{proof}
\begin{lem}\label{noB2}
The root system $\Phi$ does not contain a root subsystem of type $B_2$.
\end{lem}
\begin{proof}
Assume that $\Psi\subset\Phi$ is a root subsystem of type $B_2$. Then the subspace ${\mathbb R}\Psi$ possesses a basis $\{e_1,e_2\}$ such that $\Psi$ consists of eight roots $\pm e_1$, $\pm e_2$, and $\pm e_1\pm e_2$, and where $e_1\perp e_2$ and $e_1+e_2\perp e_1-e_2$. Thus for any $\lambda\in S$, Lemma \ref{A1A1} implies that $\lambda\perp e_i$ for some $i=1$, $2$, and that $\lambda\perp e_1\pm e_2$ for some choice of sign. Together this gives four cases, in each of which we deduce that $\lambda\perp e_1$. As $\lambda$ was arbitrary, this shows that $S\perp e_1$, contradicting condition (a).
\end{proof}
\begin{lem}\label{noG2}
The root system $\Phi$ is not of type $G_2$.
\end{lem}
\begin{proof}
Choose simple roots $\alpha_1$, $\alpha_2$ of $\Phi$ such that $\alpha_1$ is the shorter one. Then $\Phi$ contains the root $2\alpha_1+\alpha_2$ which is orthogonal to $\alpha_2$, and the root $3\alpha_1+2\alpha_2$ which is orthogonal to $\alpha_1$. Thus for any $\lambda\in S$, Lemma \ref{A1A1} implies that $\lambda\perp\alpha_2$ or $\lambda\perp 2\alpha_1+\alpha_2$, and that $\lambda\perp\alpha_1$ or $\lambda\perp3\alpha_1+2\alpha_2$. By a simple calculation, each of these four cases implies that $\lambda=0$, contradicting condition (a).
\end{proof}
\begin{lem}\label{noD4}
The root system $\Phi$ does not contain a root subsystem of type $D_4$.
\end{lem}
\begin{proof}
Assume that $\Psi\subset\Phi$ is a root subsystem of type $D_4$. Then, up to scaling the inner product on $E$, the subspace ${\mathbb R}\Psi$ possesses an orthonormal basis $\{e_1,e_2,e_3,e_4\}$ such that $\Psi$ consists of the roots $\pm e_i\pm e_j$ for all $1\leq i<j\leq4$ and all choices of signs. In particular, the roots $e_1+e_i$ and $e_1-e_i$ are orthogonal for every $2\leq i\leq4$. Thus for any $\lambda\in S$, Lemma \ref{A1A1} implies that $\lambda\perp e_1+\epsilon_ie_i$ for some $\epsilon_i=\pm1$. Since the roots $e_1-\epsilon_2e_2$ and $\epsilon_3e_3+\epsilon_4e_4$ are also orthogonal, Lemma \ref{A1A1} implies that $\lambda\perp e_1-\epsilon_2e_2$ or $\lambda\perp \epsilon_3e_3+\epsilon_4e_4$. Since
$$(e_1+\epsilon_2e_2)+(e_1-\epsilon_2e_2) \ =\
(e_1+\epsilon_3e_3)+(e_1+\epsilon_4e_4)-(\epsilon_3e_3+\epsilon_4e_4) \ =\ 2e_1,$$
in both cases we deduce that $\lambda\perp 2e_1$. As $\lambda$ was arbitrary, this shows that $S\perp 2e_1$, contradicting condition (a).
\end{proof}
Combining Lemmas \ref{simple} through \ref{noD4}, it follows that $\Phi$ is a simple root system of type $A_\ell$ for some $\ell\ge1$. Using standard notation we may identify $E$ with the vector space $\mathbb{R}^{\ell+1}/\diag(\mathbb{R})$, let $e_0,\ldots,e_\ell\in E$ denote the images of the standard basis vectors of $\mathbb{R}^{\ell+1}$, and assume that $\Phi$ consists of the roots $e_i-e_j$ for all distinct $0\leq i,j \leq\ell$. Then its Weyl group is the symmetric group $S_{\ell+1}$ on $l+1$ letters, acting on $E$ by permuting the coefficients.
Consider any $\lambda\in S$ and write $\lambda = (a_0,\ldots,a_\ell)$ modulo $\diag(\mathbb{R})$. Since $\lambda\not=0$ in~$E$, the coefficients $a_i$ are not all equal.
\begin{lem}\label{no4distinct}
Suppose that $\ell\ge3$, and consider indices $i$ and $j$ satisfying $a_i\not=a_j$. Then for all indices $i'$ and $j'$ that are distinct from $i$ and $j$ we have $a_{i'}=a_{j'}$.
\end{lem}
\begin{proof}
The assumption implies that $i\not=j$, and the assertion is trivial unless also $i'\not=j'$. Then $e_i-e_j$ and $e_{i'}-e_{j'}$ are orthogonal roots, and so Lemma \ref{A1A1} implies that $\lambda\perp e_i-e_j$ or $\lambda\perp e_{i'}-e_{j'}$. This means that $a_i=a_j$ or $a_{i'}=a_{j'}$; but by assumption only the second case is possible.
\end{proof}
\begin{lem}\label{no3distinct}
If $\ell\ge3$, there exists an index $i$ such that the $a_j$ for all $j\not=i$ are equal.
\end{lem}
\begin{proof}
Since $\ell\geq3$ and the $a_i$ are not all equal, Lemma \ref{no4distinct} implies that the $a_i$ are also not all distinct. Therefore there exist distinct indices $i,j,j'$ satisfying $a_i\not=a_j=a_{j'}$. Then for any $i'\not=i,j$, Lemma \ref{no4distinct} shows that $a_{i'}=a_{j'}$. Thus $i$ has the desired property.
\end{proof}
\begin{lem}\label{no3distinct2}
If $\ell=2$ and in addition condition (c) is satisfied, then the $a_i$ are not all distinct.
\end{lem}
\begin{proof}
Being an orbit under the Weyl group, the set $S$ consists of the vectors
\begin{eqnarray*}
(a_0,a_1,a_2), & (a_1,a_2,a_0), & (a_2,a_0,a_1), \\
(a_0,a_2,a_1), & (a_1,a_0,a_2), & (a_2,a_1,a_0)
\end{eqnarray*}
modulo $\diag(\mathbb{R})$.
If the $a_i$ are all distinct, these six vectors are all distinct in $E$, for instance because the positions of the greatest and the smallest coefficient of a vector in $\mathbb{R}^3$ depend only on its residue class modulo $\diag(\mathbb{R})$. As the sum of the three vectors in the first row is equal to the sum of those in the second row, that contradicts condition (c).
\end{proof}
We can now prove Theorem \ref{mainroot}. The statement about $\Phi$ has already been established. To show the statement about $S$, we may assume condition (c) if $\ell=2$. If $\ell\ge2$, using the action of the Weyl group $S_{\ell+1}$, Lemma \ref{no3distinct} or \ref{no3distinct2} implies that $S$ contains an element of the form $(a,b,\ldots,b) \mod \diag(\mathbb{R})$ with $a\not=b$. The same is trivially true if $\ell=1$, because then any non-zero element of $E$ has this form. But for any $\ell\geq1$, the indicated element of $E$ is equal to $ce_1$ with $c=a-b\not=0$. Since $S$ is an orbit under $S_{\ell+1}$, it follows that $S=\{ c e_i \mid 0 \leq i\leq\ell\}$, as desired. This finishes the proof of Theorem \ref{mainroot}.
\subsection{Some algebraic relations}\label{alg_relations}
From here until the end of this section we fix an integer $n\geq2$. Consider the expression
\begin{myequation}\label{Rprod}
\vcenter{\hbox{$\quad
\begin{array}{l}\displaystyle
\!\!\prod_{\substack{i_1,i_2 \\ \text{distinct}}} \!
(\alpha_{i_1}-\alpha_{i_2})
\;\cdot\!\!\prod_{\substack{i_1,i_2,i_3 \\ \text{distinct}}} \!
(\alpha_{i_1}\alpha_{i_2}-\alpha_{i_3}^2)
\;\cdot\!\! \\[25pt]
\displaystyle\hskip1.3cm
\;\cdot\!\!\prod_{\substack{i_1,\ldots, i_4 \\ \text{distinct}}} \!
(\alpha_{i_1}\alpha_{i_2}-\alpha_{i_3}\alpha_{i_4})
\;\cdot\!\!\prod_{\substack{i_1,\ldots, i_6 \\ \text{distinct}}} \!
(\alpha_{i_1}\alpha_{i_2}\alpha_{i_3}-\alpha_{i_4}\alpha_{i_5}\alpha_{i_6}),
\end{array}
$}}
\end{myequation}%
where the products are extended over all tuples of distinct indices in $\{1,\ldots,n\}$.
(Note that some of these products are empty for small $n$, but this will not cause any problems.) Clearly this is a symmetric polynomial with integral coefficients in the variables $\alpha_1,\ldots,\alpha_n$. It can therefore be written uniquely as a polynomial with integral coefficients in $\beta_1,\ldots,\beta_n$, where
$$\prod_{i=1}^n \;(T-\alpha_i)\ =\ T^n+ \beta_1 T^{n-1}+\cdots+\beta_n.$$
In particular, we can apply it to the coefficients of the characteristic polynomial $\det(T\cdot{\rm Id}_n-\gamma)$ of a matrix $\gamma \in\GL_n$ over any field $L$ and obtain an algebraic morphism
\begin{myequation}\label{Rdef}
f: \GL_{n,L} \rightarrow \mathbb{A}_L^1.
\end{myequation}%
By construction, this morphism has the following property:
\begin{lem}\label{Rprop}
For any algebraically closed field $L$ and any matrix $\gamma\in\GL_n(L)$ with eigenvalues $\alpha_1,\ldots,\alpha_n \in L$, listed with their respective multiplicities, we have $f(\gamma)=0$ if and only if one of the following holds:
\begin{enumerate}
\item[(a)] There exist distinct indices $i_1,i_2$ such that $\alpha_{i_1}=\alpha_{i_2}$.
\item[(b)] There exist distinct indices $i_1,i_2,i_3$ such that $\alpha_{i_1}\alpha_{i_2}=\alpha_{i_3}^2$.
\item[(c)] There exist distinct indices $i_1,\ldots,i_4$ such that $\alpha_{i_1}\alpha_{i_2}=\alpha_{i_3}\alpha_{i_4}$.
\item[(d)] There exist distinct indices $i_1,\ldots,i_6$ such that $\alpha_{i_1}\alpha_{i_2}\alpha_{i_3}=\alpha_{i_4}\alpha_{i_5}\alpha_{i_6}$.
\end{enumerate}
\end{lem}
\begin{lem}\label{R_nontrivial}
For any field $L$ and any integer $N \geq 1$, the morphism
$$\GL_{n,L} \to \mathbb{A}_L^1,\ \gamma \mapsto f(\gamma^N)$$
is not identically zero.
\end{lem}
\begin{proof}
Let $T\subset\GL_{n,L}$ be the subgroup of diagonal matrices. Then none of the factors in (\ref{Rprod}) is identically zero on $T$; hence $f$ is not identically zero on $T$. Since the morphism $T\to T$, $\gamma\mapsto \gamma^{nN}$ is surjective, it follows that $\gamma\mapsto f(\gamma^{nN})$ is not identically zero on $T$, and hence not on $\GL_{n,L}$. This implies the desired conclusion.
\end{proof}
\subsection{Linear algebraic groups}
\begin{thm} \label{LAG_main}
Let $n\geq2$, let $L$ be an algebraically closed field, and let $G$ be a connected semisimple linear algebraic subgroup of $\SL_{n,L}$. Assume that the tautological representation of $G$ on $L^n$ is irreducible and the morphism $f$ from (\ref{Rdef}) does not vanish identically on $G$. Then $G = \SL_{n,L}$.
\end{thm}
\begin{proof}
Let $T$ be a maximal torus of $G$, let $E=X^*(T)\otimes{\mathbb R}$ be the associated character space, let $\Phi\subset E$ be the root system of $G$ with respect to $T$, and let $W$ denote the Weyl group of $\Phi$. The assumption $n\geq2$ and the irreducibility implies that $G$ and hence $\Phi$ is non-trivial.
Let $S\subset E$ be the set of weights of $T$ in the given representation on $L^n$. The fact that the representation is faithful implies that $S$ generates $E$. Let $\lambda\in S$ denote the highest weight of the representation, and let $W\lambda\subset S$ denote its orbit under $W$. Then $S$ is contained in the convex closure of $W\lambda$;
hence $W\lambda$ also generates $E$.
Next, since the conjugates of $T$ form a Zariski dense subset of $G$, and $f$ does not vanish identically on $G$, it follows that $f$ does not vanish identically on $T$. From this we conclude that
\begin{enumerate}
\item[(a)] none of the weights $\lambda\in S$ has multiplicity $>1$;
\item[(b)] there are no distinct elements $\lambda_1,\lambda_2,\lambda_3 \in S$ such that $\lambda_1+\nobreak\lambda_2\allowbreak=2\lambda_3$;
\item[(c)] there are no distinct elements $\lambda_1,\ldots,\lambda_4 \in S$ such that $\lambda_1+ \lambda_2=\lambda_3+\lambda_4$;
\item[(d)] there are no distinct elements $\lambda_1,\ldots,\lambda_6 \in S$ such that $\lambda_1+ \lambda_2+\lambda_3=\lambda_4+\lambda_5+\lambda_6$;
\end{enumerate}
because by Lemma \ref{Rprop} any one of these relations would imply that $f|T=0$.
In particular, the assumptions of Theorem \ref{mainroot} are satisfied for $\Phi$ and the orbit $W\lambda$. It follows that $\Phi$ is simple of type $A_\ell$ for some $\ell\geq 1$ and that $W\lambda = \{ c e_i \mid 0 \leq i \leq \ell\}$ in standard notation for some constant $c \neq 0$.
Since $W\lambda$ consists of weights, $c$ is an integer. Let us use the standard ordering of $A_\ell$, where the simple roots are $e_{i-1}-e_i$ for all $1\leq i\leq\ell$. The fact that $\lambda$ is dominant then implies that $\lambda=ce_0$ with $c >0$, or $\lambda=ce_\ell$ with $c<0$. These two cases correspond to dual representations which are interchanged by the outer automorphism of $A_\ell$; hence we may assume that $\lambda=ce_0$ and $c>0$.
\begin{lem} \label{premet}
Suppose that $L$ has characteristic $p>0$. Then $0 < c \leq p-1$.
\end{lem}
\begin{proof}
For any integer $d\ge0$ let $V_{d}$ denote the irreducible representation of $\SL_{\ell+1,L}$ with highest weight $de_0$. We know already that there exists a central isogeny $\SL_{\ell+1,L}\twoheadrightarrow G$, such that the pullback of the given representation on $L^n$ is isomorphic to $V_{c}$. Write $c=a+pb$ with integers $0\leq a\leq p-1$ and $b\geq0$. Then by Steinberg's Tensor Product Theorem (cf.\ \cite{HumphMod}, Theorem 2.7) we have $V_{c} \cong V_{a} \otimes V_{b}^{(p)}$, where $(\ )^{(p)}$ denotes the pullback under the absolute Frobenius morphism ${\rm Frob}_p$, which on coordinates is given by $x\mapsto x^p$.
If $a=0$, it follows that the homomorphism $\SL_{\ell+1,L} \twoheadrightarrow G \subset \SL_{n,L}$ factors through ${\rm Frob}_p$, which is not a central isogeny. We must therefore have $a>0$. Suppose that $b>0$. Then the $ae_i$ for $0\leq i\leq\ell$ are distinct weights of $V_{a}$, and the $be_j$ for $0\leq j\leq\ell$ are distinct weights of $V_{b}$; hence the $\lambda_{ij} := ae_i+pbe_j$ for $0\leq i,j\leq\ell$ are distinct weights of $V_{c}$. In other words, the $\lambda_{ij}$ for $0\leq i,j\leq\ell$ are distinct elements of $S$. Since $\lambda_{00}+\lambda_{11}=\lambda_{01}+\lambda_{10}$, this contradicts the property (c) above. We must therefore have $b=0$ and so $0<c\leq p-1$, as desired.
\end{proof}
\begin{lem} \label{postmet}
For all $0\leq i\leq c$ we have $(c-i)e_0+ie_1\in S$.
\end{lem}
\begin{proof}
Consider the simple root $\alpha:=e_0-e_1$, and let $U_{\pm\alpha}\subset G$ denote the two root subgroups, isomorphic to ${\mathbb G}_{a,L}$ and normalized by $T$, corresponding to $\pm\alpha$. Let $H_\alpha\subset G$ denote the subgroup generated by $T$ and $U_{\alpha}$ and $U_{-\alpha}$, whose semisimple part has root system $\{\pm\alpha\}$ of type $A_1$. For any weight $\mu$ let $V_\mu \subset L^n$ denote the associated weight space, and recall that the highest weight is $\lambda=ce_0$. Then the subspace $\bigoplus_{i \in \IZ} V_{ce_0-i\alpha}$ is $H_\alpha$-invariant and irreducible with highest weight $ce_0$ by \cite{Jan}, Part II, Proposition 2.11.
If $L$ has characteristic $0$, by classical results the representation of the Lie algebra of $H_\alpha$ on this subspace is irreducible with highest weight $ce_0$. If $L$ has characteristic $p>0$, we have $0< c \leq p-1$ by Lemma \ref{premet}, and so the same conclusion holds by \cite{Premet}, Theorem 1.
{}From \cite{HumLie}, Proposition 21.3, it follows that the set of weights of this representation is saturated; in other words these weights are $ce_0-i\alpha$ for all $0\leq i\leq 2(ce_0,\alpha)/(\alpha,\alpha) = c$. They therefore appear in $S$, as desired.
\end{proof}
In particular, Lemma \ref{postmet} implies that $S$ contains the elements
\begin{eqnarray*}
\lambda_1 := ce_0, && \lambda_3 := (c-1)e_0+e_1, \\
\lambda_2 := ce_1, && \lambda_4 := e_0+(c-1)e_1.
\end{eqnarray*}
If $c\ge3$, these elements are all distinct. If $c=2$, we have $\lambda_3=\lambda_4$, but $\lambda_1,\lambda_2,\lambda_3$ are all distinct. Since $\lambda_1+ \lambda_2=\lambda_3+\lambda_4$, we obtain a contradiction to the above property (c) if $c\geq3$, respectively to (b) if $c=2$. We must therefore have $c=1$.
But then $G\cong\SL_{\ell+1,L}$ and the given representation is isomorphic to the standard representation on $L^{\ell+1}$. Thus $\ell+1=n$ and so $G=\SL_{n,L}$, as desired. This finishes the proof of Theorem \ref{LAG_main}.
\end{proof}
\subsection{Finite groups of Lie type}
In this subsection $L$ denotes an algebraically closed field of characteristic $p > 0$.
Let $G$ be a simply connected simple semisimple linear algebraic group over $L$. A surjective endomorphism $F: G \rightarrow G$ whose group of fixed points $G^F$ is finite is called a \emph{Frobenius map} on $G$. For any such $F$, any non-abelian finite simple group isomorphic to a Jordan-H\"{o}lder constituent of $G^F$ is called a \emph{finite simple group of Lie type in characteristic $p$}.
A few small finite simple groups of Lie type have idiosyncrasies that we avoid with the following ad hoc definition. Denote the center of a group $H$ by $Z(H)$.
\begin{defn}\label{regular}
Let $\Gamma$ be a finite simple group of Lie type in characteristic $p$. We call $\Gamma$ \emph{regular} if there exist $G$ and $F$ as above such that
\begin{enumerate}
\item[(a)] $\Gamma \cong G^F/Z(G^F)$,
\item[(b)] $G^F$ is perfect, and
\item[(c)] $G^F$ is the universal central covering of $\Gamma$ as an abstract group.
\end{enumerate}
\end{defn}
\begin{prop}\label{Struct_fgolt}
Up to isomorphism, there are only finitely many finite simple groups of Lie type, in any characteristic, that are not regular.
\end{prop}
\begin{proof}
Suppose that $\Gamma$ is a non-abelian Jordan-H\"{o}lder constituent of $G^F$.
Since $G$ is simply connected, by \cite{GoLySo}, \nolinebreak Theorem 2.2.6 \nolinebreak (f) the group $G^F$ is generated by elements whose order is a power of $p$. We can therefore apply \cite{GoLySo}, Theorem 2.2.7, to $G^F$. The first part of this theorem says that, with finitely many exceptions up to isomorphism, $G^F/Z(G^F)$ is non-abelian simple. It is therefore isomorphic to $\Gamma$. The second part says that, with the same exceptions as in the first part, the group $G^F$ is perfect.
As $\Gamma$ is simple and hence perfect, by \cite{GoLySo}, Theorem 5.1.2, it possesses a universal central covering $\Gamma^c\twoheadrightarrow\Gamma$ which is unique up to isomorphism. Its kernel $M(\Gamma)$ is called the \emph{Schur multiplier} of $\Gamma$.
By \cite{GoLySo}, Theorem 6.1.4, after removing another finite number of exceptions up to isomorphism (these are listed in Table 6.1.3), the Schur multiplier $M(\Gamma)$ is isomorphic to $Z(G^F)$. Since $G^F$ is already perfect with $G^F/Z(G^F)\cong\Gamma$, this implies that $G^F$ is the universal central covering of $\Gamma$.
Then $\Gamma$ is regular, and the proposition follows.
\end{proof}
The next result is a direct consequence of the stronger statements of \cite{HumphMod}, Theorems 2.11 and 20.2.
\begin{prop}\label{Irred_fgolt}
Let $G$ and $F$ be as above, and let $\rho: G^F \rightarrow \SL_n(L)$ be an irreducible representation on the vector space $L^n$. Then $\rho$ is the restriction to $G^F$ of an irreducible algebraic representation $\rho_G : G \rightarrow \SL_{n,L}$.
\end{prop}
Now we can state our analogues of Theorem \ref{LAG_main}.
\begin{thm}\label{finite_main}
Let $n\geq2$, and let $\Gamma$ be a finite subgroup of $\SL_n(L)$ that acts irreducibly on $L^n$. Assume that $\Gamma$ is perfect and that $\Gamma/Z(\Gamma)$ is a direct product of finite simple groups of Lie type in characteristic $p$ that are regular in the sense of Definition \ref{regular}. Assume moreover that the map $\Gamma\to L$, $\gamma\mapsto f(\gamma)$ is not identically zero. Then there exist a finite subfield $k'$ of $L$ and a model $G'$ of $\SL_{n,L}$ over $k'$ such that $\Gamma=G'(k').$
\end{thm}
\begin{proof}
Let $\overline{\Gamma}_1, \ldots, \overline{\Gamma}_m$ denote the simple factors of $\Gamma/Z(\Gamma)$ and $\Gamma_1,\ldots,\Gamma_m$ their inverse images in~$\Gamma$.
Then the natural homomorphism $\Gamma_1\times\ldots\times\Gamma_m \twoheadrightarrow\Gamma$ is a central extension. By \cite{GoFinite}, Theorem 3.7.1, the pullback of the given irreducible representation on $L^n$ is the exterior tensor product of irreducible representations $\Gamma_i \to \GL_{n_i}(L)$ for certain ${n_i\geq1}$. In fact every $n_i\ge2$, because the corresponding projective representation of $\overline{\Gamma}_1\times\ldots\times\overline{\Gamma}_m$ is faithful.
For each $1 \leq i \leq m$ choose a simply connected simple semisimple linear algebraic group $G_i$ over $L$, a Frobenius map $F_i : G_i \rightarrow G_i$, and an isomorphism $G_i^{F_i}/Z(G_i^{F_i}) \cong \overline{\Gamma}_i$, such that $G_i^{F_i}$ is perfect and the universal central covering of~$\overline{\Gamma}_i$. By the last property the isomorphism lifts to a homomorphism $G_i^{F_i} \to\Gamma_i$. By Proposition \ref{Irred_fgolt} the composite homomorphism $G_i^{F_i} \to\Gamma_i \to \GL_{n_i}(L)$ is the restriction of some irreducible algebraic representation $\rho_i : G_i \rightarrow \GL_{n_i,L}$. Since $G_i$ is simple and $n_i\geq2$, the kernel of this homomorphism is finite.
Set $G := G_1\times\ldots\times G_m$. Then the exterior tensor product of the above $\rho_i$ defines an irreducible algebraic representation $\rho: G \to \GL_{n,L}$. By construction its kernel is finite; in other words it induces an isogeny $G\twoheadrightarrow\rho(G)$. Moreover, with the Frobenius map $F := F_1\times\ldots\times F_m$ on $G$ the homomorphism $\rho$ induces a homomorphism $G^F\to\Gamma$ lifting the given isomorphism
$$G^F/Z(G^F) \ =\ \prod_{i=1}^m G_i^{F_i}/Z(G_i^{F_i})
\ \cong\ \prod_{i=1}^m \overline{\Gamma}_i \ \cong\ \Gamma/Z(\Gamma).$$
As $\Gamma$ is perfect, it follows that $G^F\to\Gamma$ is surjective.
Since $G$ is a connected semisimple algebraic group, so is its image $\rho(G)$, which is therefore contained in $\SL_{n,L}$. Moreover, the tautological representation of $\rho(G)$ on $L^n$ is again irreducible. Furthermore, since by assumption the morphism $f$ does not vanish identically on the subgroup $\Gamma \subset \rho(G)$, it does not vanish identically on $\rho(G)$. By Theorem \ref{LAG_main} we therefore have $\rho(G) = \SL_{n,L}$.
In particular $\rho(G)$ is simple of type $A_{n-1}$, and so $G$ itself is simple of type~$A_{n-1}$.
As $G$ is simply connected, it is therefore isomorphic to $\SL_{n,L}$. Consider the resulting isogeny $\rho: \SL_{n,L}\cong G\twoheadrightarrow\rho(G) = \SL_{n,L}$.
Its scheme-theoretic kernel is contained in the scheme-theoretic kernel of $\rho\circ F$; hence
there exists an isogeny $F': \SL_{n,L} \to \SL_{n,L}$ satisfying $F'\circ\rho = \rho\circ F$. On the other hand $\rho$ is bijective; hence it induces an isomorphism from $G^F$ to $\SL_{n,L}^{F'}$. Together it follows that $\Gamma = \SL_{n,L}^{F'}$.
Finally, by known classification results such as \cite{CarterShort}, Proposition 4.5, the Frobenius map $F'$ is standard. This means that there is a finite subfield $k' \subset L$ and a model $G'$ of $\SL_{n,L}$ over $k'$ such that $\SL_{n,L}^{F'}=G'(k')$. Thus Theorem \ref{finite_main} is proved.
\end{proof}
For the next theorem let $c$ denote the least common multiple of the orders of all finite simple groups of Lie type that are not regular in the sense of Definition \ref{regular}, which is finite by Proposition \ref{Struct_fgolt}. Let $\Gamma^{\rm der}$ denote the derived group of $\Gamma$.
\begin{thm}\label{finite_main1}
Let $n\geq2$, and let $\Gamma$ be a finite subgroup of $\GL_n(L)$ that acts irreducibly on~$L^n$. Assume that $\Gamma/Z(\Gamma)$ is a direct product of finite simple groups of Lie type in characteristic $p$. Assume moreover that the map $\Gamma\to L$, $\gamma\mapsto f(\gamma^c)$ is not identically zero. Then there exist a finite subfield $k'$ of $L$ and a model $G'$ of $\SL_{n,L}$ over $k'$ such that $\Gamma^{\rm der}=G'(k').$
\end{thm}
\begin{proof}
Let $\overline{\Gamma}_i$ and $\Gamma_i \to \GL_{n_i}(L)$ be as in the proof of Theorem \ref{finite_main}. Suppose that some factor of $\Gamma/Z(\Gamma)$, say $\overline{\Gamma}_1$, is not regular. Then for every $\gamma\in\Gamma$, the definition of $c$ implies that $\gamma^c \in \Gamma_2\cdots\Gamma_m$. Each eigenvalue of $\gamma^c$ then has multiplicity $\geq n_1\geq2$; hence $f(\gamma^c)=0$ by Lemma \ref{Rprop} (a). This contradicts the given assumptions, and so each $\overline{\Gamma}_i$ is in fact regular.
The assumptions also imply that $\Gamma^{\rm der}$ is perfect and that $\Gamma = \Gamma^{\rm der}\cdot Z(\Gamma)$. Write any $\gamma\in\Gamma$ in the form $\gamma=\gamma'\zeta$ with $\gamma'\in\Gamma^{\rm der}$ and a scalar $\zeta\in Z(\Gamma)$. By construction $f(\alpha)$ is homogeneous of some degree $d$ in the coefficients of~$\alpha$; thus we have
$f(\gamma^c)\ =\ f(\gamma^{\prime c}\zeta^c) = f(\gamma^{\prime c})\cdot\zeta^{cd}$. Since this is not identically zero and $\gamma^{\prime c}\in\Gamma^{\rm der}$, it follows that $f$ is not identically zero on $\Gamma^{\rm der}$.
Together this shows that $\Gamma^{\rm der}$ satisfies the assumptions of Theorem \ref{finite_main}, and so the desired assertion follows.
\end{proof}
The following auxiliary results will help to determine the subfield $k'$ and the model $G'$ arising in Theorems \ref{finite_main} and \ref{finite_main1}:
\begin{prop}\label{fieldcontained}
Let $n\geq2$, and let $G$, $G'$ be models of $\SL_{n,L}$ over finite subfields $k, k'\subset L$, respectively. If $G'(k') \subset G(k)$, then $|k'|\le|k|$.
\end{prop}
\begin{proof}
Let $q:=|k|$, and set $\epsilon:=1$ if $G$ is split and $\epsilon:=-1$ otherwise. Likewise, let $q':=|k'|$, and set $\epsilon':=1$ if $G'$ is split and $\epsilon':=-1$ otherwise. Then \cite{HumphMod}, Table 1.6.1, implies that
$$(q')^{\frac{n(n-1)}{2}} \prod_{i=2}^n (q^{\prime i}-\epsilon^{\prime i})
\ =\ |G'(k')| \ \le\ |G(k)| \ =\
q ^{\frac{n(n-1)}{2}} \prod_{i=2}^n (q^i-\epsilon^i).$$
Suppose that $q'>q$. Since both numbers are powers of the same prime $p$, it follows that $q'\geq pq\geq2q$. For each $2\leq i\leq n$ we then have $q^{\prime i}-\epsilon^{\prime i} \geq 4q^i-1 > q^i-\epsilon^i$, and so the left hand side of the above inequality is in fact greater than the right hand side, which is impossible. Therefore $q'\leq q$, as desired.
\end{proof}
\begin{prop}\label{same_model}
Let $n\geq2$, and let $G$, $G'$ be models of $\SL_{n,L}$ over the same finite subfield $k\subset L$. If $G'(k) \subset G(k)$, then the models are equal and $G'(k) = G(k)$.
\end{prop}
\begin{proof}
Let $F,F': \SL_{n,L}\to\SL_{n,L}$ be the Frobenius maps corresponding to the models $G,G'$, respectively. Since they belong to the same finite field, there exists an automorphism $\alpha$ of $\SL_{n,L}$ such that $F=\alpha\circ F'$. Then for any $g'\in G'(k)$ we have $g'\in G(k)$ and hence $g' = F(g') = \alpha(F'(g')) = \alpha(g')$. In other words $\alpha$ is the identity on $G'(k)$.
Suppose first that $\alpha$ is an inner automorphism. Then it is conjugation by some element of $\SL_n(L)$. This element commutes with $G'(k)$, and since the standard representation of $G'(k)$ is irreducible by Proposition \ref{Irred_fgolt}, it must be a scalar. Then $\alpha$ is the identity, and so $F=F'$ and $G=G'$, as desired.
If $\alpha$ is not an inner automorphism, we must have $n\geq3$. Let $\mathfrak{psl}_n(L)$ denote the image of the natural homomorphism of Lie algebras $\sll_n(L)\to\pgll_n(L)$, and let $\rho$ denote the representation on $\mathfrak{psl}_n(L)$ induced by the adjoint representation of $\SL_{n,L}$. Since $n\geq3$, we know that $\rho$ factors through a faithful irreducible representation of $\PGL_{n,L}$. Moreover, by Proposition \ref{Irred_fgolt} it remains irreducible when restricted to $G'(k)$. On the other hand $\alpha$ induces an automorphism $\bar\alpha$ of $\mathfrak{psl}_n(L)$ that commutes with $\rho(G'(k))$. Thus $\bar\alpha$ is multiplication by a scalar, and therefore it commutes with the algebraic representation $\rho$. It follows that $\alpha$ induces the identity on $\PGL_{n,L}$. But then it is really an inner automorphism, contrary to the assumption.
\end{proof}
\begin{prop}\label{same_k_and_model}
The subfield $k'$ and the model $G'$ in Theorems \ref{finite_main} and \ref{finite_main1} are unique.
\end{prop}
\begin{proof}
Let $k$ be another finite subfield of $L$, and let $G$ be a model of $\SL_{n,L}$ over $k$, such that $\Gamma=G(k)$. Then applying Proposition \ref{fieldcontained} in both ways shows that $|k'|=|k|$. Thus $k'=k$, and then Proposition \ref{same_model} shows that $G'=G$, as desired.
\end{proof}
\subsection{Arbitrary finite groups} \label{irre_subgroups}
The following general result was established by Larsen and the second author in \cite{L-P}, Theorem 0.2:
\begin{thm} \label{L-P}
For any integer $n \geq 1$ there exists a constant $c_n$, such that for every field $L$, of arbitrary characteristic $p\geq0$, and every finite subgroup $\Gamma\subset \GL_n(L)$, there exist normal subgroups $\Gamma_3 \subset \Gamma_2 \subset \Gamma_1$ of $\Gamma$ satisfying:
\begin{enumerate}
\item[(a)] $[\Gamma : \Gamma_1]\leq c_n$,
\item[(b)] either $\Gamma_1=\Gamma_2$, or $p>0$ and $\Gamma_1/\Gamma_2$ is a direct product of finite simple groups of Lie type in characteristic $p$,
\item[(c)] $\Gamma_2/\Gamma_3$ is abelian of order not divisible by $p$, and
\item[(d)] either $\Gamma_3=\{1\}$, or $p>0$ and $\Gamma_3$ is a $p$-group.
\end{enumerate}
\end{thm}
We are interested in the following special case:
\begin{thm} \label{L-P3}
For any integer $n \geq 1$ there exists a constant $c'_n$, such that for every algebraically closed field $L$, of arbitrary characteristic $p\geq0$, and every finite subgroup $\Gamma\subset \GL_n(L)$ acting irreducibly, there exists a normal subgroup $\Gamma' \triangleleft\Gamma$ satisfying:
\begin{enumerate}
\item[(a)] $[\Gamma : \Gamma']\leq c'_n$, and
\item[(b)] either $\Gamma'=Z(\Gamma')$, or $p>0$ and $\Gamma'/Z(\Gamma')$ is a direct product of finite simple groups of Lie type in characteristic $p$.
\end{enumerate}
\end{thm}
\begin{proof}
Let $\Gamma_3 \subset \Gamma_2 \subset \Gamma_1\subset\Gamma$ be the subgroups furnished by Theorem \ref{L-P}. First we show that $\Gamma_3$ is trivial. By assumption, this is a unipotent normal subgroup of~$\Gamma$. Set $V:=L^n$. Then the subspace of invariants $V^{\Gamma_3}$ is non-zero and stabilized by $\Gamma$. Since $V$ is an irreducible representation of $\Gamma$, it follows that $V^{\Gamma_3}=V$. This means that $\Gamma_3=\{ 1\}$, as desired.
The triviality of $\Gamma_3$ implies that $\Gamma_2$ is an abelian group of order not divisible by $p$. Let $V=V_1 \oplus \ldots \oplus V_m$ be the isotypic decomposition under $\Gamma_2$, with all summands non-zero. The number of summands then satisfies $m\le n$. Since $\Gamma_2$ is normal in $\Gamma$, the summands are permuted by $\Gamma$, and so the permutation action corresponds to a homomorphism $f$ from $\Gamma$ to the symmetric group $S_m$ on $m$ letters. Set $\Gamma' := \Gamma_1\cap\mathop{\rm ker}(f)$. By construction this is a normal subgroup of index ${[\Gamma:\Gamma']} \allowbreak \leq {[\Gamma:\Gamma_1]\cdot |S_m|} \allowbreak \leq c_n\cdot n! =:c'_n$, where $c_n$ is the constant from Theorem \ref{L-P}.
On the other hand, the fact that $\Gamma_2$ acts by scalars on each $V_i$ and $\Gamma'$ stabilizes each $V_i$ implies that $\Gamma_2$ is contained in the center of~$\Gamma'$. Moreover $\Gamma'/\Gamma_2$ is the kernel of a homomorphism $\Gamma_1/\Gamma_2\to S_m$ induced by $f$. Since $\Gamma_1/\Gamma_2$ is a direct product of non-abelian finite simple groups, this kernel is simply a direct product of some of the factors. Thus either $\Gamma'=\Gamma_1=\Gamma_2$, or $p>0$ and $\Gamma'/\Gamma_2$ is a direct product of finite simple groups of Lie type in characteristic $p$. The last statement also implies that the inclusion $\Gamma_2\subset Z(\Gamma')$ must be an equality, and everything is proved.
\end{proof}
\section{Approximation in the special linear group}\label{approx}
\section{Subgroups of $\SL_n$ over a complete valuation ring} \label{approx}
Let $R$ be a complete discrete valuation ring of equal characteristic with finite residue field $k=R/\pp$ of characteristic $p$. Fix an integer $n\geq2$. In this section we consider closed subgroups of $\SL_n(R)$ for the topology induced by~$R$ and establish suitable conditions for such a closed subgroup to be equal to $\SL_n(R)$. We use successive approximation over the congruence filtration of $\SL_n(R)$, respectively of $\GL_n(R)$, whose subquotients are related to the adjoint representation. The case $p=n=2$ presents some special subtleties here, because the Lie bracket on $\sll_2$ in characteristic $2$ is not surjective. In Subsection \ref{trace_crit} we show how a certain non-triviality condition required earlier can be guaranteed using traces in the adjoint representation. The main results are Theorems \ref{strong_approx_SLn}, \ref{strong_approx_GLn}, \ref{trace_crit_1}, and \ref{trace_crit_2}.
\subsection{Adjoint representation}
We first collect a few general results on the cohomology and subgroups of the adjoint representation. Let $\gll_n$, $\sll_n$, $\pgll_n$ denote the Lie algebras of $\GL_n$, $\SL_n$, $\text{PGL}_n$, respectively. As usual we identify elements of $\gll_n$ with $n\times n$-matrices. Let $\mathfrak{c}$ denote the subspace of scalar matrices in $\gll_n$. For any field $k$ let $\mathfrak{psl}_n(k)$ denote the image of the natural homomorphism $\sll_n(k)\to\pgll_n(k)$.
\begin{prop}\label{cohom1}
For any finite field $k$ with $|k|>9$ and any subgroup $H$ of $\GL_n(k)$ that contains $\SL_n(k)$, we have
$$H^1(H,\pgll_n(k))=0.$$
\end{prop}
\begin{proof}
Consider the short exact sequence
$$0 \longrightarrow \mathfrak{c}(k) \longrightarrow \gll_n(k) \longrightarrow \pgll_n(k) \longrightarrow 0.$$
Its associated long exact cohomology sequence contains the portion
$$H^1(\SL_n(k),\gll_n(k)) \longrightarrow H^1(\SL_n(k),\pgll_n(k)) \longrightarrow H^2(\SL_n(k),\mathfrak{c}(k)).$$
Here the group on the left is trivial by \cite{TaZa}, Theorem 9. The group on the right classifies central extensions of $\SL_n(k)$ by $\mathfrak{c}(k)$; but since $\SL_n(k)$ has no central extensions by \cite{Ste2}, Theorem 1.1, if $|k|>9$, this group is also trivial. Thus the group in the middle is trivial. Finally, since $[H:\SL_n(k)]$ divides $[\GL_n(k):\SL_n(k)]= |k|-1$, it is prime to the characteristic of $k$. By \cite{CPS1}, Proposition 2.3 (g), the restriction map
$$H^1(H,\pgll_n(k)) \longrightarrow H^1(\SL_n(k),\pgll_n(k))$$
is therefore injective. Thus $H^1(H,\pgll_n(k))$ is trivial, as desired.
\end{proof}
The next proposition is an adaptation of \cite{PR2}, Proposition 2.1.
\begin{prop}\label{Egon2.1}
Let $n\geq 2$ and $k$ be a finite field with $|k|>9$. Let $H$ be an additive subgroup of $\mathfrak{gl}_n(k)$ that is invariant under conjugation by $\SL_n(k)$. Then either $H\subset\mathfrak{c}(k)$ or $\mathfrak{sl}_n(k)\subset H$.
\end{prop}
\begin{proof}
Let $W_0\subset\gll_n(k)$ denote the subgroup of diagonal matrices. For each pair of distinct indices $1\leq i,j\leq n$, let $W_{i,j}\subset\gll_n(k)$ denote the group of matrices with all entries zero except, possibly, in the position $(i,j)$. Then we have the decomposition
$$\mathfrak{gl}_n(k)=W_0\oplus \bigoplus_{i\not=j}W_{i,j}.$$
This decomposition is invariant under conjugation by the group of diagonal matrices $T(k)\subset\GL_n(k)$. Indeed, an element $t=\diag(t_1,\ldots,t_n)\in T(k)$ acts trivially on $W_0$ and by multiplication with $\chi_{i,j}(t) := t_i/t_j$ on $W_{i,j}$. Set $T'(k) := T(k)\cap\SL_n(k)$, and let ${\mathbb F}_p$ denote the prime field of $k$.
\begin{lem}\label{tolstoi}
The $W_{i,j}$ are non-trivial and irreducible viewed as representations of $T'(k)$ over ${\mathbb F}_p$. If $|k|>9$, they are pairwise non-isomorphic. Furthermore, they are permuted transitively under conjugation by the normalizer of $T'(k)$ in $\SL_n(k)$.
\end{lem}
\begin{proof}
The last assertion follows from the fact that the $W_{i,j}$ are permuted transitively by the permutation matrices and that every permutation matrix can be moved into $\SL_n(k)$ by changing the sign of at most one entry.
For the first assertion it thus suffices to consider $W_{1,2}$. The calculation $\chi_{1,2}(\diag(x,x^{-1},1,\ldots,1))=x^2$ shows that $(k^\times)^2 \subset \chi_{1,2}(T'(k)) \subset k^\times$. Since every element of a finite field $k$ can be written as a sum of two squares, this subgroup generates $k$ as an ${\mathbb F}_p$-algebra. As $W_{1,2}$ is a one-dimensional vector space over $k$, it is therefore irreducible as a representation of $T'(k)$ over ${\mathbb F}_p$. Since $(k^\times)^2\not=\{1\}$ by assumption, this representation is non-trivial.
For the remaining assertion suppose that two distinct $W_{i,j}$ and $W_{i',j'}$ are isomorphic as representations of $T'(k)$ over ${\mathbb F}_p$. This means that $\chi_{i,j}|T'(k) = \chi_{i',j'}^{p^m}|T'(k)$ for some $m\ge0$. Without loss of generality we may assume that $(i',j')=(1,2)$. Suppose first that $(i,j)=(2,1)$. Then by applying the equation to elements of the form $\diag(x,x^{-1},1,\ldots,1)$ we find that $x^{-2} = x^{2p^m}$ for all $x\in k^\times$. By an explicit calculation that we leave to the reader, this is not possible if $|k|>9$ (and this bound cannot be improved if $n=2$!). If $i,j>2$, the element $\diag(x,x^{-1},1,\ldots,1)$ acts as multiplication by $x^2$ on $W_{1,2}$ and trivially on $W_{i,j}$. Since $(k^\times)^2\not=\{1\}$ by assumption, again the representations cannot be isomorphic. If precisely one of $i,j$ is $\le 2$, we may assume that the other is~$3$. Then the element $\diag(x,x,x^{-2},1,\ldots,1)$ acts trivially on $W_{1,2}$ and as multiplication by $x^{\pm3}$ on $W_{i,j}$. Since $(k^\times)^3\not=\{1\}$ by assumption, we again obtain a contradiction.
\end{proof}
Now let $H$ be as in the proposition. Suppose first that $H\subset W_0$. Consider an arbitrary element $h=\diag(h_1,\ldots,h_n)\in H$. Take distinct indices $1\leq i,j\leq n$ and let $g \in \SL_n(k)$ be the matrix with entries $1$ on the diagonal and in the position $(i,j)$, and entries $0$ elsewhere. Then $ghg^{-1}-h$ is the matrix with entry $h_i-h_j$ in the position $(i,j)$ and entries $0$ elsewhere. Since $H\subset W_0$, it follows that $h_i=h_j$. Varying $i$ and $j$ we deduce that $h$ is a scalar matrix, i.e., that $H\subset\mathfrak{c}(k)$.
If $H\not\subset W_0$, Lemma \ref{tolstoi} implies that $H$ contains at least one, and hence all $W_{i,j}$. Consider the trace form
$$\mathfrak{gl}_n(k) \times \mathfrak{gl}_n(k) \to {\mathbb F}_p,
\quad (X,Y)\mapsto \Tr_{k/{\mathbb F}_p}\Tr(XY),$$
which is a perfect ${\mathbb F}_p$-bilinear pairing invariant under $\SL_n(k)$. Then $H$ contains the orthogonal complement $W_0^\perp$ of $W_0$, and since taking orthogonal complements reverses inclusion relations, the orthogonal complement $H^\perp$ of $H$ is contained in $W_0$. By construction $H^\perp$ is again an $\SL_n(k)$-invariant subgroup; hence the preceding arguments show that $H^\perp\subset\mathfrak{c}(k)$. Therefore
$\mathfrak{sl}_n(k) = \mathfrak{c}(k)^\perp \subset H$, as desired.
\end{proof}
\begin{cor}\label{Egon2.1_cor}
Let $n\geq 2$ and $k$ be a finite field with $|k|>9$. Let $H$ be a non-zero additive subgroup of $\mathfrak{pgl}_n(k)$ that is invariant under conjugation by $\SL_n(k)$. Then $H$ contains $\mathfrak{psl}_n(k)$.
\end{cor}
\begin{proof}
Apply Proposition \ref{Egon2.1} to the inverse image $\tilde H\subset\mathfrak{gl}_n(k)$ of $H$. Since $H$ is non-trivial, we have $\tilde H\not\subset\mathfrak{c}(k)$ and hence $\mathfrak{sl}_n(k)\subset \tilde H$. Therefore $\mathfrak{psl}_n(k)\subset H$, as desired.
\end{proof}
\subsection{Successive approximation} \label{succc}
The congruence filtration of $\GL_n(R)$ consists of the subgroups
$$G^i := \{ g\in\GL_n(R)\mid g\equiv{\rm Id}_n\mathrel{\rm mod}\pp^i\}$$
for all $i\geq0$. Their successive subquotients possess natural isomorphisms
$$G^{[i]}\ :=\ G^i/G^{i+1}\ \cong\
\left\{\begin{array}{ll}
\GL_n(k) & \hbox{if $i=0$,}\\[5pt]
\gll_n(\pp^i/\pp^{i+1}) & \hbox{if $i>0$,}
\end{array}\right.$$
where the second isomorphism is given by $[{\rm Id}_n+X] \mapsto [X]$. For any subgroup $H$ of $\GL_n(R)$ we define $H^i:= H\cap G^i$ and $H^{[i]}:= H^i/H^{i+1}$ and identify the latter with a subgroup of $\GL_n(k)$ or $\gll_n(\pp^i/\pp^{i+1})$, respectively. For example, the induced congruence filtration of $G' := \SL_n(R)$ has the successive subquotients
$$G^{\prime[i]}\ \cong\
\left\{\begin{array}{ll}
\SL_n(k) & \hbox{if $i=0$,}\\[5pt]
\sll_n(\pp^i/\pp^{i+1}) & \hbox{if $i>0$.}
\end{array}\right.$$
As a preparation we show:
\begin{lem} \label{camilla}
Assume that $|k|>9$. Let $H$ be a subgroup of $\GL_n(R)$, and let $H'$ be a normal subgroup of $H$ that is contained in $\SL_n(R)$. Assume that $\SL_n(k) \subset H^{\prime[0]}$ and that $H^{[1]}$ contains a non-scalar matrix. Then we have $H^{\prime[1]}=\sll_n(\pp/\pp^2)$.
\end{lem}
\begin{proof}
Pick a non-scalar matrix $X\in H^{[1]} \subset \gll_n(\pp/\pp^2)$. Since $\pp/\pp^2\cong k$, Proposition \ref{Egon2.1} implies that the $k$-vector subspace generated by all $\SL_n(k)$-conjugates of $X$ contains $\sll_n(\pp/\pp^2)$. Thus there exists $\gamma\in\SL_n(k)$ such that $X$ and $\gamma X\gamma^{-1}$ and $\Id_n$ are $k$-linearly independent. Choose elements $h\in H^1$ and $h'\in H'$ that are mapped to $X$ and~$\gamma$, respectively, i.e., that satisfy $h\equiv \Id_n+X\mathrel{\rm mod}\pp^2$ and $h'\equiv\gamma\mathrel{\rm mod}\pp$. Then the commutator $h'hh^{\prime-1}h^{-1}$ lies in $H'$ and is congruent to $\Id_n + \gamma X\gamma^{-1}-X \mathrel{\rm mod}\pp^2$. By construction $\gamma X\gamma^{-1}-X \mathrel{\rm mod}\pp^2$ is not scalar; hence $H^{\prime[1]}\subset\sll_n(\pp/\pp^2)$ contains a non-scalar matrix.
On the other hand the isomorphism $G^{[1]}\stackrel{\sim}{\to}\gll_n(\pp/\pp^2)$, $[{\rm Id}_n{+}X] \mapsto [X]$ is equivariant under conjugation by $\GL_n(R)$. This conjugation action factors through an action of $\GL_n(k)$. In the present situation it follows that $H^{\prime[1]}\subset\sll_n(\pp/\pp^2)$ is an additive subgroup that is invariant under conjugation by $\SL_n(k)\subset H^{\prime[0]}$. Since by assumption it also contains a non-scalar matrix, Proposition \ref{Egon2.1} implies that $H^{\prime[1]} = \sll_n(\pp/\pp^2)$, as desired.
\end{proof}
\begin{thm}\label{strong_approx_SLn}
Assume that $|k|>9$. Let $H$ be a closed subgroup of $\SL_n(R)$ such that $H^{[0]}=\SL_n(k)$ and that $H^{[1]}$ contains a non-scalar matrix. Then $H=\SL_n(R)$.
\end{thm}
Before proving this we derive two consequences. For a closed subgroup $H$ of $\GL_n(R)$ we let $H^{\rm der}$ denote the closure of the derived group of $H$ for the topology induced by~$R$. (Probably the derived group is already closed, but we neither need nor want to worry about that.)
\begin{thm}\label{strong_approx_GLn}
Assume that $|k|>9$. Let $H$ be a closed subgroup of $\GL_n(R)$ such that $\SL_n(k)\subset H^{[0]}$ and that $H^{[1]}$ contains a non-scalar matrix.
Then $H^{\rm der}=\SL_n(R)$.
\end{thm}
\begin{proof}
Since $|k|>3$, the group $\SL_n(k)$ is perfect, and so the assumption $\SL_n(k)\subset H^{[0]}$ implies that $(H^{\rm der})^{[0]} = \SL_n(k)$. Since $H^{[1]}$ contains a non-scalar matrix, applying Lemma \ref{camilla} with $H'=H^{\rm der}$ thus shows that $(H^{\rm der})^{[1]}=\sll_n(\pp/\pp^2)$; in particular it contains a non-scalar matrix. The desired assertion results by applying Theorem \ref{strong_approx_SLn} to $H^{\rm der}$.
\end{proof}
\begin{prop}\label{normal_sa}
Assume that $|k|>9$. Then every closed normal subgroup $H\subset\SL_n(R)$ satisfying $H^{[0]}=\SL_n(k)$ is equal to $\SL_n(R)$.
\end{prop}
\begin{proof}
Applying Lemma \ref{camilla}
with $(H,\SL_n(R))$ in place of $(H',H)$
shows that $H^{[1]}$ contains a non-scalar matrix. The desired assertion now follows directly from Theorem \ref{strong_approx_SLn}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{strong_approx_SLn}.]
The proof of this will extend to the end of the next subsection. Let $H$ satisfy the assumptions of Theorem \ref{strong_approx_SLn}. Since $H$ is a closed subgroup of $\SL_n(R)$, the desired assertion is equivalent to $H^{[i]}= \SL_n(R)^{[i]} $ for all $i\geq 0$. By assumption this holds already for $i=0$. Applying Lemma \ref{camilla} with $H'=H$ implies:
\begin{lem}\label{lemma12.6}
We have $H^{[1]}=\sll_n(\pp/\pp^2)$.
\end{lem}
\begin{lem}\label{lemma12.7}
If $(p,n)\neq (2,2)$, then $H^{[i]}=\sll_n(\pp^i/\pp^{i+1})$ for all $i\geq 1$.
\end{lem}
\begin{proof}
We use induction on $i$, the case $i=1$ being covered by Lemma \ref{lemma12.6}. So suppose that the assertion holds for some $i\geq1$. Consider the commutator map $(g,g')\mapsto gg'g^{-1}g^{\prime-1}$ on $\GL_n(R)$. Direct calculation shows that it induces a map $G^1\times G^i\to G^{i+1}$ and hence a map $G^{[1]}\times G^{[i]} \to G^{[i+1]}$, which under the above isomorphisms corresponds to the Lie bracket
\begin{eqnarray*}
[\ \:,\ ] :\ \sll_n(\pp/\pp^2) \times \sll_n(\pp^i/\pp^{i+1})
\!\!\!&\longrightarrow&\!\!\! \sll_n(\pp^{i+1}/\pp^{i+2}), \\
(X,Y) &\mapsto& XY-YX.
\end{eqnarray*}
Since $(p,n)\neq (2,2)$, the image of this pairing generates $\sll_n(\pp^{i+1}/\pp^{i+2})$ as an additive group, for instance by \cite{PSTAX}, Proposition 1.2 (a).
On the other hand, by construction the pairing sends the subset $H^{[1]} \times H^{[i]}$ to the subgroup $H^{[i+1]}$.
The equality in the source thus implies equality in the target, and so the assertion holds for $i+1$, as desired.
\end{proof}
This proves Theorem \ref{strong_approx_SLn} in the case $(p,n)\neq (2,2)$. The remaining case is more complicated, because the image of the Lie bracket on $\sll_2$ in characteristic $2$ does not generate $\sll_2$. We deal with this in the next subsection.
\subsection{Successive approximation in the case $p=n=2$}
Keeping the assumptions of Theorem \ref{strong_approx_SLn}, we now consider the case $p=n=2$.
Let $\overline{G}{}^i$ for all $i\ge0$ denote the subgroups in the congruence filtration of $\PGL_2(R)$. Thus $\overline{G}{}^i$ consists of all elements of $\PGL_2(R)$ whose images in the adjoint representation are congruent to the identity modulo $\pp^i$. Their successive subquotients possess natural isomorphisms
$$\overline{G}{}^{[i]}\ :=\ \overline{G}{}^i/\overline{G}{}^{i+1}\ \cong\
\left\{\begin{array}{ll}
\PGL_2(k) & \hbox{if $i=0$,}\\[5pt]
\pgll_2(\pp^i/\pp^{i+1}) & \hbox{if $i>0$.}
\end{array}\right.$$
We will compare the congruence filtration of $\overline{G}$ with that of $G' := \SL_n(R)$. Let $\pi$ denote the projection $\SL_2\to\PGL_2$. For $i>0$ the induced map $G^{\prime[i]}\to\overline{G}{}^{[i]}$ corresponds to the derivative $d\pi: \sll_2(\pp^i/\pp^{i+1}) \to \pgll_2(\pp^i/\pp^{i+1})$. Let $\mathfrak{psl}_2(\pp^i/\pp^{i+1})$ denote its image. Being in characteristic $2$, we have short exact sequences
$$\xymatrix@R-18pt{
0\ar[r] & \mathfrak{c}(\pp^i/\pp^{i+1})
\ar[r] & \sll_2(\pp^i/\pp^{i+1})
\ar[r] & \psll_2(\pp^i/\pp^{i+1}) \ar[r] & 0, \\
0\ar[r] & \psll_2(\pp^i/\pp^{i+1})
\ar[r] & \pgll_2(\pp^i/\pp^{i+1})
\ar[r]^{(*)} & \mathfrak{c}^*(\pp^i/\pp^{i+1}) \ar[r] & 0, \\}$$
where $\mathfrak{c}(\pp^i/\pp^{i+1}) \cong \mathfrak{c}^*(\pp^i/\pp^{i+1}) \cong \pp^i/\pp^{i+1}$ and the homomorphism $(*)$ is induced by the trace on $\gll_2$.
The first few layers of $\SL_2(R)$ and $\PGL_2(R)$ are related as follows. Since $k$ has characteristic $2$, the homomorphism $\SL_2(k) \to \PGL_2(k)$ is an isomorphism of abstract groups. Therefore $\pi^{-1}(\overline{G}{}^1) = G^{\prime1}$. Next consider the subgroup $G^{\prime2-} := \pi^{-1}(\overline{G}{}^2) \subset \SL_2(R)$. What we have just said implies that $G^{\prime2} \subset G^{\prime2-} \subset G^{\prime1}$. By construction $\pi$ induces a natural homomorphism $G^{\prime2-}/G^{\prime3} \to \overline{G}{}^{[2]}$.
\begin{lem}\label{tempe}
There is an isomorphism $\mathfrak{c}(\pp/\pp^2) \stackrel{\sim}{\longrightarrow} \mathfrak{c}^*(\pp^2/\pp^3)$ such that the following diagram commutes:
$$\xymatrix@R-20pt{
0\ar[r] & G^{\prime2}/G^{\prime3} \ar[r] \ar@{=}[dd]^-\wr &
G^{\prime2-}/G^{\prime3} \ar[r] \ar[ddd]_{[\pi]} &
G^{\prime2-}/G^{\prime2} \ar[r] \ar@{=}[dd]^-\wr & 0 \\
&&&& \\
& \sll_2(\pp^2/\pp^3) \ar@{->>}[ddd]_{d\pi} &&
\mathfrak{c}(\pp/\pp^2) \ar[ddd]_\cong & \\
&& \overline{G}{}^{[2]} \ar@{=}[dd]^-\wr && \\
&&&& \\
0\ar[r] & \psll_2(\pp^2/\pp^3) \ar[r] &
\pgll_2(\pp^2/\pp^3) \ar[r] &
\mathfrak{c}^*(\pp^2/\pp^3) \ar[r] & 0\rlap{.} \\}$$
\end{lem}
\begin{proof}
The commutativity on the left hand side is already clear. Let $\varpi\in\pp$ be a uniformizer. An easy calculation shows that $G^{\prime2-}$ consists of the matrices
$$g\ =\ \left(\!\begin{array}{cc}1+\varpi x & 0 \\[5pt]
0 & (1+\varpi x)^{-1} \end{array}\!\!\right)
\cdot g_2$$
for all $x\in R$ and $g_2\in\SL_2(R)$ which satisfy $g_2\equiv{\rm Id_2}\mathrel{\rm mod}\pp^2$. The residue class of $g$ in $\mathfrak{c}(\pp/\pp^2) \cong \pp/\pp^2$ is then $\varpi x\mathrel{\rm mod}\pp^2$. On the other hand $\pi(g) \in \PGL_2(R)$ is also the image of the matrix
$$\tilde g\ :=\ (1+\varpi x)\cdot g\ =\
\left(\!\begin{array}{cc}1+\varpi^2 x^2 & 0 \\[5pt]
0 & 1 \end{array}\!\right)
\cdot g_2 \ \in\ \GL_2(R).$$
Since $\tilde g\equiv{\rm Id_2}\mathrel{\rm mod}\pp^2$, its image in $\mathfrak{c}^*(\pp^2/\pp^3) \cong \pp^2/\pp^3$ is simply $\Tr(\tilde g-\Id_2)$. But the assumptions $g_2\in\SL_2(R)$ and $g_2\equiv{\rm Id_2}\mathrel{\rm mod}\pp^2$ imply that $\Tr(g_2-\Id_2) \in \pp^3$, and so an easy calculation shows that
$$\Tr(\tilde g-\Id_2)\ \equiv\ \varpi^2 x^2 + \Tr(g_2-\Id_2)
\ \equiv\ \varpi^2 x^2 \mod \pp^3.$$
Thus the diagram in question commutes with the map
$$\mathfrak{c}(\pp/\pp^2) \longrightarrow \mathfrak{c}^*(\pp^2/\pp^3),
\quad (\varpi x\mathrel{\rm mod}\pp^2) \mapsto (\varpi^2 x^2\mathrel{\rm mod}\pp^3).$$
Up to multiplication by $\varpi$, respectively $\varpi^2$, this is simply the Frobenius map $x\mapsto x^2$ on the finite field $k$. It is therefore an isomorphism, as desired.
\end{proof}
Now let $\overline{H}$ denote the image of $H$ in $\PGL_2(R)$. Define $\overline{H}{}^i:= \overline{H}\cap \overline{G}{}^i$ and $\overline{H}{}^{[i]}:= \overline{H}{}^i/\overline{H}{}^{i+1}$ and identify the latter with a subgroup of $\PGL_2(k)$ or $\pgll_2(\pp^i/\pp^{i+1})$, respectively. The projection $\pi: \SL_2(R)\twoheadrightarrow\PGL_2(R)$ induces homomorphisms $H^i\to\overline{H}{}^i$ and $H^{[i]}\to\overline{H}{}^{[i]}$.
Consider the subgroup $H^{2-} := H\cap G^{\prime2-}$. Then by construction the natural homomorphisms $H^{2-} \to \overline{H}{}^2 \to \overline{H}{}^{[2]}$ are surjective.
\begin{lem}\label{brennan}
We have $\overline{H}{}^{[2]}=\pgll_2(\pp^2/\pp^3)$.
\end{lem}
\begin{proof}
By construction $H^{2-}$ can be described equivalently as the inverse image of $\mathfrak{c}(\pp/\pp^2) \subset \sll_2(\pp/\pp^2)$ in $H^1$. Thus Lemma \ref{lemma12.6} implies that $H^{2-}$ surjects to $\mathfrak{c}(\pp/\pp^2)$.
Thus Lemma \ref{tempe} implies that $\overline{H}{}^{[2]}$ surjects to $\mathfrak{c}^*(\pp^2/\pp^3)$. In particular $\overline{H}{}^{[2]}$ is non-zero.
Furthermore, since the embedding $\overline{H}{}^{[2]} \hookrightarrow \pgll_2(\pp^2/\pp^3)$ is equivariant under conjugation by $H$, its image is invariant under conjugation by $H^{[0]}=\SL_2(k)$. Thus Corollary \ref{Egon2.1_cor} implies that $\psll_2(\pp^2/\pp^3)\subset \overline{H}{}^{[2]}$. Combined with the surjection $\overline{H}{}^{[2]} \twoheadrightarrow \mathfrak{c}^*(\pp^2/\pp^3)$ this implies the desired equality.
\end{proof}
\begin{lem}\label{seeley}
We have $H^{[2]}=\sll_2(\pp^2/\pp^3).$
\end{lem}
\begin{proof}
Lemma \ref{brennan} says that $H^{2-}$ surjects to $\pgll_2(\pp^2/\pp^3)$. Combined with Lemma \ref{tempe} this implies that $H^2$ surjects to $\psll_2(\pp^2/\pp^3)$. In particular $H^{[2]}$ contains a non-scalar matrix. Being invariant under $H^{[0]}=\SL_2(k)$ and contained in $\sll_2(\pp^2/\pp^3)$, by Proposition \ref{Egon2.1} it is therefore equal to $\sll_2(\pp^2/\pp^3)$.
\end{proof}
\begin{lem}\label{lemma12.13}
We have $H^{[i]}=\sll_2(\pp^i/\pp^{i+1})$ for all $i\geq1$.
\end{lem}
\begin{proof}
By Lemmas \ref{lemma12.6} and \ref{seeley} we already know this assertion for $i=1,2$. Suppose that the assertion holds for some $i\geq1$. The commutator map $(g,g')\mapsto gg'g^{-1}g^{\prime-1}$ on $\GL_2(R)$ induces a map $\PGL_2(R) \times \SL_2(R) \to \SL_2(R)$. Direct calculation shows that this in turn induces a map $\overline{G}{}^2\times G^{\prime i}\to (G')^{i+2}$ and hence a map $\overline{G}{}^{[2]}\times G^{\prime[i]}\to G^{\prime[i+2]}$, which under the given isomorphisms is obtained from the Lie bracket by
\begin{eqnarray*}
[\ \:,\ ] :\ \pgll_2(\pp^2/\pp^3) \times \sll_2(\pp^i/\pp^{i+1})
\!\!\!&\longrightarrow&\!\!\! \sll_2(\pp^{i+2}/\pp^{i+3}), \\
\bigl((X\mathrel{\rm mod}\mathfrak{c}),Y\bigr) &\mapsto& XY-YX.
\end{eqnarray*}
Another direct calculation, or looking up \cite{PSTAX}, Proposition 1.2 (b), shows that the image of this pairing generates $\sll_2(\pp^{i+2}/\pp^{i+3})$ as an additive group.
On the other hand, by construction the pairing sends the subset $\overline{H}{}^{[2]} \times H^{[i]}$ to the subgroup $H^{[i+2]}$. Since $\overline{H}{}^{[2]}=\pgll_2(\pp^2/\pp^3)$ by Lemma \ref{brennan}, the equality in the source thus implies equality in the target, and so the assertion holds for $i+2$. By separate induction over all even, resp.\ odd integers, the assertion follows for all $i\geq1$.
\end{proof}
Lemma \ref{lemma12.13} finishes the proof of Theorem \ref{strong_approx_SLn} in the remaining case $p=n=2$.
\end{proof}
\subsection{Trace criteria} \label{trace_crit}
In this subsection we show how the assumption in Theorem \ref{strong_approx_GLn} that $H^{[1]}$ contain a non-scalar matrix can be guaranteed using traces in the adjoint representation.
We keep the notations of Subsection \ref{succc}, assume that $|k|>9$, and consider a closed subgroup $H$ of $\GL_n(R)$, such that $\SL_n(k)\subset H^{[0]}$. In other words, the remaining assumptions in Theorem \ref{strong_approx_GLn} are met.
Recall that $R$ is a complete valuation ring of equal characteristic. Thus the projection $R\twoheadrightarrow k$ possesses a unique splitting $k\hookrightarrow R$. Via this splitting we can view $\GL_n(k)$ as a subgroup of $\GL_n(R)$.
Let $G^{2-}$ denote the group of all matrices $g\in\GL_n(R)$ that are congruent to the identity modulo $\pp$ and congruent to a scalar modulo $\pp^2$. Then $G^2\subset G^{2-}\subset G^1$, and $G^{2-}/G^2 \cong \mathfrak{c}(\pp/\pp^2)$.
\begin{lem} \label{booth}
If $H^{[1]}$ contains only scalar matrices, then up to conjugation by an element of $\GL_n(R)$ we have $H\subset\GL_n(k)\cdot G^{2-}$.
\end{lem}
\begin{proof}
Consider the commutative diagram with exact rows
$$\xymatrix@R-7pt{
0\ar[r] & \pgll_n(\pp/\pp^2) \ar[r]^-{1+(\ )} \ar@{=}[d]^-\wr &
\GL_n(R/\pp^2)/(1+\mathfrak{c}(\pp/\pp^2)) \ar[r] \ar@{=}[d]^-\wr &
\GL_n(k) \ar[r] \ar@{=}[d]^-\wr & 1 \\
1\ar[r] & G^1/G^{2-} \ar[r] &
G^0/G^{2-} \ar[r] &
G^{[0]} \ar[r] & 1 \\
1\ar[r] & (H\cap G^1)/(H\cap G^{2-}) \ar[r] \ar@{^{ (}->}[u] &
H/(H\cap G^{2-}) \ar[r] \ar@{^{ (}->}[u] &
H^{[0]} \ar[r] \ar@{^{ (}->}[u] & 1\rlap{.} \\}$$
The assumption $H^{[1]} \subset\mathfrak{c}(\pp/\pp^2)$ means that $H^1 = H\cap G^1\subset G^{2-}$. Thus the group on the lower left is trivial, and so the group in the lower middle defines a splitting $H^{[0]} \to G^0/G^{2-}$. We compare this splitting with the splitting induced by the inclusions $H^{[0]} \subset \GL_n(k) \subset \GL_n(R)$. These two splittings differ by a $1$-cocycle $H^{[0]} \to \pgll_n(\pp/\pp^2)$. But since $\SL_n(k)\subset H^{[0]} \subset \GL_n(k)$, Proposition \ref{cohom1} shows that this cocycle is a coboundary. This means that the splittings are conjugate by an element coming from $\pgll_n(\pp/\pp^2)$, and the lemma follows.
\end{proof}
Let $\Ad$ denote the adjoint representation of $\GL_n$, and let $\Tr\Ad(H)$ denote the subset $\{\Tr\Ad(h)\mid h\in H\} \subset R$. Recall that $H^{\rm der}$ denotes the closure of the derived group of $H$.
\begin{thm} \label{trace_crit_1}
Assume that $|k|>9$. Let $H$ be a closed subgroup of $\GL_n(R)$ such that $\SL_n(k)\subset H^{[0]}$ and $\Tr\Ad(H)$ topologically generates the ring $R$. Then $H^{\rm der}=\SL_n(R)$.
\end{thm}
\begin{proof}
By Theorem \ref{strong_approx_GLn} it suffices to show that $H^{[1]}$ contains a non-scalar matrix. If that is not the case, by Lemma \ref{booth} we may assume that $H \subset \GL_n(k)\cdot\nobreak G^{2-}$.
Consider any element $h\in H$. Then by the definition of $G^{2-}$, its image $\Ad(h)$ is the product of an element of $\PGL_n(k)$ with a matrix that is congruent to the identity modulo $\pp^2$. Its trace is therefore congruent to an element of $k$ modulo $\pp^2$; in other words we have $\Tr\Ad(h) \in k+\pp^2$. This contradicts the assumption on $\Tr\Ad(H)$.
\end{proof}
If $p=n=2$, the assumption on traces in Theorem \ref{trace_crit_1} may fail in interesting cases (compare Proposition \ref{Rtradsmall}), although the conclusion is satisfied. This is due to the fact that the representation of $\GL_2$ on $\psll_2$ in characteristic $2$ is isomorphic to the pullback under $\Frob_2$ of the standard representation twisted with the inverse of the determinant, which implies that $\Tr\Ad(g) = \Tr(g)^2\cdot\det(g)^{-1} + 2$ for every $g\in \GL_2(R)$. Thus if $\det(H)$ consists of squares, which happens in particular for $H=\SL_2(R)$, the subset $\Tr\Ad(H)$ is entirely contained in the subring $R^2 := \{x^2\mid x\in R\}$. The following result provides a suitable substitute in that case:
\begin{thm} \label{trace_crit_2}
Assume that $|k|>9$ and $p=2$. Let $H$ be a closed subgroup of $\GL_2(R)$ such that $\SL_2(k)\subset H^{[0]}$. Let $H'\subset H$ denote the intersection of all closed subgroups of index $2$, and assume that $\Tr\Ad(H')$ topologically generates the subring $R^2 := \{x^2\mid x\in \nobreak R\}$.
Then $H^{\rm der}=\SL_2(R)$.
\end{thm}
\begin{proof}
Again by Theorem \ref{strong_approx_GLn} it suffices to show that $H^{[1]}$ contains a non-scalar matrix. If that is not the case, by Lemma \ref{booth} we may assume that $H \subset \GL_n(k)\cdot\nobreak G^{2-}$. Let $\varpi\in\pp$ be a uniformizer. Then every element of $H$ can be written in the form $h=\gamma\cdot g_2\cdot (1+\varpi x)$ with $\gamma\in \GL_2(k)$ and $g_2\in G^2$ and $x\in R$.
\begin{lem} \label{gormogon}
There exists a homomorphism $f:H\to\pp^2/\pp^3$ satisfying $f(h)= \Tr(g_2-\Id_2)\mathrel{\rm mod}\pp^3$ for any element $h=\gamma\cdot g_2\cdot (1+\varpi x)$ of the above form.
\end{lem}
\begin{proof}
Consider another element $h'=\gamma'\cdot g'_2\cdot (1+\varpi x')\in H$ with $\gamma'\in \GL_2(k)$ and $g'_2\in G^2$ and $x'\in R$.
To show that $f$ is well-defined, we must prove that $\Tr(g'_2-\Id_2) \equiv \Tr(g_2-\nobreak\Id_2)\allowbreak \mathrel{\rm mod}\pp^3$ whenever $h'=h$. But $h'=h$ implies that $\gamma'=\gamma$ and hence $g'_2 = g_2\cdot(1+\varpi y)$ for some $y\in R$. Therefore $\Tr(g'_2-\Id_2) = \Tr(g_2-\Id_2) + \Tr(g_2)\cdot\varpi y$. Since $g_2$ is congruent to the identity matrix modulo $\pp^2$, its trace is congruent to $2\mathrel{\rm mod}\pp^2$, i.e., congruent to $0\mathrel{\rm mod}\pp^2$. Thus $\Tr(g_2)\cdot\varpi y\in\pp^3$, and so the map is well-defined.
To show that $f$ is a homomorphism, observe that
\begin{eqnarray*}
h'h &=& \gamma'\cdot g'_2\cdot (1+\varpi x') \cdot \gamma\cdot g_2\cdot (1+\varpi x) \\
&=& (\gamma'\gamma) \cdot (\gamma^{-1}g'_2\gamma\cdot g_2)\cdot (1+\varpi x') \cdot (1+\varpi x)
\end{eqnarray*}
with $\gamma'\gamma \in \GL_2(k)$ and $\gamma^{-1}g'_2\gamma\cdot g_2 \in G^2$ and $(1+\varpi x') \cdot (1+\varpi x)= (1+\varpi y)$ for some $y\in R$. Thus
$f(h'h) = \Tr(\gamma^{-1}g'_2\gamma\cdot g_2-\Id_2)\mathrel{\rm mod}\pp^3$. Write this trace in the form
$$\Tr\bigl((\gamma^{-1}g'_2\gamma-\Id_2)(g_2-\Id_2)\bigr)
+ \Tr(\gamma^{-1}g'_2\gamma-\Id_2)
+ \Tr(g_2-\Id_2).$$
Here the first summand lies in $\pp^4$, because $\gamma^{-1}g'_2\gamma-\Id_2 \equiv g_2-\Id_2 \equiv 0$ modulo $\pp^2$, and the second summand is equal to $\Tr(g'_2-\Id_2)$, because the trace is invariant under conjugation. Therefore
$$\Tr(\gamma^{-1}g'_2\gamma\cdot g_2-\Id_2)
\ \equiv\ \Tr(g'_2-\Id_2)+ \Tr(g_2-\Id_2) \mod\pp^3,$$
and so $f(h'h) = f(h')+f(h)$, as desired.
\end{proof}
Since $\pp^2/\pp^3$ is an elementary abelian $2$-group, Lemma \ref{gormogon} implies that the restriction of $f$ to the subgroup $H'$ is trivial. In other words, for every element $h=\gamma\cdot g_2\cdot (1+\varpi x) \in H'$ with $\gamma\in \GL_2(k)$ and $g_2\in G^2$ and $x\in R$ we have $\Tr(g_2-\Id_2)\in\pp^3$. But for any such element we have
\begin{eqnarray*}
\Tr\Ad(h) &=& \Tr\Ad(\gamma g_2) \ =\
\Tr(\gamma g_2)^2\cdot\det(\gamma g_2)^{-1} + 2 \\
&=& \Tr(\gamma g_2)^2\cdot\det(\gamma)^{-1}\cdot\det(g_2)^{-1} + 2.
\end{eqnarray*}
Here the matrix $\gamma g_2$ has coefficients in $k+\pp^2$; hence its trace lies in $k+\pp^2$, and the first factor lies in $k+\pp^4$. Since $\gamma\in\GL_2(k)$, the second factor lies in $k^\times$. Moreover, the fact that $g_2\equiv\Id_2\mathrel{\rm mod}\pp^2$ implies that $\det(g_2) \equiv 1+\Tr(g_2-\Id_2) \mathrel{\rm mod}\pp^2$. But we have just seen that $\Tr(g_2-\Id_2)\in\pp^3$, and so $\det(g_2)$ and hence the third factor lies in $1+\pp^3$. Together we find that $\Tr\Ad(h)$ lies in $k+\pp^3$. This contradicts the assumption on $\Tr\Ad(H')$, and so Theorem \ref{trace_crit_2} is proved.
\end{proof}
\section{Preliminary results on Drinfeld modules} \label{known_results}
In this section we list some known results about Drinfeld modules or adapt them slightly, and create the setup on which the proof in Section \ref{sect_surjective} is based. From Subsection \ref{tatess} onwards we will restrict ourselves to the case of special characteristic. For the general theory of Drinfeld modules see Drinfeld \cite{DriI}, Deligne and Husem\"oller \cite{DeHu}, Hayes \cite{HayEx}, or Goss \cite{GossFFA}.
\subsection{Endomorphisms rings} \label{endos}
Let $\mathbb{F}_p$ denote the finite field of prime order~$p$. Let $F$ be a finitely generated field of transcendence degree $1$ over~$\mathbb{F}_p$. Let $A$ be the ring of elements of $F$ which are regular outside a fixed place $\infty$ of~$F$.
Let $K$ be another finitely generated field over $\mathbb{F}_p$ of arbitrary transcendence degree.
Then the endomorphism ring of the algebraic additive group ${\mathbb G}_{a,K}$ over $K$ is the non-commutative polynomial ring in one variable $K\{\tau\}$, where $\tau$ represents the endomorphism $u \mapsto u^p$ and satisfies the commutation relation $\tau u =u^p \tau$ for all $u\in K$.
Consider a Drinfeld $A$-module
$$\phi: A \rightarrow
K\{\tau\},\, a\mapsto \phi_a$$
of rank $r \geq 1$ over~$K$.
Let $\pp_0$ denote the characteristic of~$\phi$, that is, the kernel of the homomorphism $A \rightarrow K$ determined by the lowest coefficient of~$\phi$. This is a prime ideal of~$A$ and hence either $(0)$ or a maximal ideal, and $\phi$ is called of generic resp.\ of special characteristic accordingly.
By definition, the endomorphism ring of $\phi$ over $K$ is the centralizer
$$\End_K(\phi) := \{u\in K\{\tau\}\mid\forall a\in A: \phi_a\circ u = u\circ\phi_a\}.$$
This is a finitely generated projective $A$-module, and $\End_K(\phi)\otimes_AF$ is a\hyphenation{finite} finite dimensional division algebra over~$F$. In special characteristic this algebra can be non-commutative. We often identify $A$ with its image under the homomorphism $A\to\End_K(\phi)$, ${a\mapsto\phi_a}$.
It may happen that $\phi$ possesses endomorphisms over an overfield that are not defined over~$K$. But by \cite{GossFFA}, Proposition 4.7.4, Remark 4.7.5, we have:
\begin{prop}\label{def_field_end}
There exists a finite separable extension $K'$ of $K$ such that for every overfield $L$ of $K'$ we have $\End_L(\phi) = \End_{K'}(\phi)$.
\end{prop}
Consider any integrally closed infinite subring $B\subset A$. Then $A$ is a finitely generated projective $B$-module of some rank $m\ge1$, and the restriction $\phi|B$ is a Drinfeld $B$-module of rank $rm$ over~$K$. By definition there is a natural inclusion $\End_K(\phi) \subset \End_K(\phi|B)$ identifying $\End_K(\phi)$ with the commutant of $A$ in $\End_K(\phi|B)$. In special characteristic it is possible that the latter is non-commutative and $A$ is not contained in its center, in which case the inclusion is proper.
Dually consider any commutative $A$-subalgebra $A'\subset \End_K(\phi)$. Then $A'$ is a finitely generated projective $A$-module of some rank $m'\ge1$. If $A'$ is normal, i.e., integrally closed in its quotient field, the tautological embedding $\phi': A'\rightarrow K\{\tau\}$ is a Drinfeld $A'$-module of rank $r/m'$ over~$K$; in particular $m'$ is then a divisor of~$r$. One can prove the same fact for arbitrary $A'$ using the isogeny provided by the following subsection.
\subsection{Isogenies} \label{isogs}
Let $\phi'$ be a second Drinfeld $A$-module over~$K$. Let $f$ be an isogeny $\phi\to\phi'$ over~$K$, that is, a non-zero element $f\in K\{\tau\}$ satisfying $f\circ\phi_a = \phi'_a\circ f$ for all $a\in A$. Then $f$ induces an isomorphism of $F$-algebras
\begin{myequation}\label{EndIsog}
\End_K(\phi)\otimes_AF \ \stackrel{\sim}{\longrightarrow}\ \End_K(\phi')\otimes_AF
\end{myequation
which sends $e\otimes1$ to $e'\otimes1$ if $f\circ e=e'\circ f$.
The following proposition extends a result of \cite{HayEx}, Proposition 3.2, to the possibly non-commutative case and is established in a different way.
\begin{prop}\label{ConstIsog}
Let $\phi\colon A\to K\{\tau\}$ be any Drinfeld module, let $S$ be any $A$-subalgebra of $\End_K(\phi)$ and let $S'$ be any maximal $A$-order in $S\otimes_AF$ which contains~$S$. Then there exist a Drinfeld $A$-module $\phi'\colon A\to K\{\tau\}$ and an isogeny $f\colon\phi\to\phi'$ over $K$ such that $S'$ corresponds to $\End_K(\phi') \cap (S\otimes_AF)$ via the isomorphism (\ref{EndIsog}).
\end{prop}
\begin{proof}
To avoid confusing endomorphisms of $\phi$ with endomorphisms of the desired $\phi'$ we denote the tautological embedding $S\hookrightarrow K\{\tau\}$ by $s\mapsto \phi_s$.
Fix any non-zero element $a\in A$ satisfying $S'a\subset S$. Let $H_a$ denote the kernel of $\phi_a$ as a finite subgroup scheme of ${\mathbb G}_{a,K}$. Observe that the action of any endomorphism $s\in S$ on $H_a$ depends only on the residue class of $s$ modulo~$Sa$, and that $Sa$ has finite index in~$S'a$. Thus the sum
$$H := \sum_{s\in S'a} \phi_s(H_a)$$
is really finite and defines another finite subgroup scheme of ${\mathbb G}_{a,K}$. By construction $H$ is mapped to itself under $\phi_s$ for every $s\in S$. In particular it is therefore the scheme theoretic kernel of a non-zero element $f\in K\{\tau\}$.
Also, for each $s\in S$ we have $f(\phi_s(H)) \subset f(H) = 0$; hence $f\circ\phi_s$ annihilates $H=\Ker(f)$, and thus we have $f\circ\phi_s = \phi'_s\circ f$ for a unique element $\phi'_s\in K\{\tau\}$. For any two elements $s_1$, $s_2\in S$ we have
\begin{myequation}\label{ConstIsog0}
\phi'_{s_1}\circ\phi'_{s_2}\circ f
\ =\ \phi'_{s_1}\circ f\circ\phi_{s_2}
\ =\ f\circ\phi_{s_1}\circ\phi_{s_2}
\ =\ f\circ\phi_{s_1s_2}
\ =\ \phi'_{s_1s_2}\circ f
\end{myequation
and therefore $\phi'_{s_1}\circ\phi'_{s_2} = \phi'_{s_1s_2}$, and a similar calculation shows that $\phi'_{s_1}+\phi'_{s_2} = \phi'_{s_1+s_2}$. The resulting map $S\to K\{\tau\}$, $s\mapsto \phi'_s$ is thus a ring homomorphism. In particular its restriction to $A$ is a Drinfeld $A$-module $\phi'$ such that $f$ defines an isogeny $\phi\to\nobreak\phi'$, and the full map $s\mapsto \phi'_s$ defines an embedding $S\hookrightarrow \End_K(\phi')$ compatible with the isomorphism (\ref{EndIsog}). To extend this map to the maximal order~$S'$ we need the following preparation:
\begin{lem}\label{ConstIsog1}
Let $H_{a^2} \subset {\mathbb G}_{a,K}$ denote the kernel of $\phi_{a^2}$. Then
$$\sum_{s\in S'a} \phi_s(H_{a^2}) \ =\ \Ker(f\circ\phi_a).$$
\end{lem}
\begin{proof}
The summand for $s=a$ on the left hand side is $\phi_a(H_{a^2}) = H_a = \Ker(\phi_a)$ and therefore also contained in the right hand side. Thus it suffices to prove that the images of both sides under $\phi_a$ coincide. But
\begin{eqnarray*}
\phi_a\biggl( \sum_{s\in S'a} \phi_s(H_{a^2}) \biggr)
&=& \sum_{s\in S'a} \phi_a(\phi_s(H_{a^2}))
\ =\ \sum_{s\in S'a} \phi_s(\phi_a(H_{a^2}))
\ =\ \sum_{s\in S'a} \phi_s(H_a) \\
&\stackrel{\rm def}{=}& H
\ =\ \Ker(f)
\ =\ \phi_a(\Ker(f\circ\phi_a)),
\end{eqnarray*}
as desired.
\end{proof}
Now consider any $s'\in S'$, and observe that we have already constructed $\phi'_a$ and $\phi'_{s'a}$ in $K\{\tau\}$.
\begin{lem}\label{ConstIsog2}
There exists an element $\phi'_{s'}\in \End_K(\phi')$ which satisfies
$\phi'_{s'}\circ\phi'_a = \phi'_{s'a}$.
\end{lem}
\begin{proof}
For each $s\in S'a$ we have $s's$, $s'as\in S'a$; hence $\phi_{s'a}$ and $\phi_{s's}$ and $\phi_{s'as}$ all exist and satisfy $\phi_{s'a}\circ\nobreak\phi_s\allowbreak = \phi_{s'as} = \phi_{s's}\circ\phi_a$. Also, we have $f(\phi_{s's}(H_a)) \subset f(H) = 0$ and so
$$(f\circ\phi_{s'a})(\phi_s(H_{a^2}))
\ =\ (f\circ\phi_{s's})(\phi_a(H_{a^2}))
\ =\ (f\circ\phi_{s's})(H_a)
\ =\ 0.$$
Summing over all $s\in S'a$ and using Lemma \ref{ConstIsog1} we deduce that
$f\circ\phi_{s'a}$ annihilates $\Ker(f\circ\phi_a)$. Thus there exists a unique element $\phi'_{s'}\in K\{\tau\}$ satisfying
$f\circ\phi_{s'a} = \phi'_{s'}\circ f\circ\phi_a$.
The calculation
$$\phi'_{s'a}\circ f
\ =\ f\circ\phi_{s'a}
\ =\ \phi'_{s'}\circ f\circ\phi_a
\ =\ \phi'_{s'}\circ \phi'_a\circ f$$
now implies that $\phi'_{s'a} = \phi'_{s'}\circ \phi'_a$. Finally, a calculation like that in (\ref{ConstIsog0}) shows that $\phi'_{s'}\circ\phi'_b = \phi'_b\circ\phi'_{s'}$ for all $b\in A$.
Thus $\phi'_{s'}\in\End_K(\phi')$, as desired.
\end{proof}
By a calculation as in (\ref{ConstIsog0}) one easily shows that the map $S'\to \End_K(\phi')$, $s'\mapsto \phi'_{s'}$ is a ring homomorphism extending the previous one on~$S$. By construction it factors through a homomorphism $S'\to\End_K(\phi') \cap (S\otimes_AF)$ which becomes an isomorphism after tensoring with $F$ over~$A$. Since both sides of the latter are finitely generated torsion free $A$-modules, that homomorphism must be an inclusion of finite index. But as $S'$ is already a maximal $A$-order in $S\otimes_AF$, it follows that
$S'\to\End_K(\phi') \cap (S\otimes_AF)$ is an isomorphism. This finishes the proof of the proposition.
\end{proof}
\begin{prop}\label{numerics}
Let $\phi\colon A\to K\{\tau\}$ be a Drinfeld module of rank~$r$. Let $d^2$ be the dimension of $\End_K(\phi)\otimes_A F$ over its center~$Z$, and let $e$ denote the dimension of $Z$ over~$F$.
Then $de$ divides~$r$.
\end{prop}
\begin{proof}
Set $R := \End_K(\phi)$ and let $F'$ be any maximal commutative $F$-subalgebra of $R\otimes_A F$. Let $A'$ denote the integral closure of $A$ in~$F'$. Then by construction we have $\mathop{\rm rank}\nolimits_A(A') = [F'/F] = de$. Applying Proposition \ref{ConstIsog} to $S:= A'\cap R$ and $S':=A'$ yields a Drinfeld $A'$-module $\phi'\colon A'\to K\{\tau\}$ and an isogeny $f\colon\phi\to\phi'|A$. Then $\phi'|A$ has rank $r$, and the remarks at the end of Subsection \ref{endos} imply that $\phi'$ has rank $r/de$. Thus this quotient is an integer, as desired.
\end{proof}
\subsection{Tate modules} \label{tatess}
From now on we assume that $\phi$ has special characteristic. We abbreviate $R:=\End_K(\phi)$ and assume that $A$ is the center of~$R$. By Proposition \ref{numerics} we then have $\dim_F(R\otimes_A F) = d^2$ for some factorization in integers $r=nd$.
\medskip
Let $K^{\text{sep}}$ denote the separable closure of $K$ inside a fixed algebraic closure $\overline{K}$ of $K$. Let $\kappa$ denote the finite constant field of $K$ and $\overline{\kappa}$ its algebraic closure in $K^\text{sep}$. Then $G_K:=\Gal(K^{\text{sep}}/K)$ is the absolute Galois group and $G_K^{\geom}:=\Gal(K^{\text{sep}}/K \overline{\kappa})$ the geometric Galois group of~$K$. Moreover, the quotient $G_K/G_K^{\geom} \cong \Gal(\overline{\kappa}/\kappa)$ is the free pro-cyclic group topologically generated by the element $\Frob_\kappa$, which acts on $\overline{\kappa}$ by $u \mapsto u^{|\kappa|}$.
By a prime $\pp$ of $A$ we will mean any maximal ideal of $A$. The $\pp$-adic completions of $A$ and $F$ are denoted $A_\pp$ and $F_\pp$, respectively. For any prime $\pp\not=\pp_0$ of $A$ and any positive integer $i$, the $\pp^i$-torsion points of $\phi$
$$\phi[\pp^i]:=\{x \in K^{\text{sep}} \mid \forall a\in \pp^i : \phi_a(x)=0\}$$
form a free $A/\pp^i$-module of rank $r$. The $\pp$-adic Tate module $T_\pp (\phi) := \varprojlim \phi[\pp^i]$ is therefore a free $A_\pp$-module of rank $r$. Choosing a basis, the natural action of the Galois group $G_K$ on $T_\pp (\phi)$ is described by a continuous homomorphism
$$\rho_\pp : G_K\longrightarrow \GL_r(A_\pp).$$
The action of endomorphisms turns $T_\pp(\phi)$ into a module over $R_\pp :=R\otimes_A\nobreak A_\pp$.
Let $D_\pp$ denote the commutant of $R_\pp$ in $\End_{A_\pp}(T_\pp(\phi))$.
Since the action of $G_K$ commutes with that of $R_\pp$, the homomorphism $\rho_\pp$ factors through the multiplicative group $D_\pp^\times$ of $D_\pp$. We can thus view $\rho_\pp$ as a homomorphism $G_K\to D_\pp^\times$. The associated adelic Galois representation then becomes a homomorphism
$$\rho_{\text{ad}} := (\rho_\pp)_\pp:\
G_K \longrightarrow \prod_{\pp\not=\pp_0} D_\pp^\times
\ \subset\ \prod_{\pp\not=\pp_0}\GL_r(A_\pp).$$
Let $V_\pp(\phi) := T_\pp(\phi)\otimes_{A_\pp}F_\pp$ denote the rational Tate module of $\phi$ at $\pp$. Then by construction $D_\pp\otimes_{A_\pp}F_\pp$ is the commutant of $R\otimes_AF_\pp$ in $\End_{F_\pp}(V_\pp(\phi))$.
\medskip
For the next technical results we choose a maximal commutative $F$-subalgebra $F'\subset R\otimes_A F$, let $A'$ denote the integral closure of $A$ in~$F'$, and choose a Drinfeld $A'$-module $\phi'\colon A'\to K\{\tau\}$ and an isogeny $f\colon\phi\to\phi'|A$, as in the proof of Proposition \ref{numerics}. Then $\phi'$ has rank~$n$ and its characteristic $\pp'_0$ is a prime of $A'$ above~$\pp_0$. For any prime $\pp\neq\pp_0$ of $A$ the isogeny $f$ induces a $G_K$-equivariant isomorphism
\begin{myequation}\label{eq:tensisom}
V_\pp(\phi) \ \cong\ V_\pp(\phi'|A) \ \cong\ \prod_{\pp'|\pp}V_{\pp'}(\phi').
\end{myequation
\begin{prop}\label{boalo}
For any prime $\pp\not=\pp_0$ of $A$ and any prime $\pp'$ of $A'$ above $\pp$ we have:
\begin{enumerate}
\item[(a)] $D_\pp\otimes_{A_\pp}F_\pp$ is a central simple algebra of dimension $n^2$ over~$F_\pp$.
\item[(b)] There is a natural isomorphism
$D_\pp\otimes_{A_\pp}F'_{\pp'} \stackrel{\sim}{\longrightarrow}
\End_{F'_{\pp'}}(V_{\pp'}(\phi'))$.
\item[(c)] The action of $G_K$ on $V_{\pp'}(\phi')$ is induced by the homomorphism $\rho_\pp\!: {G_K\to D_\pp^\times}$ and the isomorphism (b).
\end{enumerate}
\end{prop}
\begin{proof}
By construction $R_\pp\otimes_{A_\pp}F_\pp$
is a central simple algebra of dimension $d^2$ over~$F_\pp$, and $D_\pp\otimes_{A_\pp}F_\pp$ is its commutant in the action on the $R_\pp\otimes_{A_\pp}F_\pp$-module $V_\pp(\phi)$ of dimension $r=nd$ over~$F_\pp$. With general facts on semisimple algebras this implies~(a).
Next the isomorphism (\ref{eq:tensisom}) is really the isotypic decomposition of $V_\pp(\phi)$ over $A'\otimes_AF_\pp \cong \prod_{\pp'|\pp}F'_{\pp'}$. Since the action of $A'\otimes_AF_\pp \subset R\otimes_AF_\pp$ commutes with $D_\pp$, the decomposition is $D_\pp$-invariant. Thus each $V_{\pp'}(\phi')$ is a $D_\pp$-module. The actions of both $D_\pp$ and $F'_{\pp'}$ agree on~$A_\pp$; hence they induce a non-zero homomorphism
$$D_\pp\otimes_{A_\pp}F'_{\pp'} \longrightarrow
\End_{F'_{\pp'}}(V_{\pp'}(\phi')).$$
Here by (a) the left hand side is a central simple algebra of dimension $n^2$ over~$F'_{\pp'}$. But since $V_{\pp'}(\phi')$ has dimension $n$ over $F'_{\pp'}$, the same is true for the right hand side as well. Thus the homomorphism must be an isomorphism, proving (b). Finally, the natural construction implies (c).
\end{proof}
For any prime $\pp\not=\pp_0$ of $A$, let $D^1_\pp$ denote the subgroup of all elements of $D_\pp^\times$ whose reduced norm over $F_\pp$ is~$1$.
\begin{prop} \label{Ggeom_in_SL}
There exists a finite extension $K'\subset K^{\text{sep}}$ of $K$ such that
$$\rho_{\text{ad}}(G_{K'}^{\geom}) \subset \prod_{\pp\not=\pp_0} D_\pp^1.$$
\end{prop}
\begin{proof}
Let $\phi'\colon A'\to K\{\tau\}$ be as above. By Anderson \cite{Anderson_t_Motives}, \S4.2, the determinant of the adelic Galois representation associated to $\phi'$ is the adelic Galois representation associated to some Drinfeld $A'$-module of rank $1$ of special characteristic~$\pp_0'$. By Proposition \ref{finite} below the image of $G_K^{\geom}$ in that representation is finite. Choose a finite extension $K'\subset K^{\text{sep}}$ of $K$ such that $G_{K'}^{\geom}$ lies in its kernel. Then for any prime $\pp\not=\pp_0$ of $A$ and any prime $\pp'$ of $A'$ above $\pp$, Proposition \ref{boalo} (b) and (c) implies that $\rho_\pp(G_{K'}^{\geom}) \subset D^1_\pp$, as desired.
\end{proof}
\begin{prop}\label{labacu}
For almost all primes $\pp\not=\pp_0$ of $A$, we have $D_\pp \cong {\rm Mat}_{n\times n}(A_\pp)$ and $D_\pp^\times \cong \GL_n(A_\pp)$ and $D_\pp^1 \cong \SL_n(A_\pp)$.
\end{prop}
\begin{proof}
For almost all $\pp$, the central simple algebra $R\otimes_AF_\pp$ is split and $R_\pp = R\otimes_AA_\pp$ is a maximal order therein. For these $\pp$ we have $R_\pp \cong {\rm Mat}_{d\times d}(A_\pp)$, and $T_\pp(\phi)$ is a direct sum of $n$ copies of the tautological representation $A_\pp^{d}$. Its commutant $D_\pp$ is then isomorphic to ${\rm Mat}_{n\times n}(A_\pp)$, and everything follows.
\end{proof}
\subsection{Non-singular model}\label{model}
Any integral scheme of finite type over $\mathbb{F}_p$ with function field $K$ is called a \emph{model of $K$}. By de Jong's theorem on alterations \cite[Th.$\;$4.1]{deJong1996} we have:
\begin{thm}\label{deJong}
There exists a finite separable extension $K'$ of $K$ which possesses a smooth projective model.
\end{thm}
In the following we assume that $\overline{X}$ is a smooth projective model of~$K$. Then we have:
\begin{thm}\label{Pop}
For any finite group $H$ there exist only finitely many continuous homomorphisms $G_K\to H$ which are unramified at all points of codimension $1$ of $\overline{X}$.
\end{thm}
\begin{proof}
By the Zariski-Nagata purity theorem of the branch locus \cite{ZariskiPurity}, \cite{NagataPurity}, any such extension comes from a finite \'etale covering of $\overline{X}$. In other words it factors through the \'etale fundamental group $\pi_1^{\rm et}(\,\overline{X}\,)$. This group lies in a short exact sequence
$$\xymatrix{
1 \ar[r] & \pi_1^{\rm et}(\,\overline{X}_{\overline{\kappa}}) \ar[r] &
\pi_1^{\rm et}(\,\overline{X}\,) \ar[r] &
\Gal(\overline{\kappa}/\kappa) \ar[r] & 1,\\}$$
where $\pi_1^{\rm et}(\,\overline{X}_{\overline{\kappa}}\,)$ is topologically finitely generated by Grothendieck \cite[Exp.$\,$X, Th.$\,$2.9]{SGA1}, and $\Gal(\overline{\kappa}/\kappa)$ is the free pro-cyclic group topologically generated by Frobenius. Thus $\pi_1^{\rm et}(\,\overline{X}\,)$ is topologically finitely generated and so possesses only finitely many continuous homomorphisms to $H$, as desired.
\end{proof}
We choose an open dense subscheme $X\subset\overline{X}$ such that $\phi$ extends to a family of Drinfeld $A$-modules of rank $r$ over $X$. Since $\phi$ has special characteristic $\pp_0$, the extended family has characteristic $\pp_0$ everywhere. For any $\pp\not=\pp_0$, the action of $G_K$ on $T_\pp(\phi)$ factors through the \'etale fundamental group $\pi_1^{\rm et}(X)$. In particular it is unramified at all points of codimension $1$ in~$X$.
In the next three subsections we look separately at information coming from points in $X$, respectively in $\overline{X}\smallsetminus X$.
\subsection{Frobenius action}\label{chapter_frob}
Consider any closed point $x\in X$ with finite residue field $\kappa_x$. By a Frobenius element $\Frob_x \in G_K$ we mean any element whose image in $\pi_1^{\rm et}(X)$ lies in a decomposition group above $x$ and acts by $u \mapsto u^{|\kappa_x|}$ on an algebraic closure of~$\kappa_x$. The action of $\Frob_x$ on $T_\pp(\phi)$ corresponds to the action on the Tate module $T_\pp(\phi_x)$, where $\phi_x$ denotes the reduction of $\phi$ at $x$.
Let $\pp$ be any prime of $A$ for which Proposition \ref{labacu} holds. Then $\rho_\pp(\Frob_x)\in D_\pp^\times \cong\GL_n(A_\pp)$, and we can consider its characteristic polynomial
\begin{myequation}\label{eq:charpol}
f_x(T) := \det\bigl(T\cdot{\rm Id}_n-\rho_\pp(\Frob_x)\bigr) \in A_\pp[T].
\end{myequation}%
\begin{prop}\label{F_charpol}
The polynomial $f_x$ has coefficients in $A$ and is independent of $\pp$.
\end{prop}
\begin{proof}
Let $F'$ and $\phi'\colon A'\to K\{\tau\}$ be as in Subsection \ref{tatess}. Then Proposition \ref{boalo} shows that, for every $\pp$ as above and every $\pp'|\pp$, the image of $f_x(T)$ in $F'_{\pp'}[T]$ is the characteristic polynomial of the image of $\Frob_x$ in its representation on $V_{\pp'}(\phi')$. Applying \cite{GossFFA}, Theorem 4.12.12 (b), to the Drinfeld $A'$-module $\phi'$ shows that this image has coefficients in $F'$ and is independent of $\pp'$. Fixing $\pp$ and varying $\pp'|\pp$ it follows that the coefficients of $f_x(T)$ lie in $\diag(F') \subset \prod_{\pp'|\pp}F'_{\pp'}$, in other words, in the subring $A'\otimes_AF \subset A'\otimes_AF_\pp$. But by definition they also lie in the subring $A_\pp \cong A\otimes_AA_\pp$, whose intersection with the former is just $A$. Varying both $\pp$ and $\pp'$ then shows that $f_x(T)$ is independent of $\pp$.
\end{proof}
\begin{prop}\label{F_eigenvalues}
Let $\alpha_1, \ldots, \alpha_n$ be the roots of $f_x$ in an algebraic closure $\overline{F}$ of $F$, with repetitions if necessary. Consider any normalized valuation $v$ of $F$ and an extension $\overline{v}$ of $v$ to $\overline{F}$. Let $k_v$ denote the residue field at $v$.
\begin{enumerate}
\item[(a)] If $v$ does not correspond to $\pp_0$ or $\infty$, then for all $1 \leq i \leq n$ we have
$$\overline{v}(\alpha_i)=0.$$
\item[(b)] If $v$ corresponds to $\infty$, then for all $1 \leq i \leq n$ we have
$$\overline{v}(\alpha_i)=-\frac{1}{nd}\cdot \frac{[\kappa_x/\mathbb{F}_p]}{[k_v/\mathbb{F}_p]}.$$
\item[(c)] If $v$ corresponds to $\pp_0$, then there exists an integer $1\leq n_x \leq n$ such that
$$\overline{v}(\alpha_i)=\left\{ \begin{array}{ll}
\frac{1}{n_x d}\cdot \frac{[\kappa_x/\mathbb{F}_p]}{[k_v/\mathbb{F}_p]} & \text{ for precisely } n_x \text{ of the } \alpha_i, \text{ and} \\[7pt]
0 & \text{ for the remaining } n-n_x \text{ of the } \alpha_i.
\end{array}\right.$$
\end{enumerate}
\end{prop}
\begin{proof}
By construction the $\alpha_i$ are the roots of the characteristic polynomial of $\rho_\pp(\Frob_x)$ associated to the Drinfeld module $\phi$ of rank $r=nd$, except that their multiplicities are divided by $d$. Thus the proposition is a direct consequence of \cite{DriII}, Proposition 2.1, to~$\phi$.
\end{proof}
\subsection{Good reduction and lattices} \label{lettuce}
In this subsection we briefly leave the current setting and consider the following general situation.
Let $L$ be a field containing $\mathbb{F}_p$ with a non-trivial discrete valuation $v$. Let $R\subset L$ denote the associated discrete valuation ring and ${\mathfrak m}$ its maximal ideal. Let $\psi:A\to R\{\tau\}$, $a\mapsto\psi_a$ be a Drinfeld $A$-module of rank $s>0$ with good reduction, i.e., such that for every $a\in A\smallsetminus\{0\}$ the highest non-zero coefficient of $\psi_a$ is a unit in $R$.
We view $L$ as an $A$-module with respect to the action $a\cdot u :=\psi_a(u)$ for all $a\in A$ and $u\in L$. Then $R$ is a submodule for this action, and we are interested in the structure of the $A$-module $L/R$.
To any $A$-module $M$ are associated the following notions. The \emph{rank of $M$} is the maximal number of $A$-linearly independent elements of~$M$, or $\infty$ if the maximum does not exist. Of course, any finitely generated $A$-module has finite rank. Next, the \emph{division hull} of an $A$-submodule $N\subset M$ is defined as
\begin{myequation}\label{DivHull}
\sqrt{N}\ :=\ \bigl\{ \overline{u}\in M \bigm|
\exists\, a\in A\smallsetminus\{0\}: a\cdot\overline{u}\in N\bigr\},
\end{myequation}%
which is an $A$-module of the same rank as $N$. The $A$-module $M$ is called \emph{tame} if every finitely generated $A$-submodule $N\subset M$ satisfies $[\sqrt{N}:N] < \infty$.
The following result was obtained by Poonen in \cite[Lemma$\,$5]{PoonenMordellWeil} when $L$ is a global field and $\psi$ has generic characteristic, and by Wang \cite{WangMordellWeil} in general.
\begin{prop}\label{lettprop1}
$L/R$ is a tame $A$-module.
\end{prop}
\subsection{Bad reduction} \label{bad_red}
Now we return to the situation and the notations of Subsections \ref{endos} through \ref{chapter_frob}. We assume in addition that there exists a prime $\pp\not=\pp_0$ of $A$ such that all $\pp$-torsion points of $\phi$ are defined over~$K$. This can be achieved on replacing $K$ by the finite separable extension corresponding to the action of $G_K$ on $\phi[\pp]$.
Let $x$ be one of the finitely many generic points of $\overline{X}\smallsetminus X$. Let $K_x$ denote the completion of $K$ with respect to the valuation at $x$, and let $R_x\subset K_x$ denote the associated discrete valuation ring. Since $\phi$ possesses a full level structure of some level $\pp\not=\pp_0$ over~$K$, it is known to have semistable reduction over $K_x$. Its Tate uniformization at $x$ (see \cite{DriI}, \S 7) then consists of a Drinfeld $A$-module $\psi_x$ over $R_x$ of some rank $1\leq r_x\leq r$ with good reduction and an $A$-lattice $\Lambda_x \subset K_x^\text{sep}$ of rank $r-r_x$ for the action of $A$ on $K_x^\text{sep}$ via $\psi_x$. Here by definition an $A$-lattice is a finitely generated projective $A$-submodule whose intersection with any ball of finite radius is finite. This implies that any non-zero element of $\Lambda_x$ has valuation $<0$. Also, being finitely generated, the lattice is already contained in some finite Galois extension $K_x'$ of $K_x$.
Let $I_x\subset D_x\subset G_K$ denote the inertia group, respectively the decomposition group, at a fixed place of $K^{\text{sep}}$ above $x$. Then $D_x$ is also the absolute Galois group of $K_x$. Let $D'_x\triangleleft D_x$ denote the absolute Galois group of $K'_x$, and set $I'_x := I_x\cap D'_x$. Then $D_x$ acts on $\Lambda_x$ through the finite quotient $D_x/D'_x$.
For any prime $\pp\not=\pp_0$ of $A$ and any positive integer $i$ the Tate uniformization yields a $D_x$-equivariant isomorphism
\begin{myequation}\label{TateSeq0}
\phi[\pp^i]\ \cong\
\bigl\{ u \in K_x^{\text{sep}} \bigm|
\forall a\in \pp^i : \psi_{x,a}(u)\in\Lambda_x \bigr\} \bigm/ \Lambda_x
\end{myequation}%
and hence a $D_x$-equivariant short exact sequence
$$0 \longrightarrow \psi_x[\pp^i] \longrightarrow \phi[\pp^i] \longrightarrow \Lambda_x\otimes_A(\pp^{-i}/A)\longrightarrow 0.$$
Taking the inverse limit over $i$ yields a $D_x$-equivariant short exact sequence
$$0 \longrightarrow T_\pp(\psi_x) \longrightarrow T_\pp(\phi) \longrightarrow \Lambda_x\otimes_AA_\pp \longrightarrow 0.$$
Here $I_x$ acts trivially on $T_\pp(\psi_x)$, and $D'_x$ acts trivially on $\Lambda_x\otimes_AA_\pp$. Thus in a suitable basis $\rho_\pp(D'_x)$ is contained in the group of block triangular matrices of the form
$$\left(\begin{array}{c|c} \ast & \ast \\ \hline 0 & 1 \\ \end{array}\right)
\ \subset\ \GL_r(A_\pp),$$
and $\rho_\pp(I'_x)$ is a $\rho_\pp(D_x)$-invariant subgroup of the group of block triangular matrices of the form
\begin{myequation}\label{inertunip}
\left(\begin{array}{c|c} 1 & \ast \\ \hline 0 & 1 \\ \end{array}\right)
\ \cong\ \Hom_A(\Lambda_x,T_\pp(\psi_x))
\ \cong\ T_\pp(\psi_x)^{r-r_x}.
\end{myequation}%
We are interested in the following three consequences:
\begin{lem}\label{no_inert}
Fix an integer $c\geq1$. Then for almost all primes $\pp\not=\pp_0$ of $A$, any continuous homomorphism from $\rho_\pp(D_x)$ to a finite group of order $\leq c$ is trivial on $\rho_\pp(I'_x)$.
\end{lem}
\begin{proof}
Fix a Drinfeld $A$-module $\psi_y$ of rank $r_x$ over a finite field that arises by good reduction from $\psi_x$. Let $\Frob_y$ be an associated Frobenius element in $D_x/I_x$, the absolute Galois group of the residue field at $x$. Then by \cite{GossFFA}, Theorem 4.12.12 (b), the characteristic polynomial of $\Frob_y$ on $T_\pp(\psi_x)$ has coefficients in $A$ and is independent of $\pp$. Moreover, \cite{DriII}, Proposition 2.1, implies that none of its eigenvalues $\beta_1,\ldots,\beta_{r_x}\in \overline{F}$ is a root of unity. Thus $a := \prod_{i=1}^{r_x}(\beta_i^{c!}-1)$ is a non-zero element of $A$. We claim that the assertion holds for all $\pp\not=\pp_0$ that do not divide $a$.
Indeed, let $f: \rho_\pp(D_x)\to H$ be a continuous homomorphism to a finite group of order $\leq c$, such that $f|\rho_\pp(I'_x)$ is non-trivial. Then $\mathop{\rm ker} f|\rho_\pp(I'_x)$ is a $\rho_\pp(D_x)$-invariant proper closed subgroup of $\rho_\pp(I'_x)$ of index $\leq c$. Thus $T_\pp(\psi_x)^{r-r_x}$ and hence $T_\pp(\psi_x)$, as a representation of $\rho_\pp(D_x)$, possesses a non-trivial finite subquotient of order $\leq c$. Then $\Frob_y^{c!}$ acts trivially on this subquotient. But this requires that some $\beta_i^{c!}$ is congruent to $1$ modulo a prime of $\overline{F}$ above $\pp$, or equivalently that $\pp|a$. This proves the claim.
\end{proof}
\begin{lem}\label{inert_big}
For almost all primes $\pp\not=\pp_0$ of $A$ we have $\psi_x[\pp] = \phi[\pp]^{I_x} = \phi[\pp]^{I'_x}$.
\end{lem}
\begin{proof}
The inclusions $\psi_x[\pp] \subset \phi[\pp]^{I_x} \subset \phi[\pp]^{I'_x}$ result from the fact that $I_x$ acts trivially on $\psi_x[\pp]$. To prove equality take any element of
$\phi[\pp]^{I'_x}$. By (\ref{TateSeq0}) it corresponds to the residue class $u+\Lambda_x$ for some $u \in K_x^{\text{sep}}$ satisfying $\psi_{x,a}(u)\in\Lambda_x$ for all $a\in \pp$. That this residue class is $I'_x$-invariant means that $\sigma u-u \in \Lambda_x$ for all $\sigma\in I'_x$. But $\sigma\in I'_x$ acts trivially on $\psi_{x,a}(u)\in\Lambda_x$ for all $a\in\pp$; hence $\psi_{x,a}(\sigma u-u) = \sigma\psi_{x,a}(u) - \psi_{x,a}(u) = 0$. Since the homomorphism $\psi_{x,a}: \Lambda_x\to\Lambda_x$ is injective whenever $a\not=0$, it follows that $\sigma u-u=0$ and hence $u$ is $I'_x$-invariant.
Let $L$ denote the maximal unramified extension of $K'_x$, and $R\subset L$ its discrete valuation ring. As in Subsection \ref{lettuce} we denote the residue class in $L/R$ of an element $v\in L$ by $\overline{v}$ and abbreviate $a\cdot\overline{v} := \overline{\psi_{x,a}(v)}$ for all $a\in A$. Since every non-zero element of $\Lambda_x$ has valuation $< 0$, we have $\Lambda_x \cap R =\{0\}$ and thus the natural map $\Lambda_x\to L/R$ is injective; let $N_x$ denote its image.
The fact that $u$ is $I'_x$-invariant means that $u\in L$. On the other hand, the fact that $\psi_{x,a}(u)\in\Lambda_x$ for all $a\in \pp$ implies that $\pp\cdot\overline{u} \subset N_x$. In particular we have $\overline{u}\in\sqrt{N_x}$ in the notation of (\ref{DivHull}). But since $[\sqrt{N_x}:N_x] < \infty$ by Proposition \ref{lettprop1}, for almost all $\pp$ we can deduce that $\overline{u}\in N_x$. Then $u=v+\lambda$ for some $v\in R$ and $\lambda\in\Lambda_x$. For all $a\in \pp$ we then have $\psi_{x,a}(v)\in\Lambda_x\cap R=\{0\}$; in other words $v\in\psi_x[\pp]$. Thus the residue class in question $u+\Lambda_x$ comes from an element of $\psi_x[\pp]$, as desired.
\end{proof}
\begin{lem}\label{small_quotients1}
For any finite abelian group $H$ there exists a finite set $P'$ of primes of $A$, such that the number of continuous homomorphisms $G_K\to H$, which are trivial on $\mathop{\rm ker}(\rho_\pp)$ for some $\pp\not\in P'$, is finite.
\end{lem}
\begin{proof}
For each of the finitely many generic points $x$ of $\overline{X}\smallsetminus X$, let $P_x$ denote the finite set of primes of $A$ excluded by Lemma \ref{no_inert} with $c:=|H|$. We claim that the assertion holds with $P'$ the union of these sets $P_x$.
Indeed, let $f: G_K \to H$ be a continuous homomorphism which is trivial on $\mathop{\rm ker}(\rho_\pp)$ for some $\pp\not\in P'$. From Subsection \ref{model} we know that $\rho_\pp$ and hence $f$ factors through the \'etale fundamental group $\pi_1^{\rm et}(X)$. Also, the restriction $f|I'_x$ is trivial for every generic point $x$ of $\overline{X}\smallsetminus X$ by Lemma \ref{no_inert}. There are therefore only finitely many possibilities for the restriction $f|I_x$. Since there are only finitely many $x$, it suffices to prove that the number of such $f$ with fixed restrictions $f|I_x$ for all $x$ is finite.
But since $H$ is abelian, any two such homomorphisms $f$ differ by a continuous homomorphism $g: G_K \to H$ which is unramified over $X$ and at all generic points of
$\overline{X}\smallsetminus X$. By Theorem \ref{Pop} there are only finitely many possibilities for such $g$, and the desired finiteness follows.
\end{proof}
\subsection{Setup}\label{setup}
From here on we assume that $\phi$ satisfies the conditions of Theorem \ref{main_theorem}. Since we are only interested in the image of Galois groups up to commensurability, we may replace $K$ by a finite extension. We first replace it by the composite of the extensions provided by Propositions \ref{def_field_end} and \ref{Ggeom_in_SL} and the fields of definition of all $\pp$-torsion points of $\phi$ for some chosen prime $\pp\not=\pp_0$ of $A$. Thereafter we replace it by the extension from Theorem \ref{deJong}.
By Proposition \ref{finite} below the assumption on $\End_{K^{\text{sep}}}(\phi|B)=R$ in Theorem \ref{main_theorem} implies that $n\geq2$.
Thus altogether we have the following assumptions:
\begin{asses}\label{asses}
\begin{enumerate}
\item[(a)] $R := \End_K(\phi) = \End_{K^{\text{sep}}}(\phi)$.
\item[(b)] The center of $R$ is $A$.
\item[(c)] $n:=r/d \geq2$.
\item[(d)] For every integrally closed infinite subring $B \subset A$ we have $\End_{K^{\text{sep}}}(\phi|B)=\nobreak R$.
\item[(e)] $\rho_{\text{ad}}(G_K^{\geom}) \subset \prod_{\pp\not=\pp_0} D_\pp^1$.
\item[(f)] There exists a prime $\pp\not=\pp_0$ of $A$ such that all $\pp$-torsion points of $\phi$ are defined over~$K$.
\item[(g)] $K$ possesses a smooth projective model $\overline{X}$.
\end{enumerate}
\end{asses}
\subsection{Images of Galois groups}\label{known}
Throughout the following we let $P_0$ denote the finite set of primes excluded by Proposition \ref{labacu}. For any $\pp\not\in P_0$ we set
$$\begin{array}{lll}
\Gamma_\pp & :=\ \rho_\pp(G_K)
& \!\!\!\subset\ \GL_n(A_\pp), \quad\hbox{and} \\[5pt]
\Gamma_\pp^{\geom}\! & :=\ \rho_\pp(G_K^{\geom})
& \!\!\!\subset\ \SL_n(A_\pp).
\end{array}$$
By construction the latter is a closed normal subgroup of the former and the quotient is pro-cyclic. Combining Proposition \ref{boalo} (c) with \cite{PinI}, Theorem 1.1, and applying \cite{PinIII}, Lemma 3.7, we obtain:
\begin{thm}\label{Zariski_dense}
For any $\pp$ as above $\Gamma_\pp$ is Zariski dense in $\GL_{n,F_\pp}$, and $\Gamma_\pp^{\geom}$ is Zariski dense in $\SL_{n,F_\pp}$.
\end{thm}
The next result concerns the image of the group ring. By \cite{PinIII}, Theorem B, in the case that $K$ has transcendence degree $1$, and by \cite{PR1}, Theorem 0.2, in the general case, we know:
\begin{thm}\label{split_at_p}
For almost all primes $\pp$ of $A$ we have $A_\pp[\Gamma_\pp] ={\rm Mat}_{n\times n}(A_\pp)$.
\end{thm}
Let $k_\pp := A/\pp$ denote the residue field of $\pp$, and let $\overline{\rho}_\pp: G_K\to \GL_n(k_\pp)$ denote the reduction of $\rho_\pp$ modulo $\pp$.
Theorem \ref{split_at_p} immediately implies:
\begin{cor}\label{abs_irred}
For almost all primes $\pp$ of $A$ the representation $\overline{\rho}_\pp$ on $k_\pp^n$ is absolutely irreducible.
\end{cor}
\begin{thm} \label{openness}
For any finite set $P$ of primes $\not=\pp_0$ of $A$, consider the combined representation
$$\rho_{P} := (\rho_\pp)_\pp:\
G_K \longrightarrow \prod_{\pp\in P} D_\pp^\times
\ \subset\ \prod_{\pp\in P} \GL_r(A_\pp).$$
Then $\rho_{P}(G_K^{\geom})$ has finite index in $\prod_{\pp\in P} D^1_\pp$.
\end{thm}
\begin{proof}
Since $n\ge2$, Proposition \ref{finite} below shows that $\phi$ is not isomorphic over $K^{\text{sep}}$ to a Drinfeld module defined over a finite field. We may thus apply \cite{PinII}, Theorems 6.1 and 6.2. The subfield $E$ given there is contained in the center $F$ of $R\otimes_AF$, such that $B := E\cap A$ is infinite and $\End_{K^{\text{sep}}}(\phi|B)\otimes_BE$ has center $E$. But by Assumption \ref{asses} (d) we have $\End_{K^{\text{sep}}}(\phi|B)=R$ with center $A$. Thus we must have $E=F$.
The group $G_Q(E_Q)$ described in \cite{PinII}, Theorem 6.2, is then the centralizer of $R\otimes_A\prod_{\pp\in P}F_\pp$ in $\prod_{\pp\in P}\Aut_{F_\pp}(V_\pp(\phi))$. In our situation it is therefore equal to $\prod_{\pp\in P}(D_\pp\otimes_{A_\pp}F_\pp)^\times$.
The subgroup $G_Q^{\rm der}(E_Q)$ is the subgroup of elements of reduced norm $1$.
Theorem 6.1 of \cite{PinII} says that $\rho_{P}(G_K^{\geom})$ is commensurable to an open subgroup of $G_Q^{\rm der}(E_Q)$. Since $\rho_{P}(G_K^{\geom})$ is already contained in $\prod_{\pp\in P} D^1_\pp$ by Assumption \ref{asses} (e), which is an open compact subgroup of $G_Q^{\rm der}(E_Q)$, the index must be finite, as desired.
\end{proof}
\subsection{Ring of traces} \label{ring_of_trances}
Let $\Ad$ denote the adjoint representation of $\GL_n$. Proposition \ref{F_charpol} implies that the trace $\Tr\Ad(\rho_\pp(\Frob_x))$ lies in $F$ and is independent of $\pp$. We let $R^{\rm trad}$ denote the subring of $F$ generated by $\Tr\Ad(\rho_\pp(\Frob_x))$ for all closed points $x\in X$, and let $F^{\rm trad} \subset F$ denote the quotient field of $R^{\rm trad}$.
\begin{thm} \label{field_trad}
Either $F^{\rm trad} = F$, or $n=p=2$ and $F^{\rm trad}=F^2 := \{x^2\mid x\in F\}$.
\end{thm}
\begin{proof}
Applying \cite{PinII}, Theorem 1.2, to the Drinfeld $A'$-module $\phi'$ from Subsection \ref{tatess} yields a subfield $E\subset F'$, which by Assumption \ref{asses} (d) turns out to be $F$. (One may equivalently combine \cite{PinII}, Theorem 1.1, for $\phi'$ with Theorem \ref{openness} above.) Thus by \cite{PinII}, Theorem 1.3, the subfield generated by the traces of Frobeniuses in the adjoint representation associated to $\phi'$ has the desired properties. But by Proposition \ref{boalo}, those traces are just $\Tr\Ad(\rho_\pp(\Frob_x))$; hence this subfield is $F^{\rm trad}$.
\end{proof}
As the following proposition shows, the second case in Theorem \ref{field_trad} really does occur:
\begin{prop}\label{Rtradsmall}
Let $\kappa'\subset\overline{\kappa}$ denote the extension of degree $2$ of the constant field $\kappa$. If $n=p=2$, then after replacing $K$ by $K\kappa'$, we have $F^{\rm trad}=F^2$.
\end{prop}
\begin{proof}
In characteristic $p=2$, let ${\rm \std}^{(2)}$ denote the pullback under $\Frob_2$ of the standard representation of $\GL_2$, and let $\det:\GL_2\to{\mathbb G}_m$ denote the determinant. Then the adjoint representation of $\GL_2$ is an extension of ${\rm \std}^{(2)} \otimes \det^{-1}$ with two copies of the trivial representation of dimension~$1$. Thus for every $g\in \GL_2$ we have $\Tr\Ad(g) = \Tr(g)^2\cdot\det(g)^{-1} + 2$.
Recall from Assumption \ref{asses} (e) that $\rho_\pp(G_K^{\geom}) \subset \SL_n(A_\pp)$. Thus $\det\circ\rho_\pp$ factors through a homomorphism $\Gal(\overline{\kappa}/\kappa) \to A_\pp^\times$. Its value on any element of $\Gal(\overline{\kappa}/\kappa')$ is therefore a square. After replacing $K$ by $K\kappa'$ we find that $\Tr\Ad(\rho_\pp(\Frob_x)) \in F\cap F_\pp^2 = F^2$ for every closed point $x\in X$. Thus now only the second case in Theorem \ref{field_trad} is possible.
\end{proof}
\begin{prop}\label{Rtradprop}
Let $A_0$ be the ring of elements of $F$ which are regular outside $\pp_0$. Then either $R^{\rm trad}$ is a subring of finite index of $A_0$, or $n=p=2$ and $R^{\rm trad}$ is a subring of finite index in $A_0^2 := \{x^2\mid x\in A_0\}$.
\end{prop}
\begin{proof}
Let $\alpha_1, \ldots, \alpha_n \in \overline{F}$ denote the eigenvalues of $\rho_\pp(\Frob_x)$. By Proposition \ref{F_eigenvalues} they have valuation $0$ at all places not above $\pp_0$ or $\infty$, and the same negative valuation at any place above $\infty$. Thus their ratios $\alpha_i/\alpha_j$ have trivial valuation at all places not above $\pp_0$. The sum over all $i,j$ of these ratios is therefore regular at all places $\not=\pp_0$. This sum is just $\Tr\Ad(\rho_\pp(\Frob_x))$, proving that $R^{\rm trad} \subset A_0$.
By Theorem \ref{field_trad} the ring $R^{\rm trad}$ must contain some non-constant element $x$. Then $F$ is a finite field extension of ${\mathbb F}_p(x)$. Moreover, $x$ as an element of $F$ is regular outside $\pp_0$, and therefore $\pp_0$ is the unique place of $F$ above the place of ${\mathbb F}_p(x)$ where $x$ has a pole. This implies that $A_0$ is the integral closure of ${\mathbb F}_p[x]$ in~$F$. It is therefore a module of finite type over ${\mathbb F}_p[x]$, and so $R^{\rm trad}$ is a submodule that is again of finite type. In particular, $R^{\rm trad}$ is already generated by finitely many traces.
Also, it follows that $R^{\rm trad}$ is of finite index in its normalization. Depending on the case in Theorem \ref{field_trad}, this normalization is either $A_0$ or $A_0^2$, and we are done.
\end{proof}
By construction any prime $\pp\not=\pp_0$ of $A$ corresponds to a unique prime of $A_0$. Thus there are natural homomorphisms $R^{\rm trad} \hookrightarrow A_0 \hookrightarrow A_\pp \twoheadrightarrow k_\pp$.
\begin{prop}\label{residue_field_trad1}
There exists a finite set $P^{\rm trad}$ of primes of $A$, containing $\pp_0$, such that:
\begin{enumerate}
\item[(a)] For any prime $\pp \not\in P^{\rm trad}$ of $A$, the homomorphism $R^{\rm trad} \to k_\pp$ is surjective.
\item[(b)] For any two disctinct primes $\pp_1,\pp_2 \not\in P^{\rm trad}$ of $A$, the homomorphism $R^{\rm trad} \to k_{\pp_1}\times k_{\pp_2}$ is surjective.
\item[(c)] For any prime $\pp \not\in P^{\rm trad}$ of $A$, the image of the homomorphism $R^{\rm trad} \to A_\pp$ is dense in $A_\pp$ if $F^{\rm trad} = F$, respectively dense in $A_\pp^2 := \{a^2\mid a\in A_\pp\}$ if $F^{\rm trad}=F^2$.
\end{enumerate}
\end{prop}
\begin{proof}
Depending on the case, Proposition \ref{Rtradprop} implies that the annihilator of $A_0/R^{\rm trad}$, respectively the annihilator of $A_0^2/R^{\rm trad}$, as an $R^{\rm trad}$-module contains a non-zero element $x\in R^{\rm trad}$. Then $R^{\rm trad}[x^{-1}]$ is equal to $A_0[x^{-1}]$, respectively to $A_0^2[x^{-1}]$. Let $P^{\rm trad}$ be the finite set of primes of $A$ consisting of $\pp_0$ and all those dividing $x$ within $A_0$. Then $P^{\rm trad}$ has all the desired properties.
\end{proof}
\section{Surjectivity of the residual representation}\label{sect_surjective}
\section{Proof of the main result}\label{sect_surjective}
In this section we prove Theorem \ref{main_theorem}. Subsections \ref{51} through \ref{54} deal with the image of the geometric Galois group $G_K^{\geom}$, while Subsection \ref{55} finishes with the image of the absolute Galois group~$G_K$.
We keep all the notations from the preceding section and impose Assumptions \ref{asses}.
\subsection{Residual surjectivity at a single prime}\label{51}
Recall that $P_0$ denotes the finite set of primes excluded by Proposition \ref{labacu}. For any prime $\pp\not\in P_0$ of $A$, we let $\Delta_\pp^{\text{geom}} \triangleleft \Delta_\mathfrak{p}\subset\GL_n(k_\pp)$ denote the images of $G_K^{\text{geom}} \triangleleft G_K$ under the residual representation $\overline{\rho}_\pp$. Thus
$$\begin{array}{lll}
\Gamma_\pp
& \twoheadrightarrow\ \Delta_\pp
& \!\!\!\subset\ \GL_n(k_\pp),
\\[5pt]
\Gamma_\pp^{\text{geom}}\!\!\!
& \twoheadrightarrow\ \Delta_\pp^{\text{geom}}
& \!\!\!\subset\ \SL_n(k_\pp),
\end{array}$$
and the quotient $\Delta_\pp/\Delta_\pp^{\text{geom}}$ is cyclic. We will prove that $\Delta_\pp^{{\geom}} = \SL_n(k_\pp)$ for almost all $\pp$.
\begin{lem}\label{fgh_on_delta1}
Fix any integer $c\geq 1$, and let $f$ denote the morphism from (\ref{Rdef}). Then for almost all primes $\pp\not\in P_0$ of $A$, the map $\Delta_\pp \to k_\pp$, $\delta \mapsto f(\delta^c)$ is not identically zero.
\end{lem}
\begin{proof}
Take any prime $\pp\not\in P_0$ of $A$. Then by Theorem \ref{Zariski_dense} together with Lemma \ref{R_nontrivial}, the map $\Gamma_\pp \to F_\pp$, $\gamma\mapsto f(\gamma^c)$ is not identically zero. Since this map is continuous and the images of Frobenius elements are dense in $\Gamma_\pp$, we may fix a closed point $x\in X$ such that $a := f(\rho_\pp(\Frob_x)^c)\not=0$. By the definition of $f$, this value is a polynomial with coefficients in ${\mathbb Z}$ in the coefficients of the characteristic polynomial of $\rho_\pp(\Frob_x)^c$. With Proposition \ref{F_charpol} it follows that $a$ lies in $A$ and is independent of $\pp$. In other words, having found $x$ and $a\in A\smallsetminus\{0\}$ with the help of \emph{some} auxiliary prime $\pp\not\in P_0$, we then have $f(\rho_\pp(\Frob_x)^c) = a$ for \emph{every} prime $\pp\not\in P_0$.
Thus for $\delta := \overline{\rho}_\pp(\Frob_x) \in \Delta_\pp$ we now deduce that $f(\delta^c) = a\mathop{\rm mod}\pp$. This is non-zero whenever $\pp\nmid a$; hence the desired assertion holds whenever $\pp\not\in P_0$ and $\pp\nmid a$.
\end{proof}
Let $\bar k_\pp$ denote an algebraic closure of $k_\pp$, and set $W_\pp := \phi[\pp]\otimes_{k_\pp}\bar k_\pp$. By Corollary \ref{abs_irred} this is an irreducible representation of $\Delta_\pp$ over $\bar k_\pp$ for all $\pp$ outside some finite set of primes $P^{\rm irr}$. By Theorem \ref{L-P3} there then exists a normal subgroup $\Delta'_\pp \triangleleft\Delta_\pp$ of index $\leq c'_n$, such that $\Delta'_\pp/Z(\Delta'_\pp)$ is a direct product of finite simple groups of Lie type in characteristic~$p$. We fix such a subgroup $\Delta'_\pp$ for every $\pp\not\in P^{\rm irr}$.
\begin{lem}\label{Wdecomp}
For almost all primes $\pp\not\in P^{\rm irr}$ of $A$, we have $W_\pp = W_{\pp,1}\oplus\ldots\oplus W_{\pp,m_\pp}$ for pairwise inequivalent irreducible representations $W_{\pp,i}$ of $\Delta'_\pp$.
\end{lem}
\begin{proof}
Let $W_{\pp,1}$ be any irreducible representation of $\Delta'_\pp$ contained in $W_\pp$. Then the sum of the conjugates $\delta W_{\pp,1}$ for all $\delta\in\Delta_\pp$ is a non-zero $\Delta_\pp$-invariant subspace. By irreducibility it is therefore equal to $W_\pp$ for all $\pp\not\in P^{\rm irr}$. Thus $W_\pp$ is the direct sum of certain conjugates $\delta W_{\pp,1}$.
It remains to show that these summands are pairwise inequivalent. For this suppose that $\delta_1W_{\pp,1}$ and $\delta_2W_{\pp,1}$ are distinct but equivalent as representations of $\Delta'_\pp$ for some $\delta_1,\delta_2\in\Delta_\pp$. Then for every $\delta\in\Delta_\pp$, we have $\delta^{c'_n!} \in \Delta'_\pp$, and this element has the same eigenvalues on $\delta_1W_{\pp,1}$ and $\delta_2W_{\pp,1}$. By Lemma \ref{Rprop} (a) we thus have $f(\delta^{c'_n!})=0$. But since $\delta\in\Delta_\pp$ is arbitrary, by Lemma \ref{fgh_on_delta1} with $c=c'_n!$ this can happen only for finitely many primes~$\pp$, as desired.
\end{proof}
The stated properties imply that the decomposition in Lemma \ref{Wdecomp} is the isotypic decomposition of $W_\pp$ under $\Delta'_\pp$. It is therefore normalized by $\Delta_\pp$, and so the permutation action is given by a homomorphism from $\Delta_\pp$ to the symmetric group $S_{m_\pp}$ on $m_\pp$ letters. Let $\sigma_\pp$ denote the composite homomorphism $G_K\twoheadrightarrow \Delta_\pp \to S_{m_\pp}$.
\begin{lem}\label{PermRepUnram}
For almost all primes $\pp\not\in P^{\rm irr}$ of $A$, the homomorphism $\sigma_\pp$ is unramified at all points of codimension $1$ of $\overline{X}$.
\end{lem}
\begin{proof}
This is clear for points in $X$, because $\overline{\rho}_\pp$ is already unramified there. So let $x$ be one of the finitely many generic points of $\overline{X}\smallsetminus X$. Since $|S_{m_\pp}| \leq m_\pp! \leq n!$ is bounded, Lemma \ref{no_inert} implies that $\sigma_\pp|I'_x$ is trivial for almost all $\pp$. Then $I'_x$ stabilizes each summand $W_{\pp,i}$. Since $I'_x$ acts unipotently by (\ref{inertunip}), we deduce that $W_{\pp,i}^{I'_x}\not=0$ for every $i$. On the other hand Lemma \ref{inert_big} implies that $W_\pp^{I_x} = W_\pp^{I'_x}$ for almost all~$\pp$. This means that $I_x$ acts trivially on $W_\pp^{I'_x} = W_{\pp,1}^{I'_x}\oplus\ldots\oplus W_{\pp,m_\pp}^{I'_x}$. But as all these summands are non-zero, and $I_x$ permutes them according to the restriction of the homomorphism $\sigma_\pp$, it follows that $\sigma_\pp|I_x$ is trivial, as desired.
\end{proof}
\begin{lem}\label{DeltaPrimeIrred}
For almost all primes $\pp\not\in P^{\rm irr}$ of $A$, the group $\Delta'_\pp$ acts irreducibly on~$W_\pp$.
\end{lem}
\begin{proof}
Combining Lemma \ref{PermRepUnram}, the inequality $m_\pp\leq n$, and Theorem \ref{Pop}, we find that there are only finitely many possibilities for the homomorphism $\sigma_\pp$. The intersection of their kernels is therefore equal to $G_{K'}$ for some subextension $K'\subset K^{\text{sep}}$ that is finite over $K$. Applying Corollary \ref{abs_irred} with $K'$ in place of $K$ implies that $\overline{\rho}_\pp(G_{K'})$ acts irreducibly on $W_\pp$ for almost all $\pp$. But by construction $\overline{\rho}_\pp(G_{K'})$ stabilizes each summand of the decomposition in Lemma \ref{Wdecomp}; hence $m_\pp=1$ and $W_\pp = W_{\pp,1}$ for almost all $\pp$. Then $\Delta'_\pp$ acts irreducibly on $W_\pp$, as desired.
\end{proof}
\begin{lem}\label{almost_surjective2}
For almost all primes $\pp\not\in P^{\rm irr}$ of $A$, there exist a finite subfield $k_\pp'$ of $\bar k_\pp$ and a model $G'_\pp$ of $\SL_{n,\bar k_\pp}$ over $k'_\pp$, such that $\Delta_\pp^{\prime\rm der} = G'_\pp(k'_\pp)$.
\end{lem}
\begin{proof}
By Lemma \ref{DeltaPrimeIrred} the group $\Delta'_\pp$ acts irreducibly on $W_\pp$ for almost all~$\pp$. On the other hand let $c$ be the constant from Theorem \ref{finite_main1}. Then for almost all $\pp$, Lemma \ref{fgh_on_delta1} shows that the map $\Delta_\pp \to k_\pp$, $\delta \mapsto f(\delta^{c'_nc})$ is not identically zero. Since $\delta^{c'_n}\in\Delta'_\pp$ for all $\delta\in\Delta_\pp$, it follows that the map $\Delta'_\pp \to k_\pp$, $\delta' \mapsto f(\delta^{\prime c})$ is not identically zero. Together we find that $\Delta'_\pp$ satisfies the assumptions of Theorem \ref{finite_main1}, and so the desired assertion follows.
\end{proof}
\begin{lem}\label{kp_is_kp1}
For almost all primes $\mathfrak{p}$ of $A$ as in Lemma \ref{almost_surjective2} we have $k_\pp \subset k_\pp'$.
\end{lem}
\begin{proof}
Let $P'$ be the finite set of primes excluded by Lemma \ref{almost_surjective2}, and let $P^{\rm trad}$ be the finite set of primes from Proposition \ref{residue_field_trad1}. We claim that the assertion holds whenever $\pp\not\in P'\cup P^{\rm trad}$.
To prove this let $\Ad$ denote the adjoint representation of $\GL_n$. Take any element $\delta\in\Delta_\pp$, and let $\mathop{\rm int}(\delta)$ denote the corresponding inner automorphism of $\GL_{n,k_\pp}$. Then $\Ad(\delta)$ is the derivative $d(\mathop{\rm int}(\delta))$, and its trace is an element of $k_\pp$.
On the other hand $\mathop{\rm int}(\delta)$ induces an algebraic automorphism of $\SL_{n,\bar k_\pp}$ which normalizes $\Delta_\pp^{\prime\rm der} = G'_\pp(k'_\pp)$. By the uniqueness in Proposition \ref{same_k_and_model} it therefore induces an algebraic automorphism of the model $G'_\pp$ over $k_\pp'$. The derivative of this automorphism is an automorphism of the Lie algebra $\mathop{\rm Lie}G'_\pp$, whose trace is therefore an element of $k_\pp'$. But the fact that $G'_\pp$ is a model of $\SL_{n,\bar k_\pp}$ yields an equivariant isomorphism $\mathop{\rm Lie}G'_\pp \otimes_{k_\pp'} \nobreak \bar k_\pp \allowbreak \cong \sll_n(\bar k_\pp)$, and so the trace in question is equal to the trace of $d(\mathop{\rm int}(\delta))|\sll_n(\bar k_\pp)$. Together we deduce that
$$\Tr\Ad(\delta) \ =\ \Tr\bigl(d(\mathop{\rm int}(\delta))|\sll_n(\bar k_\pp)\bigr) + 1
\ \in\ k'_\pp.$$
In particular, we can apply this to $\delta = \overline{\rho}_\pp(\Frob_x)$ for any closed point $x\in X$. Then $\Tr\Ad(\overline{\rho}_\pp(\Frob_x))$ is the image of $\Ad(\rho_\pp(\Frob_x))$ in the residue field $k_\pp$. Varying $x$, the elements $\Ad(\rho_\pp(\Frob_x))$ generate the ring of traces $R^{\text{trad}}$ from Subsection \ref{ring_of_trances}. Thus by Proposition \ref{residue_field_trad1} (a) their images generate the residue field $k_\pp$. Since these images also lie in $k'_\pp$, we deduce that $k_\pp \subset k_\pp'$, as desired.
\end{proof}
\begin{prop}\label{res_surj1}
For almost all primes $\pp\not\in P_0$ of $A$, we have $\Delta_\pp^{{\geom}} = \SL_n(k_\pp)$.
\end{prop}
\begin{proof}
We prove that this holds for all primes $\pp$ satisfying Lemmas \ref{almost_surjective2} and \ref{kp_is_kp1}. Indeed, Lemma \ref{almost_surjective2} shows that $G'_\pp(k'_\pp) = \Delta_\pp^{\prime\rm der} \subset \GL_n(k_\pp)^{\rm der} = \SL_n(k_\pp)$. Applying Proposition \ref{fieldcontained} with $G=\SL_{n,k_\pp}$, this implies that $|k_\pp'|\le|k_\pp|$. On the other hand we have $k_\pp \subset k_\pp'$ by Lemma \ref{kp_is_kp1}; hence together we deduce that $k_\pp = k_\pp'$. Applying Proposition \ref{same_model} with $G=\SL_{n,k_\pp}$ then shows that $G'(k'_\pp) = \SL_n(k_\pp)$. In particular we have $\SL_n(k_\pp) = \Delta_\pp^{\prime\rm der} \subset \Delta_\pp^{\rm der} \subset \Delta_\pp^{\text{geom}} \subset \SL_n(k_\pp)$, and so these inclusions are equalities, as desired.
\end{proof}
\subsection{Surjectivity at a single prime}\label{52}
\begin{prop}\label{surj1}
For almost all primes $\pp\not\in P_0$ of $A$,
we have $\Gamma_\pp^{\geom}=\SL_n(A_\pp)$.
\end{prop}
\begin{proof}
Let $P'$ be the finite set of primes $\pp$ excluded by Proposition \ref{res_surj1} or satisfying $|k_\pp|\leq9$. For all $\pp\not\in P'$ we have a surjective homomorphism $\Gamma_\pp^{{\geom}}\to\SL_n(k_\pp)$.
Suppose first that $F^{\text{trad}} = F$. Let $P^{\rm trad}$ be the finite set of primes from Proposition \ref{residue_field_trad1}. Then for any prime $\pp\not\in P'\cup P^{\rm trad}$, the set of traces $\Tr\Ad(\Gamma_\pp)$ topologically generates $A_\pp$. Applying Theorem \ref{trace_crit_1} to $\Gamma_\pp \subset \GL_n(A_\pp)$ thus shows that $\Gamma_\pp^{\rm der}=\SL_n(A_\pp)$.
Suppose now that $F^{\text{trad}}=F^2$. Then $p=n=2$ by Theorem \ref{field_trad}. By Lemma \ref{small_quotients1} there exists a finite set $P''\supset P_0$ of primes of $A$, such that the number of continuous homomorphisms from $G_K$ to a finite group of order $2$, which factor through the surjection $\rho_\pp:G_K\twoheadrightarrow\Gamma_\pp$ for some $\pp\not\in P''$, is finite. The intersection of the kernels of these homomorphisms is then $G_{K'}$ for some finite extension $K'\subset K^{\text{sep}}$ of $K$.
If $\Gamma_\pp'$ denotes the intersection of all closed subgroups of index $2$ of $\Gamma_\pp$, it follows that for all primes $\pp\not\in P''$ of $A$ we have $\rho_\pp(G_{K'}) \subset \Gamma_\pp'$.
Let $P^{\rm trad}$ be the finite set of primes obtained by applying Proposition \ref{residue_field_trad1} with $K'$ in place of $K$. Then for any prime $\pp\not\in P'\cup P''\cup P^{\rm trad}$, the set of traces $\Tr\Ad(\rho_\pp(G_{K'}))$, and hence a fortiori the set of traces $\Tr\Ad(\Gamma_\pp')$, topologically generates the subring $A_\pp^2 := \{a^2\mid a\in A_\pp\}$. Applying Theorem \ref{trace_crit_2} to $\Gamma_\pp \subset \GL_2(A_\pp)$ thus shows that $\Gamma_\pp^{\rm der}=\SL_2(A_\pp)$.
Since $\Gamma_\pp^{\rm der} \subset \Gamma_\pp^{{\geom}} \subset \SL_n(A_\pp)$, the proposition follows in either case.
\end{proof}
\subsection{Residual surjectivity at several primes}\label{53}
For any finite set $P$ of primes $\pp\not=\pp_0$ of $A$, we let
$$\Delta_P^{\geom}
\ \triangleleft\ \Delta_P\ \subset\ \prod_{\pp\in P} \;
(D_\pp/\pp D_\pp)^\times$$
denote the images of $G_K^{\geom} \triangleleft G_K$ under the combined representation induced by $\overline{\rho}_\pp$. Recall that $(D_\pp/\pp D_\pp)^\times \cong \GL_n(k_\pp)$ and $\Delta_\pp^{\geom} \subset \SL_n(k_\pp)$ whenever $\pp\not\in P_0$. Thus whenever $P\cap P_0=\varnothing$, we have
$$\Delta_P^{\geom} \ \subset\ \prod_{\pp\in P} \SL_n(k_\pp).$$
\begin{prop}\label{res_surj_fin}
There exists a finite set $P_1$ of primes of $A$ containing $P_0$, such that for any finite set of primes $P$ of $A$ satisfying $P\cap P_1=\varnothing$,
we have $\Delta_P^{\geom}= \prod_{\pp\in P} \SL_n(k_\pp)$.
\end{prop}
\begin{proof}
Let $P'$ be the finite set of primes $\pp$ excluded by Proposition \ref{res_surj1} or satisfying $|k_\pp|\leq3$. Let $P^{\rm trad}$ be the finite set of primes from Proposition \ref{residue_field_trad1}, and set $P_1:=P'\cup P^{\rm trad}$. We claim that the assertion holds whenever $P\cap P_1=\varnothing$.
For any $\pp\in P$ abbreviate $\PSL(n,k_\pp) := \SL_n(k_\pp) / Z(\SL_n(k_\pp))$. The assumption $|k_\pp|>3$ implies that this is a non-abelian finite simple group and that $\SL_n(k_\pp)$ is perfect. Let
$$\overline{\Delta_P^{\geom}} \ \subset\ \prod_{\pp\in P} \PSL(n,k_\pp)$$
denote the image of $\Delta_P^{\geom}$. Then it suffices to prove that this inclusion is an equality.
Assume otherwise. {}From Proposition \ref{res_surj1} we know that $\overline{\Delta_P^{\geom}}$ surjects to all factors. Since these factors are non-abelian simple groups, Goursat's Lemma implies that $\overline{\Delta_P^{\geom}}$ lies over the graph of an isomorphism between two factors, say associated to distinct primes $\pp_1,\pp_2\in P$. Then the situation persists after replacing $P$ by $\{\pp_1,\pp_2\}$; hence we may without loss of generality assume that $P=\{\pp_1,\pp_2\}$.
The isomorphism $\PSL(n,k_{\pp_1}) \stackrel{\sim}{\to} \PSL(n,k_{\pp_2})$ is induced by a field isomorphism $\sigma: k_{\pp_1} \stackrel{\sim}{\to} k_{\pp_2}$ and a corresponding isomorphism of algebraic groups $\alpha: \sigma^* \PGL_{r,k_{\pp_1}} \stackrel{\sim}{\to} \PGL_{r,k_{\pp_2}}$ (see \cite{PSTAX}, Lemmas 9.4 and 9.5). Since $\SL_n(k_{\pp_1}) \times \SL_n(k_{\pp_2})$ is a central extension of $\PSL(n,k_{\pp_1}) \times \PSL(n,k_{\pp_2})$, the derived group $(\Delta_P^{\geom})^{\rm der}$ of $\Delta_P^{\geom}$ depends only on $\overline{\Delta_P^{\geom}}$. It is therefore the graph of the isomorphism $\SL_n(k_{\pp_1}) \stackrel{\sim}{\to} \SL_n(k_{\pp_2})$ induced by the unique isomorphism $\tilde\alpha: \sigma^* \SL_{r,k_{\pp_1}} \stackrel{\sim}{\to} \SL_{r,k_{\pp_2}}$ lifting $\alpha$.
The uniqueness of the model from Proposition \ref{same_model} implies that the isomorphism $\tilde\alpha$ depends only on $(\Delta_P^{\geom})^{\rm der}$. Thus its graph depends only on $(\Delta_P^{\geom})^{\rm der}$. Since $\Delta_P$ normalizes $(\Delta_P^{\geom})^{\rm der}$ by construction, it thus also normalizes the graph of $\tilde\alpha$. In other words, for every $\delta=(\delta_1,\delta_2)\in\Delta_P$, the following diagram commutes:
$$\xymatrix{
\sigma^* \SL_{r,k_{\pp_1}} \ar[r]^-{\tilde\alpha}
\ar[d]_{\sigma^*{\rm int}(\delta_1)}
& \SL_{r,k_{\pp_2}} \ar[d]^{{\rm int}(\delta_2)} \\
\sigma^* \SL_{r,k_{\pp_1}} \ar[r]^-{\tilde\alpha}
& \SL_{r,k_{\pp_2}} \rlap{.}\\}$$
Taking traces and recalling that the trace on $\gll_n$ is the trace on $\sll_n$ plus $1$, we deduce that $\sigma(\Tr\Ad(\delta_1)) = \Tr\Ad(\delta_2)$.
In particular, we can apply this when $\delta$ is the image of $\Frob_x$ for any closed point $x\in X$. Then $\Tr\Ad(\delta_i) = \Tr\Ad(\overline{\rho}_{\pp_i}(\Frob_x))$ is the image of $\Ad(\rho_{\pp_i}(\Frob_x))$ in the residue field $k_{\pp_i}$, where $\Ad(\rho_{\pp_i}(\Frob_x)) \in R^{\text{trad}}$ is independent of $i$. Thus $\Tr\Ad(\delta) = \bigl( \Tr\Ad(\delta_1) , \sigma(\Tr\Ad(\delta_1)) \bigr)$ is the image of $\Ad(\rho_{\pp_1}(\Frob_x)) \in R^{\text{trad}}$ in the product of the residue fields $k_{\pp_1} \times k_{\pp_2}$. Since the elements $\Ad(\rho_{\pp_1}(\Frob_x))$ for all $x$ generate the ring of trances $R^{\text{trad}}$, it follows that the image of the reduction map $R^{\text{trad}} \to k_{\pp_1} \times k_{\pp_2}$ is contained in the graph of $\sigma$. But since $P
\cap P^{\rm trad}=\varnothing$, this contradicts Proposition \ref{residue_field_trad1} (b). Therefore $\overline{\Delta_P^{\geom}}$ cannot be a proper subgroup, and we are finished.
\end{proof}
\begin{lem} \label{res_indie}
There exists a finite set $P_2$ of primes $\pp\not=\pp_0$ of $A$ containing $P_0$, such that for every finite $P\supset P_2$ and every $\pp\not\in P$, we have
$$\Delta^{\geom}_{P\cup\{\pp\}} \ =\ \Delta^{\geom}_P \times \SL_n(k_\pp).$$
\end{lem}
\begin{proof}
Let $P_1$ be the finite set of primes excluded by Proposition \ref{res_surj_fin}.
Let $N$ be the maximum of the orders of all Jordan-H\"older constituents of the finite group $\Delta^\text{geom}_{P_1}$. Let $P_2$ be the union of $P_1$ with the set of primes $\pp$ for which $|\PSL(n,k_\pp)|\leq N$ or $|k_\pp|\leq9$. We will prove the assertion whenever $P\supset P_2$.
Consider the natural inclusion
$$\Delta^{\geom}_{P\cup\{\pp\}} \ \subset\
\Delta^{\geom}_{P_1} \times \Delta^{\geom}_{P\smallsetminus P_1} \times \SL_n(k_\pp).$$
By definition the image of $\Delta^{\geom}_{P\cup\{\pp\}}$ under the projection to the second and third factors is the subgroup
$$\Delta^{\geom}_{(P\smallsetminus P_1)\cup\{\pp\}} \ \subset\
\Delta^{\geom}_{P\smallsetminus P_1} \times \SL_n(k_\pp)
\ \subset\prod_{\pp'\in P\smallsetminus P_1}\!\! \SL_n(k_{\pp'}) \ \times\ \SL_n(k_\pp).$$
These inclusions are equalities by Proposition \ref{res_surj_fin}. Therefore the projection homomorphism $\Delta^{\geom}_{P\cup\{\pp\}} \to \Delta^{\geom}_{P\smallsetminus P_1} \times \SL_n(k_\pp)$ is surjective. From this it follows that
$$E\ :=\ \Delta^{\geom}_{P\cup\{\pp\}} \cap \bigl(\Delta^{\geom}_{P_1} {\times} \{1\} {\times} \SL_n(k_\pp) \bigr)$$
surjects to $\SL_n(k_\pp)$. In particular $\PSL(n,k_\pp)$ is a Jordan-H\"older factor of $E$. The assumption $\pp\not\in P_1$ implies that the order of $\PSL(n,k_\pp)$ is greater than the order of any Jordan-H\"older constituent of $\Delta^\text{geom}_{P_1}$. Thus $\PSL(n,k_\pp)$ cannot be a Jordan-H\"older constituent of the image of $E$ in $\Delta^\text{geom}_{P_1}$. It must therefore be a Jordan-H\"older factor of $\Delta^{\geom}_{P\cup\{\pp\}} \cap \bigl(\{1\} {\times} \{1\} {\times} \SL_n(k_\pp) \bigr)$. Since $\SL_n(k_\pp)$ is perfect, it follows that
$$\{1\} {\times} \{1\} {\times} \SL_n(k_\pp) \ \subset\
E \ \subset\ \Delta^{\geom}_{P\cup\{\pp\}}.$$
The short exact sequence
$$\xymatrix{
1 \ar[r] &
\{1\} {\times} \{1\} {\times} \SL_n(k_\pp) \ar[r] &
\Delta^{\geom}_{P\cup\{\pp\}} \ar[r] &
\Delta^{\geom}_P \ar[r] & 1\\}$$
and the $5$-Lemma then show that $\Delta^{\geom}_{P\cup\{\pp\}} \ =\ \Delta^{\geom}_P \times \SL_n(k_\pp)$, as desired.
\end{proof}
\subsection{Adelic openness}\label{54}
For any finite set $P$ of primes $\pp\not=\pp_0$ of $A$, we let $\Gamma_P^{\text{geom}}$ denote the image of the combined homomorphism
$$(\rho_\pp)_{\pp\in P}:\
G_K^{\text{geom}} \longrightarrow \prod_{\pp\in P} D_\pp^1.$$
Recall that $D_\pp^1 \cong \SL_n(A_\pp)$ whenever $\pp\not\in P_0$.
\begin{lem} \label{indie}
There exists a finite set $P_3$ of primes $\pp\not=\pp_0$ of $A$ containing $P_0$, such that for every finite $P\supset P_3$ and every $\pp\not\in P$, we have
$$\Gamma^{\geom}_{P\cup\{\pp\}} \ =\ \Gamma^{\geom}_P \times \SL_n(A_\pp).$$
\end{lem}
\begin{proof}
Let $P'$ be the finite set of primes $\pp$ excluded by Proposition \ref{surj1} or satisfying $|k_\pp|\leq9$. Let $P_3$ be the union of $P'$ with the set of primes $P_2$ from Lemma \ref{res_indie}. We will prove the assertion whenever $P\supset P_3$.
For this we consider the commutative diagram
\def\hbox{$\xymatrix@R-4pt{\ar@{->>}[d]\\ \\}$}{\hbox{$\xymatrix@R-4pt{\ar@{->>}[d]\\ \\}$}}
$$\begin{array}{ccccc}
\Gamma^{\geom}_{P\cup\{\pp\}}
&\subset& \Gamma^{\geom}_P
&\!\!\!\times\!\!\!& \SL_n(A_\pp) \\[-7pt]
\hbox{$\xymatrix@R-4pt{\ar@{->>}[d]\\ \\}$} && \hbox{$\xymatrix@R-4pt{\ar@{->>}[d]\\ \\}$} && \hbox{$\xymatrix@R-4pt{\ar@{->>}[d]\\ \\}$} \\
\Delta^{\geom}_{P\cup\{\pp\}}
&\subset& \Delta^{\geom}_P
&\!\!\!\times\!\!\!& \SL_n(k_\pp)\rlap{.}
\end{array}$$
The inclusion in the lower row is an equality by Lemma \ref{res_indie}. Thus if $H$ denotes the kernel of the surjection $\Gamma^{\geom}_P \twoheadrightarrow \Delta^{\geom}_P$, it follows that $\Gamma^{\geom}_{P\cup\{\pp\}} \cap \bigl(H {\times} \SL_n(A_\pp) \bigr)$ surjects to $\{1\}\times\SL_n(k_\pp)$. But by construction $H$ is a pro-$p$-group, and $\SL_n(k_\pp)$ has no Jordan-H\"older factor of order $p$. Since all groups in question are pro-finite, we deduce that
$$\Gamma_\pp'\ :=\ \Gamma^{\geom}_{P\cup\{\pp\}}
\cap \bigl(\{1\} {\times} \SL_n(A_\pp) \bigr)$$
also surjects to $\SL_n(k_\pp)$.
By construction $\Gamma'_\pp$ is a closed normal subgroup of $\Gamma^{\geom}_{\{\pp\}\cup P}$, and the conjugation action of $\Gamma^{\geom}_{\{\pp\}\cup P}$ on it factors through the projection $\Gamma^{\geom}_{\{\pp\}\cup P} \twoheadrightarrow \Gamma^{\geom}_\pp \subset \SL_n(A_\pp)$. Since $\pp\not\in P'$, the last inclusion is an equality by Proposition \ref{surj1}. Together this implies that $\Gamma'_\pp$ is normalized by $\SL_n(A_\pp)$.
Combining this with the assumption $|k_\pp|>9$ and the fact that $\Gamma_\pp'$ surjects to $\SL_n(k_\pp)$, Proposition \ref{normal_sa} now implies that $\Gamma'_\pp = \{1\}\times\SL_n(A_\pp)$. The short exact sequence
$$\xymatrix{
1 \ar[r] &
\{1\} {\times} \SL_n(A_\pp) \ar[r] &
\Gamma^{\geom}_{P\cup\{\pp\}} \ar[r] &
\Gamma^{\geom}_P \ar[r] & 1\\}$$
and the $5$-Lemma then show that $\Gamma^{\geom}_{P\cup\{\pp\}} \ =\ \Gamma^{\geom}_P \times \SL_n(A_\pp)$, as desired.
\end{proof}
\begin{proof}[Proof of Theorem \ref{main_theorem} (a).]
Let $P_3$ be as in Lemma \ref{indie}. Then induction on $P$ shows that for every finite $P\supset P_3$, we have
$$\Gamma^{\geom}_P \ =\ \Gamma^{\geom}_{P_3} \times \prod_{\pp\in P\smallsetminus P_3} \SL_n(A_\pp).$$
In the limit this implies that
$$\rho_{\text{ad}}(G_K^{\text{geom}}) \ =\ \Gamma^{\geom}_{P_3} \times \prod_{\pp\not\in P_3} \SL_n(A_\pp).$$
But $\Gamma^{\text{geom}}_{P_3}$ has finite index in $\prod_{\pp\in P_3} D^1_\pp$ by Theorem \ref{openness}. Therefore $\rho_{\text{ad}}(G_K^{\text{geom}})$ has finite index in $\prod_{\pp\not=\pp_0} D^1_\pp$, as desired.
\end{proof}
\subsection{Absolute Galois group}\label{55}
\begin{proof}[Proof of Theorem \ref{main_theorem} (b).]
Recall that $R := \End_K(\phi) = \End_{K^{\text{sep}}}(\phi)$ by Assumption \ref{asses} (a), and that $D_\pp$ was defined as the commutant of $R_\pp := R\otimes_AA_\pp$ in $\End_{A_\pp}(T_\pp(\phi))$. Thus $\rho_{\text{ad}}(G_K)$ is contained in $\prod_{\pp\not=\pp_0} D^\times_\pp$. We will look at its image under the determinant map.
Let $F'$ be a maximal commutative $F$-subalgebra of $R\otimes_A F$, let $A'$ denote the integral closure of $A$ in~$F'$, and choose a Drinfeld $A'$-module $\phi'\colon A'\to K\{\tau\}$ and an isogeny $f\colon\phi\to\phi'|A$, as in Subsection \ref{tatess}. The characteristic of $\phi'$ is then a prime $\pp_0'$ of $A'$ that divides~$\pp_0$.
By Anderson \cite{Anderson_t_Motives}, \S4.2, there exists a Drinfeld $A'$-module $\psi'\colon A'\to K\{\tau\}$ of rank~$1$ and characteristic $\pp_0'$ whose adelic Galois representation is isomorphic to the determinant of the adelic Galois representation associated to~$\phi'$. With Proposition \ref{boalo} it follows that the composite homomorphism
$$\xymatrix{
\det\rho_{\text{ad}}\colon\ G_K \ar[r]^-{\rho_{\text{ad}}} &
\prod\limits_{\pp\not=\pp_0}\!\! D^\times_\pp \ar[r]^-{\det} &
\prod\limits_{\pp\not=\pp_0}\!\! A^\times_\pp \ar@{^{ (}->}[r] &
\prod\limits_{\pp'\nmid\pp_0}\!\! A^{\prime\times}_{\pp'} \\ }$$
describes the Galois representation on the Tate modules $\prod_{\pp'\nmid\pp_0} T_{\pp'}(\psi')$.
Without loss of generality we may assume that $\psi'$ is defined over the finite field~$\kappa$. Let $m$ denote the degree of $\kappa$ over~$\mathbb{F}_p$. Then $\Frob_\kappa = \tau^m$ lies in the center of $\kappa\{\tau\}$. In particular it commutes with $\psi'_{a'}$ for all $a'\in A'$ and is therefore an endomorphism of~$\psi'$. As $\psi'$ has rank $1$, its endomorphism ring is equal to~$A'$; hence $\Frob_\kappa$ represents an element $a'\in A'$. The action of $\Frob_\kappa$ as an element of the Galois group $G_\kappa$ on all Tate modules of $\psi'$ is then just multiplication by~$a'$. Since $a'$ is the single eigenvalue of $\Frob_\kappa$ associated to $\psi'$, Proposition \ref{F_eigenvalues} implies that $a'$ is divisible by $\pp'_0$ but not by any other prime of~$A'$.
For every element $\sigma\in G_K$ whose restriction to $\bar\kappa$ is $\Frob_\kappa$ we thus have $\det\rho_{\text{ad}}(\sigma) = a'$ diagonally embedded into $\prod_{\pp'\nmid\pp_0} A^{\prime\times}_{\pp'}$. But it also lies in the subgroup $\prod_{\pp\not=\pp_0} A^\times_\pp$, whose intersection with the diagonally embedded $A'$ is~$A$. Thus $a'$ is actually an element of~$A$, divisible by $\pp_0$ but not by any other prime of~$A$. Moreover, we have $\det\rho_{\text{ad}}(G_K) = \smash{\overline{\langle a'\rangle}}$, the pro-cyclic subgroup topologically generated by~$a'$.
Now both $a'$ and the $a_0$ in Theorem \ref{main_theorem} are elements of $A$ that are divisible by $\pp_0$ but not by any other prime of~$A$. Thus the corresponding ideals are $(a')=\pp_0^i$ and $(a_0)=\pp_0^j$ for some positive integers $i$ and~$j$. Together it follows that $(a^{\prime j}) = \pp_0^{ij} = (a_0^i)$, and so $a^{\prime j}/a_0^i$ is a unit in~$A^\times$. As the group of units is finite, we deduce that $a^{\prime j\ell} = a_0^{i\ell}$ for some positive integer~$\ell$. Thus the subgroup $\smash{\overline{\langle a'\rangle}}$ is commensurable to $\smash{\overline{\langle a_0\rangle}}$.
On adjoining to $K$ a suitable finite extension of the constant field~$\kappa$ we can replace $a'$ by any positive integral power. We can therefore reduce ourselves to the case that $\smash{\overline{\langle a'\rangle}} \subset \smash{\overline{\langle a_0^n\rangle}}$ with $n$ as in Assumption \ref{asses} (c). Then $\det(a_0) = a_0^n$, and from this we see that the middle row in the following commutative diagram is exact and the upper right rectangle is cartesian. This together with the inclusion $\smash{\overline{\langle a'\rangle}} \subset \smash{\overline{\langle a_0^n\rangle}}$ yields the inclusions in the lower half of the diagram:
$$\xymatrix@R-10pt@C+10pt{
1 \ar[r] &
\prod\limits_{\pp\not=\pp_0}\!\!D^1_\pp \ar[r] &
\prod\limits_{\pp\not=\pp_0}\!\!D^\times_\pp \ar[r]^-{\det} &
\prod\limits_{\pp\not=\pp_0}\!\!A^\times_\pp & \\
1 \ar[r] &
\prod\limits_{\pp\not=\pp_0}\!\!D^1_\pp \ar[r] \ar@{}[u]|<<<<{\big\Vert} &
\overline{\langle a_0\rangle} \cdot \!\prod\limits_{\pp\not=\pp_0}\!\!D^1_\pp \ar[r] \ar@{}[u]|<<<<{\bigcup} &
\overline{\langle a_0^n\rangle} \ar[r] \ar@{}[u]|<<<<{\bigcup} & 1 \\
1 \ar[r] &
\rho_{\text{ad}}(G_K) \cap \!\prod\limits_{\pp\not=\pp_0}\!\!D^1_\pp \ar[r] \ar@{}[u]|<<<<{\bigcup} &
\rho_{\text{ad}}(G_K) \ar[r] \ar@{}[u]|<<<<{\bigcup} &
\overline{\langle a'\rangle} \ar[r] \ar@{}[u]|<<<<{\bigcup} &1\rlap{.} \\}$$
Theorem \ref{main_theorem} (a) implies that the inclusion at the lower left is of finite index. By the above the same is true for the inclusion at the lower right. Since the bottom row is also exact, it follows that the inclusion at the lower middle is also of finite index. This shows that $\rho_{\text{ad}}( G_K)$ is commensurable to $\overline{\langle a_0\rangle} \cdot \prod_{\pp\not=\pp_0} D^1_\pp$, finishing the proof of Theorem \ref{main_theorem}~(b).
\end{proof}
\section{Arbitrary endomorphism ring}
\label{generalcase}
As in Section \ref{notation}, we let $K$ be a field that is finitely generated over a finite field $\kappa$ and let $\phi: A \rightarrow K\{\tau\}$ be a Drinfeld $A$-module of rank $r$ over $K$ of special characteristic~$\pp_0$. We keep the relevant notations of Section \ref{notation}, but do not impose any other restrictions.
Set $R:= \End_{K^{\text{sep}}}(\phi)$ and $F := \mathop{\rm Quot}(A)$. Then $R \otimes_A F$ is a division algebra of finite dimension over $F$. Let $Z$ denote its center and write
$$\dim_Z (R\otimes_A F) =d^2 \quad \text{and} \quad [Z/F]=e.$$
Then $de$ divides $r$ by Proposition \ref{numerics}.
\subsection{The isotrivial case}\label{61}
\begin{defn}\label{isotrivialdef}
We call $\phi$ \emph{isotrivial} if over some field extension it is isomorphic to a Drinfeld $A$-module defined over a finite field.
\end{defn}
Clearly this property is invariant under extending~$K$.
\begin{prop}\label{isotrivialprop}
\begin{enumerate}
\item[(a)] $\phi$ is isotrivial if and only if it is isomorphic over $K^{\text{sep}}$ to a Drinfeld $A$-module defined over a finite subfield of $K^{\text{sep}}$.
\item[(b)] Let $\phi'$ be another Drinfeld $A$-module over $K$ that is isogenous to~$\phi$. Then $\phi$ is isotrivial if and only if $\phi'$ is isotrivial.
\item[(c)] Let $B$ be any integrally closed infinite subring of~$A$. Then $\phi$ is isotrivial if and only if $\phi|B$ is isotrivial.
\end{enumerate}
\end{prop}
\begin{proof}
In (a) the `if' part is obvious. For the `only if' part assume that $L$ is a field extension of $K$ such that $\phi$ is isomorphic over $L$ to a Drinfeld $A$-module $\psi$ defined over a finite subfield $\ell\subset L$. By the definition of isomorphisms there is then an element $u\in L^\times$ such that $\phi_a=u\circ\psi_a\circ u^{-1}$ in $L\{\tau\}$ for all $a\in A$. Choose a prime $\pp\not=\pp_0$ of $A$ and, after replacing $L$ by a finite extension, a non-zero torsion point $t\in\phi[\pp](L)$. Then $t$ is separably algebraic over~$K$. On the other hand $ut$ is a non-zero torsion point of $\psi$ and therefore algebraic over~$\ell$. Since $\ell$ is finite, $ut$ is actually separable over~$\ell$. Thus the subfiend $K\ell(u,ut)\subset L$ is separably algebraic over~$K$ and can therefore be embedded into $K^{\text{sep}}$. Then $u=ut/t$ defines an isomorphism $\phi\cong\psi$ over $K^{\text{sep}}$, as desired.
In (b) by symmetry it suffices to prove the `if' part. So assume that $L$ is a field extension of $K$ such that $\phi$ is isomorphic over $L$ to a Drinfeld $A$-module $\psi$ defined over a finite subfield $\ell\subset L$. Then $\phi'$ is isogenous to $\psi$ over~$L$. By the definition of isogenies this means that there is a non-zero element $f\in L\{\tau\}$ such that $\phi'_a\circ f=f\circ\psi_a$ for all $a\in A$. Its scheme theoretic kernel $\mathop{\rm ker}(f)$ is then a finite subgroup scheme of ${\mathbb G}_{a,L}$ that is mapped to itself under $\psi_a$ for all $a\in A$. Its identity component is a finite infinitesimal subgroup scheme of ${\mathbb G}_{a,L}$ and therefore the kernel of some power of~$\tau$. On the other hand all its geometric points are torsion points of $\psi$ and therefore algebraic over~$\ell$. Together it follows that $\mathop{\rm ker}(f)$ is defined over some finite extension $\ell'\subset L$ of $\ell$ and is therefore the kernel of some non-zero element $g\in L\{\tau\}$. Since $\mathop{\rm ker}(f) = \mathop{\rm ker}(g)$, it now follows that $f=u\circ g$ for some element $u\in L^\times$. Consider the Drinfeld $A$-module $\psi': A\to L\{\tau\}$ defined by $\psi'_a := u^{-1}\circ\phi'_a\circ u$. Then the relation $\phi'_a\circ f=f\circ\psi_a$ implies that $\psi'_a\circ g=g\circ\psi_a$ for all $a\in A$. Since $g$ and $\psi_a$ have coefficients in~$\ell'$, this relation implies that $\psi'_a$ also has coefficients in~$\ell'$. In other words $\psi'$ is really defined over~$\ell'$, and since $\phi'\cong\psi'$, it follows that $\phi'$ is isotrivial, as desired.
In (c) the `only if' part is obvious. For the `if' part assume that $L$ is a field extension of $K$ such that $\phi|B$ is isomorphic over $L$ to a Drinfeld $B$-module $\psi'$ defined over a finite subfield $\ell\subset L$. By the definition of isomorphisms there is then an element $u\in L^\times$ such that $\phi_b=u\circ\psi'_b\circ u^{-1}$ in $L\{\tau\}$ for all $b\in B$. Consider the Drinfeld $A$-module $\psi: A\to L\{\tau\}$ defined by $\psi_a := u^{-1}\circ\phi_a\circ u$. By construction it satisfies $\psi|B=\psi'$; hence it defines an embedding $B\hookrightarrow \End_L(\psi')$. Thus by Proposition \ref{def_field_end} applied to $\psi'$ over $\ell$ the coefficients of $\psi_a$ for all $a\in A$ lie in some fixed finite extension $\ell'$ of~$\ell$. This means that $\psi$ is really defined over~$\ell'$, and since $\phi\cong\psi$, it follows that $\phi$ is isotrivial, as desired.
\end{proof}
\begin{prop}\label{finite}
The following assertions are equivalent:
\begin{enumerate}
\item[(a)] $\phi$ is isotrivial.
\item[(b)] $\rho_{\text{ad}}( G_K^{\geom})$ is finite.
\item[(c)] $\rho_\pp( G_K^{\geom})$ is finite for every prime $\pp\not=\pp_0$ of $A$.
\item[(d)] $\rho_\pp( G_K^{\geom})$ is finite for some prime $\pp\not=\pp_0$ of $A$.
\item[(e)] $de=r$.
\end{enumerate}
\end{prop}
\begin{proof}
(Compare \cite{PinII}, Proposition 2.2.)
The implications (a)$\Rightarrow$(b)$\Rightarrow$(c)$\Rightarrow$(d) are obvious.
For the rest of the proof we may assume that $\End_K(\phi)=R$ after replacing $K$ by a finite extension, using Proposition \ref{def_field_end}. Let $F'$ be a maximal commutative $F$-subalgebra of $R\otimes_A F$, let $A'$ denote the integral closure of $A$ in~$F'\!$, and choose a Drinfeld $A'$-module $\phi'\colon A'\to K\{\tau\}$ and an isogeny $f\colon\phi\to\phi'|A$, as in the proof of Proposition \ref{numerics}. Then $\phi'$ has rank~$r/de$ and endomorphism ring $\End_K(\phi') = A'$.
If (d) holds, there exist a prime $\pp\not=\pp_0$ of $A$ and a finite extension $K'\subset K^{\text{sep}}$ of $K$ such that $\rho_\pp( G_{K'}^{\geom})$ is trivial and hence $\rho_\pp(G_{K'})$ is abelian. After replacing $K$ by $K'$ we may therefore assume that $\rho_\pp(G_K)$ is abelian. Moreover, as in (\ref{eq:tensisom}) we have a $G_K$-equivariant isomorphism $V_\pp(\phi) \cong V_\pp(\phi'|A) \cong \prod_{\pp'|\pp} V_{\pp'}(\phi')$. Thus for any prime $\pp'|\pp$, the image $\rho_{\pp'}(G_K)$ of the Galois representation $\rho_{\pp'}$ on $V_{\pp'}(\phi')$ is abelian, and so the subring $F'_{\pp'}[\rho_{\pp'}(G_K)]$ of $\End_{F'_{\smash{\pp'}}}(V_{\pp'}(\phi'))$ is commutative. By the semisimplicity and Tate conjectures for Drinfeld modules (see \cite{Tagu}, \cite{Tama2}, \cite{Tama1}, \cite{Tama3}) this subring is the commutant of $\End_K(\phi')\otimes_{A'}F'_{\pp'}$. But as $\End_K(\phi') = A'$, this commutant is equal to $\End_{F'_{\smash{\pp'}}}(V_{\pp'}(\phi'))$. It is therefore commutative if and only if $r/de = \dim_{F'_{\smash{\pp'}}}(V_{\pp'}(\phi')) \leq 1$. Thus (d) implies (e).
If (e) holds, then $\phi'$ is a Drinfeld $A'$-module of rank~$1$ and of special characteristic. Since the moduli stack of Drinfeld $A'$-modules of rank $1$ is finite over $\Spec A'$, the Drinfeld module $\phi'$ is isomorphic to one defined over a finite field, i.e., isotrivial.
By Proposition \ref{isotrivialprop} the same then also follows for $\phi'|A$ and for~$\phi$. Thus (e) implies (a), and we are done.
\end{proof}
To determine the images of Galois up to commensurability for an isotrivial Drinfeld module we may reduce ourselves to the case of a Drinfeld module defined over a finite field. In that case the situation is as follows:
\begin{prop}\label{FinGal}
Suppose that $\phi$ is defined over a finite field~$\kappa$. Let $C$ denote the center of $\End_\kappa(\phi)$ and $C'$ the normalization of~$C$. Then there exists an element $c_0\in C$ with the properties:
\begin{enumerate}
\item[(a)] $c_0$ generates a positive power of a unique prime $\pp'_0$ of $C'$ above~$\pp_0$.
\item[(b)] $\rho_{\text{ad}}(\Frob_\kappa)$ coincides with the action of $c_0$ on $\prod_{\pp\not=\pp_0} T_\pp(\phi)$.
\item[(c)]
$\rho_{\text{ad}}(G_\kappa) = \smash{\overline{\langle c_0\rangle}}$, the pro-cyclic subgroup topologically generated by~$c_0$.
\end{enumerate}
\end{prop}
\begin{proof}
Let $m$ denote the degree of $\kappa$ over~$\mathbb{F}_p$. Then $\Frob_\kappa = \tau^m$ lies in the center of $\kappa\{\tau\}$. In particular it commutes with $\phi_a$ for all $a\in A$ and is therefore an endomorphism of~$\phi$, and more specifically it lies in the center $C$ of $\End_\kappa(\phi)$. As such let us denote it by~$c_0$. The action of $\Frob_\kappa$ as an element of the Galois group $G_\kappa$ on all Tate modules of $\phi$ is then the same as that obtained from the natural action of $c_0$ as an endomorphism. This directly implies (b) and (c).
For (a) we apply Proposition \ref{ConstIsog} to $S:= C$ and $S':=C'$, obtaining a Drinfeld $C'$-module $\phi'\colon C'\to \kappa\{\tau\}$ and an isogeny $f\colon\phi\to\phi'|A$. The characteristic of $\phi'$ is then a prime $\pp'_0$ of $C'$ above~$\pp_0$. Also, the endomorphism $\Frob_\kappa$ of $\phi'$ still corresponds to the same element $c_0\in C$. Since $c_0$ acts as a scalar on the Tate modules of $\phi'$, it constitutes the single eigenvalue of $\Frob_\kappa$. Thus Proposition \ref{F_eigenvalues} implies that $c_0$ is divisible by $\pp'_0$ but not by any other prime of~$C'$. This shows (a), and we are done.
\end{proof}
\subsection{The non-isotrivial case}\label{62}
To determine the images of Galois in the general non-isotrivial case, we will use some reduction steps which end in the situation of Theorem \ref{main_theorem}. Recall that Theorem \ref{main_theorem} involves two conditions, namely that $A$ is the center of $R:=\End_{K^{\text{sep}}}(\phi)$ and that the endomorphism ring does not grow under restriction of~$A$. We will achieve the first condition by enlarging~$A$, and then the second condition by shrinking $A$ again until the endomorphism ring stops growing. That this process terminates is a non-trivial fact from \cite{PinII}.
\medskip
To enlarge $A$ we first choose a finite extension $K' \subset K^{\text{sep}}$ of $K$ such that $R=\End_{K'}(\phi)$. Recall that $Z$ denotes the center of $R \otimes_A F$; hence $C := Z\cap R$ is the center of~$R$. Let $C'$ denote the normalization of~$C$. Applying Proposition \ref{ConstIsog} to $S:= C$ and $S':=C'$ over~$K'$, we obtain a Drinfeld $C'$-module $\phi'\colon C'\to K'\{\tau\}$ and an isogeny $f\colon\phi\to\phi'|A$ over~$K'$. The characteristic of $\phi'$ is then a prime $\pp'_0$ of $C'$ above~$\pp_0$. Since $R\otimes_AF \cong \End_{K^{\text{sep}}}(\phi')\otimes_AF$, the construction implies that $C'$ is the center of $\End_{K^{\text{sep}}}(\phi')$. Also, the isogeny $f$ induces a $G_{K'}$-equivariant inclusion of finite index
\begin{myequation}\label{TateIsog}
\prod_{\pp\not=\pp_0} T_\pp(\phi)
\ \hookrightarrow\
\prod_{\pp\not=\pp_0} T_\pp(\phi'|A).
\end{myequation}%
Next we restrict $\phi'$ to suitable subrings $B$ of~$C'$. By Theorem 6.2 of \cite{PinII} there is a canonical choice for which the endomorphism ring of $\phi'|B$ is maximal:
\begin{prop}\label{ringB}
If $\phi$ is not isotrivial, there exists a unique integrally closed infinite subring $B$ of $C'$ with the following properties:
\begin{enumerate}
\item[(a)] The center of $\End_{K^{\text{sep}}}(\phi'|B)$ is $B$.
\item[(b)] For every integrally closed infinite subring $B'$ of $C'$ we have $\End_{K^{\text{sep}}}(\phi'|B') \allowbreak \subset\nobreak
\End_{K^{\text{sep}}}(\phi'|B)$.
\end{enumerate}
\end{prop}
With $B$ as in Proposition \ref{ringB} we abbreviate $\psi := \phi'|B$ and $S:= \End_{K^{\text{sep}}}(\psi)$. The characteristic of $\psi$ is $\pq_0:=B\cap\pp_0'$ and hence a maximal ideal of~$B$. By Proposition 3.5 of \cite{PinII} we have:
\begin{prop}\label{uniqueprime}
$\pp_0'$ is the unique prime of $C'$ above $\pq_0$.
\end{prop}
Let $P_0'$ denote the finite set of primes of $C'$ lying above $\pp_0$; then in particular $\pp'_0\in P_0'$. Let $Q_0$ denote the finite set of primes $\pq$ of $B$ such that all primes of $C'$ above $\pq$ lie in~$P_0'$. Then Proposition \ref{uniqueprime} implies that $\pq_0\in Q_0$. Combining the natural isomorphisms $T_\pq(\psi) \cong \prod_{\pp'|\pq} T_{\pp'}(\phi')$ for all primes $\pq\not\in Q_0$ of $B$ and the natural isomorphisms $T_\pp(\phi'|A) \cong \prod_{\pp'|\pp} T_{\pp'}(\phi')$ for all primes $\pp\not=\pp_0$ of~$A$, we obtain a natural $G_{K'}$-equivariant surjection
\begin{myequation}\label{TateCompare}
\prod_{\pq\not\in Q_0} T_\pq(\psi) \ \cong\
\prod_{\pq\not\in Q_0} \prod_{\pp'|\pq} T_{\pp'}(\phi') \ \twoheadrightarrow\
\prod_{\pp'\not\in P'_0} T_{\pp'}(\phi') \ \cong\
\prod_{\pp\not=\pp_0} T_\pp(\phi'|A).
\end{myequation}%
For every prime $\pq\not\in Q_0$ of $B$ let $D_\pq$ denote the commutant of $S\otimes_BB_\pq$ in $\End_{B_\pq}(T_\pq(\psi))$. As in Proposition \ref{boalo} (a) this is an order in a central simple algebra over the quotient field of~$B_\pq$. The product of these rings acts on the left hand side in (\ref{TateCompare}).
\begin{lem}\label{equivariance}
The kernel of the surjection (\ref{TateCompare}) is a $\vphantom{\Big|}\prod_{\pq\not\in Q_0} D_\pq$-submodule, and the induced action of\/ $\prod_{\pq\not\in Q_0} D_\pq$ on the quotient\/ $\prod_{\pp\not=\pp_0} T_\pp(\phi'|A)$ is faithful.
\end{lem}
\begin{proof}
For every prime $\pq\not\in Q_0$ of~$B$, the isomorphism $T_\pq(\psi) \cong \prod_{\pp'|\pq} T_{\pp'}(\phi')$ is the isotypic decomposition of $T_\pq(\psi)$ under $C'\otimes_BB_\pq$. Since $C'$ is contained in~$S$, the definition of $D_\pq$ shows that the actions of $C'\otimes_BB_\pq$ and $D_\pq$ commute; hence the decomposition is $D_\pq$-invariant. As the kernel of the surjection (\ref{TateCompare}) is a product of certain factors $T_{\pp'}(\phi')$, this implies the first assertion of the lemma. For the second note that, by the construction of~$Q_0$, for every prime $\pq\not\in Q_0$ of $B$ there exists a prime $\pp'\not\in P_0'$ of $C'$ with $\pp'|\pq$. Then $T_{\pp'}(\phi')$ is a non-trivial module over $D_\pq$, and it remains so after tensoring with the quotient field of~$B_\pq$; hence $D_\pq$ acts faithfully on it. Taking the product over all $\pq\not\in Q_0$ proves the second assertion.
\end{proof}
Let ${\mathcal D}$ denote the stabilizer in $\prod_{\pq\not\in Q_0} D_\pq$ of the image of the homomorphism (\ref{TateIsog}). By construction this is a closed subring of finite index, and Lemma \ref{equivariance} implies that ${\mathcal D}$ acts faithfully on $\prod_{\pp\not=\pp_0} T_\pp(\phi)$. For each $\pq\not\in Q_0$ let $D^1_\pq$ denote the multiplicative group of elements of $D_\pq$ of reduced norm~$1$. Then ${\mathcal D}^1 := {\mathcal D}^\times \cap \prod_{\pq\not\in Q_0} D_\pq^1$ is a closed subgroup of finite index of $\prod_{\pq\not\in Q_0} D_\pq^1$. We can identify ${\mathcal D}^\times$ and ${\mathcal D}^1$ with closed subgroups of $\prod_{\pp\not=\pp_0}\Aut_{A_\pp}(T_\pp(\phi))$.
Finally let $c_0$ be any element of $C'$ that generates a positive power of~$\pp'_0$. Let ${\mathfrak c}' \subset C'$ be the annihilator ideal of the cokernel of the inclusion (\ref{TateIsog}). Then ${\mathfrak c}'\not\subset\pp'_0$; hence it is relatively prime to~$c_0$. Thus after replacing $c_0$ by some positive power we may assume that $c_0\equiv1$ modulo~${\mathfrak c}'$. Then multiplication by $c_0$ is an automorphism of $\prod_{\pp\not=\pp_0} T_\pp(\phi'|A)$ that maps the image of (\ref{TateIsog}) to itself. We can thus view it as an element of $\prod_{\pp\not=\pp_0}\Aut_{A_\pp}(T_\pp(\phi))$. Let $\smash{\overline{\langle c_0\rangle}}$ denote the pro-cyclic subgroup of $\prod_{\pp\not=\pp_0}\Aut_{A_\pp}(T_\pp(\phi))$ that is topologically generated by it. Since $c_0\in C' \subset S$, this subgroup commutes with the action of ${\mathcal D}$ and hence with~${\mathcal D}^1$.
\begin{thm} \label{main_theorem_2}
Let $\phi$ be a non-isotrivial Drinfeld $A$-module over a finitely generated field $K$ of special characteristic~$\pp_0$. Let ${\mathcal D}^1$ and $\smash{\overline{\langle c_0\rangle}}$ denote the subgroups of\/ $\prod_{\pp\not=\pp_0}\Aut_{A_\pp}(T_\pp(\phi))$ defined above. Then
\begin{enumerate}
\item[(a)] $\rho_{\text{ad}}( G_K^{\geom})$ is commensurable to ${\mathcal D}^1$, and
\item[(b)] $\rho_{\text{ad}}( G_K)$ is commensurable to $\overline{\langle c_0\rangle} \cdot {\mathcal D}^1$.
\end{enumerate}
\end{thm}
\begin{proof}
By Proposition \ref{ringB} the assumptions of Theorem \ref{main_theorem} are satisfied for the Drinfeld $B$-module $\psi$ over~$K'$. Let $\rho^\psi_{\text{ad}} :\ G_{K'} \to \prod_{\pq\not\in Q_0} D_\pq^\times$ denote the homomorphism describing the action of $G_{K'}$ on $\prod_{\pq\not\in Q_0} T_\pq(\psi)$, and let $b_0$ be any element of $B$ that is divisible by $\pq_0$ but not by any other prime of~$B$. Then Theorem \ref{main_theorem} implies that $\rho^\psi_{\text{ad}}( G_{K'}^{\geom})$ is commensurable to $\prod_{\pq\not\in Q_0} D^1_\pq$ and $\rho^\psi_{\text{ad}}(G_{K'})$ is commensurable to $\overline{\langle b_0\rangle} \cdot \prod_{\pq\not\in Q_0} D^1_\pq$.
Viewing $b_0$ as an element of~$C'$, Proposition \ref{uniqueprime} implies that $b_0$ is divisible by~$\pp'_0$ but not by any other prime of~$C'$. The same argument as in Section \ref{55} for $a'$ and~$a_0$ shows here that some positive power of $b_0$ is equal to some positive power of~$c_0$. Thus $\rho^\psi_{\text{ad}}(G_{K'})$ is commensurable to $\overline{\langle c_0\rangle} \cdot \prod_{\pq\not\in Q_0} D^1_\pq$.
By (\ref{TateCompare}) and Lemma \ref{equivariance} the group $G_{K'}$ acts on $\prod_{\pp\not=\pp_0} T_\pp(\phi'|A)$ through the composite of $\rho^\psi_{\text{ad}}$ with the faithful action of\/ $\prod_{\pq\not\in Q_0} D_\pq$. Combining this with (\ref{TateIsog}) and the construction of ${\mathcal D}^1$ we deduce that $\rho_{\text{ad}}( G_{K'}^{\geom})$ is commensurable to ${\mathcal D}^1$ and $\rho_{\text{ad}}( G_{K'})$ is commensurable to $\overline{\langle c_0\rangle} \cdot {\mathcal D}^1$. Since $K'$ is a finite extension of~$K$, the same then follows with $K$ in place of $K'$, as desired.
\end{proof}
|
\section{Introduction}
Quantum phases of matter with exotic types of order have continued to emerge
over the past decades. Examples include
fractional quantum Hall states\cite{TSG8259,L8395}, 1D Haldane
phase\cite{H8364}, chiral spin liquids,\cite{KL8795,WWZ8913} $Z_2$ spin
liquids,\cite{RS9173,W9164,MS0181} non-Abelian fractional quantum Hall
states,\cite{MR9162,W9102,WES8776,RMM0899} quantum orders characterized by
projective symmetry group (PSG),\cite{W0213} topological
insulators\cite{KM0501,BZ0602,KM0502,MB0706,FKM0703,QHZ0824}, etc. Why are
there different orders? What is a general framework to understand all these
seemingly very different phases? How to classify all possible phases and
identify new ones? Much effort has been devoted to these questions, yet the
picture is not complete.
First, we want to emphasize that quantum phase is a property of a class of
Hamiltonians, not of a single Hamiltonian. We call such a class of Hamiltonian
an H-class. For an H-class, of a certain dimension and with possible symmetry
constraints, we ask whether the Hamiltonians in it are separated into different
groups by phase transition and hence form different phases. Two Hamiltonians in
an H-class are in the same/different phase if they can/cannot be connected
\emph{within the H-class} without going through phase transition. We see that
without identifying the class of Hamiltonians under consideration, it is not
meaningful to ask which phase a Hamiltonian belongs to. Two Hamiltonians can
belong to the same/different phases if we embed them in different H-classes.
For an H-class with certain symmetry constraints, one mechanism leading to
distinct phases is symmetry breaking.\cite{L3726,GL5064} Starting from
Hamiltonians with the same symmetry, the ground states of them can have
different symmetries, hence resulting in different phases. This symmetry
breaking mechanism for phases and phase transitions
is well understood.\cite{L3726,LanL58}
However, it has been realized that systems can be in different phases even
without breaking any symmetry. Such phases are often said to be `topological'
or `exotic'. However, the term `topological' in literature actually refers to
two different types of quantum order.
The first type has `intrinsic' topological order. This type of order is defined
for the class of systems without any symmetry constraint, which corresponds to the
original definition of `topological order'.\cite{W8987,W9039} That is, it refers
to quantum phases in an H-class which includes all local Hamiltonians (of a
certain dimension). If we believed that Landau symmetry breaking theory describes
all possibles phases, this whole H-class would belong to the same phase as there
is no symmetry to break. However, in two and three dimensions, there are
actually distinct phases even in the H-class that has no symmetries. These
phases have universal properties stable against any small local perturbation to
the Hamiltonian. To change these universal properties, the system has to go
through a phase transition. Phases in this class include quantum Hall
systems\cite{WN9077}, chiral spin liquids,\cite{KL8795,WWZ8913} $Z_2$ spin
liquids,\cite{RS9173,W9164,MS0181} quantum double model\cite{K0302} and
string-net model\cite{LW0510}. Ground states of these systems have `long-range
entanglement' as discussed in \Ref{CGW1038}.
The `topological' quantum order of the second type is `symmetry protected'. The class of
systems under consideration have certain symmetry and the ground states have
only short-range entanglement\cite{CGW1038}, like in the symmetry breaking
case. However unlike in symmetry breaking phases, the ground states have the same
symmetry as the Hamiltonian and, even so, the ground states can be in different
phases. This quantum order is protected by symmetry; as according to the discussion
in \Ref{CGW1038}, if the symmetry constraint on the class of systems is
removed, all short-range entangled states belong to the same phase. Only when
symmetry is enforced, can short range entangled states with the same symmetry
belong to different phases. Examples of this type include the Haldane
phase\cite{H8364} and topological
insulators.\cite{KM0501,BZ0602,KM0502,MB0706,FKM0703,QHZ0824}
Phases in these two classes share some similarities. For example, they both are
beyond Landau symmetry breaking theory. Also quantum Hall systems and
topological insulators both have stable gapless edge
states.\cite{W9125,BHT0657,FK0702} However, the latter requires symmetry
protection while the former does not.
Despite the similarities, these two classes of topological phases are
fundamentally different, as we can see from quantities that are sensitive to
long-rang entanglement. For example, `intrinsic' topological order has a
robust ground state degeneracy that depends on the topology of the
space.\cite{W8987,W9039} The ground states with `intrinsic' topological order
also have non-zero topological entanglement entropy\cite{LW0605,KP0604} while
ground states in `symmetry protected' topological phases are short range
entangled and therefore have zeros topological entanglement entropy. Also,
the low energy excitations in `symmetry protected' topological phases do not
have non-trivial anyon statistics, unlike in `intrinsic' topological phases.
\cite{ASW8422,WWZ8913,K0602}
In the following discussion, we will use `topological phase' to refer only to the
first type of phases (\ie `intrinsic' topological phases). For the second type,
we will call them `Symmetry Protected Topological' (SPT) phases, as in
\Ref{GW0931}. Similar to the quantum orders characterized by PSG,\cite{W0213} different
SPT phases are also characterized by the projective representations of the
symmetry group of the Hamiltonian.\cite{CGW1107}
The PSG and projective
representations of a symmetry group can be viewed as a `fractionalization' of
the symmetry. Thus, we may say that different SPT phases are caused by
`symmetry fractionalization'.
Long-rang entanglement, symmetry fractionalization, and symmetry breaking
represent three different mechanisms to separate phases and can be combined to
generate a very rich quantum phase diagram. Fig. \ref{fig:phase} shows a phase
diagram with possible phases generated by these three mechanisms. In order to
identify the kind of quantum order in a system, we first need to know whether
topological orders exist. That is, whether the ground state has long range
entanglement. Next, we need to identify the symmetry of the system (of the
Hamiltonian and allowed perturbation). Then we can find out whether all or part
of the symmetry is spontaneously broken in the ground state. If only part is broken, what is the SPT order due to
the fractionalization of the unbroken symmetry. Combining these data together
gives a general description of a quantum phase. Most of the phases studied
before involve only one of the three mechanisms. Examples where two of
them coexist can be found in \Ref{WWZ8913,KLW0902} which combine long
range entanglement (the intrinsic topological order) and symmetry breaking and
in \Ref{W0213,KLW0834,KW0906,YFQ1070} which combine long range entanglement
and symmetry fractionalization. In fact, the PSG provides a quite
comprehensive framework for symmetry fractionalization in topologically ordered
states.\cite{W0213,KLW0834,KW0906}
\begin{figure}
\begin{center}
\includegraphics[scale=0.4]{phase}
\end{center}
\caption{
(Color online) (a) The possible phases for class of Hamiltonians without any
symmetry. (b) The possible phases for class of Hamiltonians with some
symmetries. Each phase is labeled by the phase separating mechanisms involved.
The shaded regions in (a) and (b) represent the phases with long range
entanglement. SRE stands for short range entanglement, LRE for long range
entanglement, SB for symmetry breaking, SF for symmetry fractionalization.
}
\label{fig:phase}
\end{figure}
Based on this general understanding of quantum phases, we address the following
question in this paper: what quantum phases exist in one-dimensional gapped
spin systems. The systems we consider can have any finite strength finite range
interactions among the spins.
In \Ref{CGW1107}, we gave a partial answer to this question. We first showed
that one-dimensional gapped spin systems do not have non-trivial topological
order. So to understand possible 1D gapped phases, we just need to
understand symmetry fractionalization and symmetry breaking in short-range
entangled states. In other words, quantum phases are only different because
of symmetry breaking and symmetry fractionalization.
In \Ref{CGW1107}, we
then considered symmetry fractionalization, and gave a classification of
possible SPT phases with time reversal, parity and on-site unitary symmetry
respectively. In this paper, we complete the classification by considering SPT
phases of combined time reversal, parity and/or on-site unitary symmetry and
finally incorporate the possibility of symmetry breaking. We find that in 1D
gapped spin system with on-site symmetry of group $G$, the quantum phase are
basically labeled by:\\
(1) the unbroken symmetry subgroup $G'$\\
(2) projective representation of the unbroken part of on-site unitary and anti-unitary symmetry respectively\\
(3) `projective' commutation relation between representations of
unbroken symmetries.\\
Here the projective representation and the `projective' commutation relation
represent the symmetry fractionalization. Parity is not an on-site symmetry and its SPT phases are not characterized by projective representations. The classification involving parity does not fall into the general framework above, but proceeds in a very similar way, as we will show in section \ref{CS}. Actually, (2) and (3) combined gives the projective representation of $G'$ and if parity is present, it should be treated as an anti-unitary $Z_2$ element. Our result is consistent with those obtained by Schuch et al\cite{SPC1032}.
Our discussion is based on the matrix product state
representation\cite{FNW9243, PVW0701} of gapped 1D ground states. The matrix
product formalism allows us to directly deal with interacting systems and its
entangled ground state. In particular, SPT order of a system can be identified directly from the way matrices in the representation transform under symmetries.\cite{PTB1039}
Moreover, symmetry breaking in entangled systems can be represented in a
nice way using matrix product states.\cite{FNW9243,PVW0701,N9665} The traditional understanding of symmetry
breaking in quantum systems actually comes from intuition about classical
systems. For example, in the ferromagnetic phase of classical Ising model, the
spins have to choose between two possible states: either all pointing up or all
pointing down. Both states break the spin flip symmetry of the system.
However, for quantum system, the definition of symmetry breaking becomes a
little tricky. In ferromagnetic phase of quantum Ising model, the ground space
is two fold degenerate with basis states $|\uparrow \uparrow ...\uparrow\>$,
$|\downarrow \downarrow ...\downarrow\>$. Each basis state breaks the spin flip
symmetry of the system. However, quantum systems can exist in any superposition
of the basis states and in fact the superposition $\frac{1}{\sqrt{2}}|\uparrow
\uparrow ...\uparrow\> + \frac{1}{\sqrt{2}}|\downarrow \downarrow
...\downarrow\>$ is symmetric under spin flip. What is meant when a
\emph{quantum} system is said to be in a symmetry breaking phase?
Can we understand symmetry breaking in a quantum system
without relying on its classical picture?
In matrix product representation, the symmetry breaking pattern can be seen
directly from the matrices representing the ground state. If we choose the symmetric ground state in the
ground space and write it in matrix product form, the matrices can be reduced to a
block diagonal `canonical form'. The `canonical form' contains more than one
block if the system is in a symmetry breaking phase. If symmetry is not
broken, it contains only one block.\cite{FNW9243,PVW0701,N9665} Hence, the
`canonical form' of the matrices give a nice illustration of the symmetry
breaking pattern of the system. This relation will be discussed in more detail
in section \ref{SB}.
Our classification is focused on 1D interacting spin systems, however it also applies to 1D
interacting fermion systems as they are related by Jordan-Wigner
transformation. As an application of our classification result, we study
quantum phases(especially SPT phases) in gapped 1D fermion systems. Our
result is consistent with previous studies.\cite{TPB1102,LK1103}
The paper is organized as follows: in section \ref{SPT}, we review the previous
classification results of SPT phases with time reversal, parity and on-site
unitary symmetry respectively. We also introduce notations for matrix product
representation; in section \ref{CS}, we present classification result of SPT
phases with combined time reversal, parity and/or on-site unitary symmetry; in
section \ref{SB}, we incorporate the possibility of symmetry breaking; in
section \ref{fermion}, we apply classification results of spins to the study of phases in 1D fermion
systems; and finally we conclude in section \ref{conclude}.
\section{Review: Matrix product states and SPT classification}
\label{SPT}
In \Ref{CGW1107}, we considered the classification of SPT phases with time
reversal, parity, and on-site unitary symmetry respectively. Instead of
starting from Hamiltonians, we classified 1D gapped ground states which do not
break the symmetry of the system. The set of states under consideration can be
represented as short-range correlated matrix product states and we used the
local unitary equivalence between gapped ground states, which was established
in \Ref{CGW1038}, to classify phases. Here we introduce the matrix product
representation and give a brief review of previous classification result and
how it was achieved.
Matrix product states give an efficient representation of 1D gapped spin states\cite{VC0623,SWV0804} and hence provide a useful tool to deal with strongly interacting systems with many-body entangled ground states. A matrix product states (MPS) is expressed as
\begin{equation}
\label{MPS}
|\phi\rangle = \sum_{i_1,i_2,...,i_N}
\Tr (A_{i_1}A_{i_2}...A_{i_N})|i_1i_2...i_N\rangle
\end{equation}
where $i_k=1...d$ with $d$ being the physical dimension of a spin at each site, $A_{i_k}$'s are $D\times D$ matrices on site $k$ with $D$ being the inner dimension of the MPS. In our previous studies\cite{CGW1107} and also in this paper, we consider states which can be represented with a finite inner dimension $D$ and assume that they represent all possible phases in 1D gapped systems. In our following discussion, we will focus on states represented with site-independent matrices $A_{i}$ and discuss classification of phases with/without translational symmetry. Non-translational invariant systems have in general ground states represented by site dependent matrices. However, site dependence of matrices does not lead to extra features in the phase classification and their discussion involves complicated notation. Therefore, we will not present the analysis based on site dependent MPS. A detailed discussion of site dependent MPS can be found in \Ref{CGW1107} and all results in this paper can be obtained using similar method.
A mathematical construction that will be useful is the double tensor
\begin{equation}
\mathbb{E}_{\alpha\gamma,\beta\chi}=\sum_{i}
A_{i,\alpha\beta} \times (A_{i,\gamma\chi})^*
\end{equation}
of the MPS. $\mathbb{E}$ is useful because it uniquely determines the a matrix product state up to a local change of basis on each site\cite{NC2000,PVW0701}. Therefore, all correlation and entanglement information of the state is contained in $\mathbb{E}$ and can be extracted.
First we identify the set of matrix product states that need to be considered for the classification of SPT phases. The ground state of SPT phases does not break any symmetry of the system and hence is non-degenerate. The unique ground state must be short-range correlated due to the existence of gap, which requires that $\mathbb{E}$ has a non-degenerate largest eigenvalue(set to be $1$)\cite{FNW9243, PVW0701}. This is equivalent to an `injectivity' condition on the matrices $A_{i}$. That is, for large enough $n$, the set of matrices corresponding to physical states on $n$ consecutive sites $\mathcal{A}_I = A_{i_1}...A_{i_n}$($I \equiv i_1...i_n$) span the space of $D \times D$ matrices\cite{PVW0701}. On the other hand, all MPS satisfying this injectivity condition are the unique gapped ground states of a local Hamiltonian which has the same symmetry\cite{FNW9243, PVW0701}. Therefore, we only need to consider states in this set.
The symmetry of the system and hence of the ground state sets a non-trivial transformation condition on the matrices $A_{i}$. With on-site unitary symmetry of group $G$, for example, the $A_{i}$'s transform as\cite{PWS0802,PTB1039}
\begin{equation}
\sum_{j} u(g)_{ij} A_{j} = \alpha(g)R^{-1}(g)A_{i}R(g)
\label{eqn:U}
\end{equation}
where $u(g)$ is a linear representation of $G$ on the physical space, $\alpha(g)$ is a one-dimensional representation of $G$. One important realization from this equation is that, in order to satisfy this equation, $R(g)$ only has to satisfy the multiplication rule of group $G$ up to a phase factor\cite{PTB1039}. That is, $R(g_1g_2)=\omega(g_1,g_2)R(g_1)R(g_2)$, $|\omega(g_1,g_2)|=1$. $\omega(g_1,g_2)$ is called the factor system. $R(g)$ is hence a projective representation of group $G$ and belongs to different equivalence classes labeled by the elements in the second cohomology group of $G$, $\{\omega|\omega \in H^2(G,\mathbb{C})\}$.
We showed in \Ref{CGW1107} that two matrix product states symmetric under $G$ are in the same SPT phase if and only if they are related to $R(g)$ in the same equivalence class $\omega$. For two states with equivalent $R(g)$, we constructed explicitly a smooth path connecting the Hamiltonian for the first state to that for the second state without closing the gap or breaking the symmetry of the system. In this way, we gave a `local unitary transformation' as defined in \Ref{CGW1038} connecting the two state and showed that they are in the same SPT phase. On the other hand, for two states associated with inequivalent $R(g)$, we showed that no `local unitary transformation' could connect them without breaking the symmetry. Therefore, they belong to different SPT phases.
If translation symmetry is required in addition to symmetry $G$, $\alpha(g)$ is also a good quantum number and cannot be changed without breaking translation symmetry. Therefore, SPT phases with on-site unitary symmetry and translation symmetry are labeled by $\{\alpha(g),\omega\}$, where $\alpha(g)$ is a one-dimensional representation of $G$ and $\omega$ is an element in $H^2(G,\mathbb{C})$.
The projective representation can be interpreted in terms of boundary spins. A representative state in the phase labeled by $\{\alpha(g),\omega\}$ can be given as in Fig.\ref{fig:rep_1}. Each box represents one site, containing four spins. The symmetry transformations on the two black spins form projective representations of $G$, belonging to class $\omega$, $\omega'$ respectively. The factor systems of the two classes are related by $\omega(g_1,g_2)\times \omega'(g_1,g_2)=1$, that is $\omega'=\omega^*$. Therefore, the inter-site black pair can form a singlet under symmetry $G$. Suppose that the pair forms a 1D representation $\alpha_1(g)$ of $G$. The on-site white pair also forms a 1D representation $\alpha_2(g)$ of $G$. $\alpha_1(g)\alpha_2(g)=\alpha(g)$. It can be checked that, if written in matrix product representation, the matrices satisfy condition \ref{eqn:U}. Now look at any finite segment of the chain. There are unpaired black spins at each end of the chain, transforming under $G$ as projective representation $\omega$ and $\omega^*$. Different $\omega$ cannot be smoothly mapped to each other if the on-site linear symmetry is maintained. Therefore, by looking at the boundary of a finite chain, we can distinguish different SPT phases with on-site unitary symmetry.
\begin{figure}
\begin{center}
\includegraphics[width=3.2in]{rep_1}
\end{center}
\caption{
(Color online)
Representative states for different SPT phases. Each box represents one site, containing four spins. Every two connected spins form an entangled pair.}
\label{fig:rep_1}
\end{figure}
For projective representations in the same class, we can always choose the phases of $R(g)$ such that $\omega(g_1,g_2)$ is the same. In the following discussion we will always assume that $\omega(g_1,g_2)$ is fixed for each class and the phase of $R(g)$ is chosen accordingly. But this does not fix the phase of $R(g)$ completely. For any 1D linear representation $\tilde{\alpha}(g)$, $\tilde{\alpha}(g)R(g)$ always has the same factor system $\omega(g_1,g_2)$ as $R(g)$. This fact will be useful in our discussion of the next section.
The classifications for SPT phases with time reversal and parity symmetry proceed in a similar way.
For time reversal, the physical symmetry operation is $T=v\otimes v... \otimes vK$, where $K$ is complex conjugation and $v$ is an on-site unitary operation satisfying $vv^*=I$, that is $T^2=I$ on each site. (We showed in \Ref{CGW1107} that if $T^2=-I$ on each site, there are no gapped symmetric phases with translation symmetry. Without translation symmetry, it is equivalent to the $T^2=I$ case.) The symmetry transformation of matrices $A_{i}$ is
\begin{equation}
\sum_{j} v_{ij} A^*_{j} = M^{-1} A_{i} M
\label{eqn:TR}
\end{equation}
where $M$ satisfies $MM^*=\beta(T) I = \pm I$. It can be shown, in a way similar to the on-site unitary case, that states with the same $\beta(T)$ can be connected with local unitary transformations that do not break time reversal symmetry while states with different $\beta(T)$ cannot. Therefore, $\beta(T)$ labels the two SPT phases for time reversal symmetry. We can again understand this result using boundary spins. Time reversal on the boundary spin can be defined as $\hat{T}=MK$. It squares to $\pm I$ depending on $\beta(T)$. Therefore, while time reversal acting on the physical spin at each site always squares to $I$, it can act in two different ways on the boundary spin, hence distinguishing two phases. $\hat{T}=MK$ forms a projective representation of time reversal on the boundary spin, with $\hat{T}^2=\pm I$. The result is unchanged if translational symmetry is required.
For parity symmetry, the physical symmetry operation is $P=w\otimes w...\otimes w P_1$, where $P_1$ is exchange of sites and $w$ is on-site unitary satisfying $w^2=1$. As parity symmetry cannot be established in disordered systems, we will always assume translation invariance when discussing parity. The symmetry transformation of matrices $A_{i}$ is
\begin{equation}
\sum_{j} w_{ij} A^T_{j} = \alpha(P) N^{-1} A_{i} N
\label{eqn:P}
\end{equation}
where $\alpha(P)=\pm 1$ labels parity even/odd and $N^T=\beta(P)N=\pm N$. The four SPT phases are labeled by $\{\alpha(P),\beta(P)\}$. A representative state can again be constructed as in Fig.\ref{fig:rep_1}. The black spins form an entangled pair $(N\otimes I)\sum_i|i\>\otimes |i\>$, $i=1...D$, $D$ being the dimension of $N$. ($|i\>\otimes |j\>$ denotes a product state of two spins in state $|i\>$ and $|j\>$ respectively. This is equivalent to notation $|i\>|j\>$ and $|ij\>$ in different literatures.) The white spins form an entangled pair $|0\>\otimes |1\> + \alpha(P)\beta(P)|1\>\otimes |0\>$. Parity is defined as reflection of the whole chain. It can be checked that if written as matrix product states, the matrices satisfy condition Eq.\ref{eqn:P} and parity of the black pair is determined by $\beta(P)$. Therefore, $\beta(P)$ can be interpreted as even/oddness of parity between sites and $\alpha(P)$ represents total even/oddness of parity of the whole chain.
\section{Classification with combined symmetry}
\label{CS}
In this section we are going to consider the classification of SPT phases with
combined translation, on-site unitary, time reversal, and/or parity symmetry in
1D gapped spin systems. The ground state does not break any of the combined
symmetry and can be described as a short-range correlated matrix product state.
For each combination of symmetries, we are going to list all possible SPT
phases and give a label and representative state for each of them. In the end of
this section, we comment on the general scheme to classify SPT phases with all possible kinds of symmetries in 1D gapped spin systems.
\subsection{Parity + On-site $G$}
Consider a system symmetric under both parity $P=w\otimes w...\otimes w P_1$
and on-site unitary of group $G$ $u(g)\otimes u(g) ... \otimes u(g)$.
Eq.\ref{eqn:U} and \ref{eqn:P} give transformation rules of matrices $A_{i}$
under the two symmetries separately in terms of $\alpha(g)$, $R(g)$,
$\alpha(P)$, $N$. $\alpha(g)$ labels the 1D representation the state forms under $G$, $R(g)\in \omega$ is the projective representation of $G$ on the boundary spin, $\alpha(P)$ labels parity even or odd and $N^T=\beta(P)N=\pm N$ corresponds to parity even/odd between sites.
Moreover, parity and on-site $u(g)$ commute. First, it is easy to see that $P_1$ and $u(g) \otimes u(g)...\otimes u(g)$ commute. Therefore, $P_2=w \otimes w ... \otimes w$ must also commute with $u(g) \otimes u(g)...\otimes u(g)$. WLOG, we will consider the case where $w$ and $u(g)$ commute, $wu(g)=u(g)w$. This leads to certain commutation relation between $N$ and $R(g)$ as shown below.
If we act parity first and on-site symmetry next, the matrices are transformed as
\begin{equation}
\begin{array}{lll}
A_i \xrightarrow{P} A'_i &= & \sum_j w_{ij} A^T_j =\alpha(P)N^{-1}A_iN \\
A'_i \xrightarrow{G} A''_i &= & \sum_j u_{ij}(g) A'_j \\
&=& \alpha(P)N^{-1}\left(\sum_j u_{ij}A_j\right)N \\
&= & \alpha(g)\alpha(P) N^{-1}R^{-1}(g)A_iR(g)N
\end{array}
\end{equation}
Combining the two steps together we find that
\begin{equation}
\sum_{j} \sum_{k} u_{ij}(g) w_{jk} A^T_k = \alpha(g)\alpha(P) N^{-1}R^{-1}(g)A_iR(g)N
\label{eqn:PG}
\end{equation}
If on-site symmetry is acted first and then parity follows, the matrices are transformed as
\begin{equation}
\begin{array}{lll}
A_i \xrightarrow{G} A'_i &=& \sum_j u_{ij}(g) A_j= \alpha(g)R^{-1}(g)A_iR(g) \\
A'_i \xrightarrow{P} A''_i &=& \sum_j w_{ij} (A')^T_j \\
&=& \alpha(g)R^T(g)(\sum_j w_{ij} A^T_j)(R^T)^{-1}(g) \\
&=& \alpha(P)\alpha(g) R^T(g)N^{-1}A_iNR^*(g)
\end{array}
\end{equation}
The combined operation is then
\begin{equation}
\sum_{j} \sum_k w_{ij} u_{jk}(g) A^T_k
= \alpha(P)\alpha(g) R^T(g)N^{-1}A_iNR^*(g)
\label{eqn:GP}
\end{equation}
Because $w$ and $u(g)$ commute, $\sum_{j} u_{ij}(g) w_{jk}=\sum_{j} w_{ij} u_{jk}(g)$. Therefore, the combined operation in Eq. \ref{eqn:PG} and \ref{eqn:GP} should be equivalent.
\begin{equation}
N^{-1}R^{-1}(g)A_iR(g)N = R^T(g)N^{-1}A_iNR^*(g)
\end{equation}
This condition is derived for matrices on each site, $i=1...d$. However, it is easy to verify that it also holds if $n$ consecutive sites are combined together with representing matrices $\mathcal{A}_I=A_{i_1}A_{i_2}...A_{i_n}$.
\begin{equation}
N^{-1}R^{-1}(g)\mathcal{A}_IR(g)N = R^T(g)N^{-1}\mathcal{A}_INR^*(g)
\end{equation}
As $\mathcal{A}_I$ is injective(spans the whole space of $D\times D$ matrices), we find $R(g)NR^T(g)N^{-1} \propto I$. That is
\begin{equation}
N^{-1}R(g)N=e^{i\theta(g)}(R^T)^{-1}(g)=e^{i\theta(g)}R^*(g)
\label{eqn:R_gamma}
\end{equation}
Different $e^{i\theta(g)}$ corresponds to different `projective' commutation relations between parity and on-site unitary. It must satisfy certain conditions. As
\begin{equation}
\begin{array}{lll}
N^{-1}R(g_1g_2)N & = & e^{i\theta(g_1g_2)}R^*(g_1g_2)\\
& = & e^{i\theta(g_1g_2)}\omega^{-1}(g_1,g_2)R^*(g_1)R^*(g_2) \\
N^{-1}R(g_1g_2)N & = & \omega(g_1,g_2)N^{-1}R(g_1)NN^{-1}R(g_2)N \\
& = & \omega(g_1,g_2)e^{i\theta(g_1)}R^*(g_1)e^{i\theta(g_2)}R^*(g_2)
\end{array}
\end{equation}
Therefore
\begin{equation}
e^{i\theta(g_1g_2)}e^{-i\theta(g_1)}e^{-i\theta(g_2)}=\omega^2(g_1,g_2)
\end{equation}
Hence, $\omega^2$ must be trivial. WLOG, assume that the factor systems we have chosen(as discussed in section \ref{SPT}) satisfy $\omega^2=1$ and $e^{i\theta(g)}$ forms a linear representation of $G$, denoted by $\gamma(g)$.
Let us interpret this result.
First we see that the combination of parity with on-site $G$ restricts the projective representation that can be realized on the boundary spin to those $\omega$ that square to identity. This can be clearly seen from the structure of the representative state as in Fig.\ref{fig:rep_1}. Because of on-site symmetry $G$, the left and right black spin in a pair form projective representation in class $\omega$ and $\omega^*$ respectively. If the chain has further reflection symmetry, $\omega^*=\omega$ and therefore $\omega^2=1$. However, if $\omega^2\neq 1$, then $\omega^* \neq \omega$. The chain has a direction and cannot have reflection symmetry.
Different $\gamma(g)$ corresponds to different `projective' commutation relation between parity and on-site $G$ on the boundary spin. For example, suppose $G=Z_2$. Consider the state as in Fig.\ref{fig:rep_1} where each black pair consists of two qubits. Suppose that parity on the black pair is defined as exchanged of the two qubits($P_1$) and unitary operation $P_2=Z\otimes Z$ on the two qubits. $Z_2$ symmetry on the pair can be defined as $Z\otimes Z$ or $X\otimes X$. They both commute with parity. However, if we only look at one end of pair, $Z_2$ either commute or anti-commute with $P_2$. Hence these two cases correspond to two different $\gamma(Z_2)$. Because of this, the two phases cannot be connected without breaking the symmetry group generated by parity and $Z_2$.
However, $\gamma(g)$ can be changed by changing the phase of $R(g)$. Remember that the phase of $R(g)$ is only determined(by fixing $\omega$) up to a 1D representation, $\tilde{\alpha}(g)$. From Eq.\ref{eqn:R_gamma}, we can see that if the phase of $R(g)$ is changed by $\tilde{\alpha}(g)$, $\gamma(g)$ is changed to $\gamma(g)/\tilde{\alpha}^2(g)$. Therefore, $\gamma(g)$ and $\gamma'(g)$ which differ by the square of another 1D representation $\tilde{\alpha}(g)$ are equivalent. On the other hand, for a fixed $\omega$($\omega^2=1$) any $\gamma(g)$ can be realized as the `projective' commutation relation, as we will show in Appendix \ref{append_B}.
Therefore, with commuting parity and on-site unitary symmetry, SPT phases in 1D spin chain can be classified by the following data:
\begin{enumerate}
\item $\alpha(P)$, parity even/odd;
\item $\beta(P)$, parity even/odd between sites;
\item $\alpha(g)$, 1D representation of $G$;
\item $\omega$, projective representation of $G$ on boundary spin, $\omega^2=1$;
\item $\gamma(g) \in \mathcal{G}/\mathcal{G}_2$, 1D representation of $G$ related to commutation relation between parity and on-site $G$, where $\mathcal{G}$ is the group of 1D representation of $G$, $\mathcal{G}_2$ is the group of 1D representation squared of $G$.
\end{enumerate}
Following the method used in Appendix G of \Ref{CGW1107}, we can show that
states symmetric under parity and on-site $G$ are in the same SPT phase if and
only if they are labeled by the same set of data as given above. We will not
repeat the proof here.
\begin{figure}
\begin{center}
\includegraphics[width=3.2in]{rep_2}
\end{center}
\caption{
(Color online)
Representative states for different SPT phases. Each box represents one site, containing six spins. Every two connected spins form an entangled pair.}
\label{fig:rep_2}
\end{figure}
A representative state for each phase labeled by $\alpha(P)$, $\beta(P)$,
$\alpha(g)$, $\omega$ and $\gamma(g)$ can be constructed as in
Fig.\ref{fig:rep_2}. We will describe the state of each pair and how it
transforms under symmetry operations. The parameters describing the state will
then be related to the phase labels.
Each entangled pair is invariant under $P$ and on-site $G$. First, the on-site
white pair forms 1D representations $\eta(P)$ and $\eta(g)$ of parity and $G$.
For projective representation class $\omega$ that satisfies $\omega^2=1$ and
any 1D representation $\lambda(g)$, we show in appendix \ref{append_A} and
\ref{append_B} that there always exist projective representation $R(g) \in
\omega$ and symmetric matrix $N$($N^T=N$) such that
$N^{-1}R(g)N=\lambda(g)R^*(g)$. Suppose that $R(g)$ and $N$ are $D$-dimensional
matrices, choose the inter-site black pair to be composed of two
$D$-dimensional spins. Define parity on this pair to be exchange of two spins
and define on-site symmetry to be $R(g)\otimes R(g)$. If the state of the black
pair is chosen to be $(N\otimes I) \sum_i|i\rangle\otimes |i\rangle$, $i=1,...,D$, it
is easy to check that it is parity even, forms 1D representation $\lambda(g)$
for on-site $G$ and contains projective representation $\omega$ at each end.
Finally, define the state of the inter-site grey pair to be
$|0\>\otimes |1\>+\rho(P)|1\>\otimes |0\>$, $\rho(P)=\pm 1$. Parity acts on it as exchange of
spins. On-site $G$ acts trivially on it. The 1D spin state constructed in this
way is symmetric under parity and on-site unitary $G$ and belongs to the SPT
phase labeled by $\alpha(P)=\eta(P)\rho(P)$, $\beta(P)=\rho(P)$,
$\alpha(g)=\eta(g)\lambda(g)$, $\omega$, and $\gamma(g)=\lambda(g)$.
Finally, we consider some specific case.
\begin{enumerate}
\item For translation+parity+$SO(3)$, there are $2\times 2\times 1\times 2\times 1=8$ types of phases
\item For translation+parity+$D_2$, there are $2\times 2\times 4\times 2\times 4=128$ types of phases
\end{enumerate}
\subsection{Time reversal + On-site $G$}
Now consider a 1D spin system symmetric under both time reversal $T=v\otimes v ...\otimes v K$ and on-site unitary $u(g)\otimes u(g)...\otimes u(g)$. On each site, $T^2=vv^*=I$ and $u(g)$ forms a linear representation of $G$. Eq. \ref{eqn:U} and \ref{eqn:TR} are satisfied due to the two symmetries separately with some choice of $\alpha(g)$, $R(g)$, and $M$. $\alpha(g)$ labels the 1D representation the state forms under $G$, $R(g)\in \omega$ is the projective representation of $G$ on the boundary spin, and $\hat{T}=MK$ is the time reversal operator on the boundary spin which squares to $\beta(T) I=\pm I$. Note that while $\hat{T}^2=\pm I$ on the boundary spin, $T$ always square to $I$ on the physical spin at each site.
Moreover, the two symmetries commute, i.e. $u(g)v=vu^*(g)$. This leads to non-trivial relations between $R(g)$ in Eq.\ref{eqn:U} and $M$ in Eq.\ref{eqn:TR}. In particular, suppose we act $G$ first followed by $T$, the matrices $A_{i}$ transform as,
\begin{equation}
\begin{array}{lll}
A_i \xrightarrow{G} A'_i &=& \sum_j u_{ij}(g)A_j = \alpha(g)R^{-1}(g)A_iR(g) \\
A'_i \xrightarrow{T} A''_i &=& \sum_j v_{ij} (A')^*_j \\
&=& \alpha^*(g) (R^*)^{-1}(g)(\sum_j v_{ij} A^*_j)R^*(g) \\
&=& \alpha^*(g) (R^*)^{-1}(g)M^{-1}A_iMR^*(g)
\end{array}
\end{equation}
Acting $T$ first followed by $G$ gives
\begin{equation}
\begin{array}{lll}
A_i \xrightarrow{T} A'_i &=& \sum_j v_{ij} A^*_j = M^{-1}A_iM \\
A'_i \xrightarrow{G} A''_i &=& \sum_j u_{ij}(g) A'_j \\
&=& M^{-1}(\sum_j u_{ij}(g) A_j)M \\
&=& \alpha(g) M^{-1}R^{-1}(g)A_i R(g)M
\end{array}
\end{equation}
Because $u(g)v=vu^*(g)$, the previous two transformations should be equivalent. That is
\begin{equation}
\begin{array}{l}
\alpha^*(g) (R^*)^{-1}(g)M^{-1}A_iMR^*(g) = \\
\alpha(g) M^{-1}R^{-1}(g)A_i R(g)M
\end{array}
\label{UT_equiv}
\end{equation}
Denote $Q=MR^*M^{-1}R^{-1}$. It follows that $A_i=\alpha^2(g)QA_iQ^{-1}$. Suppose that the MPS is injective with blocks larger than $n$ sites, hence $\mathcal{A}_I=A_{i_1}A_{i_2}...A_{i_n}$ satisfies $\mathcal{A}_I=\alpha^{2n}(g)Q\mathcal{A}_IQ^{-1}$. But $\mathcal{A}_I$ spans the whole space of $D\times D$ matrices, therefore $Q\propto I$. That is
\begin{equation}
M^{-1}R(g)M=e^{i\theta(g)}R^*(g)
\end{equation}
Moreover, it follows from Eq.\ref{UT_equiv} that $\alpha^2(g)=1$. That is, the 1D representation of $G$ must have order $2$.
Similar to the parity + on-site $G$ case, we find, $\omega^2=1$, and $e^{i\theta(g)}$ form a 1D representation, denoted by $\gamma(g)$. Two $\gamma(g)$'s that differ by the square of a third 1D representation are equivalent.
Therefore, different phases with translation,
time reversal, and on-site $G$ symmetries are labeled by
\begin{enumerate}
\item $\beta(T)$, $\hat{T}^2=\pm I$ on the boundary spin
\item $\alpha(g)$, 1D representation of $G$, $\alpha^2(g)=1$.
\item $\omega$, projective representation of $G$ on boundary spin, $\omega^2=1$.
\item $\gamma(g) \in \mathcal{G}/\mathcal{G}_2$,1D representation of $G$ related to commutation relation between time reversal and on-site $G$, where $\mathcal{G}$ is the group of 1D representation of $G$, $\mathcal{G}_2$ is the group of 1D representation squared of $G$.
\end{enumerate}
Representative states are again given by Fig.\ref{fig:rep_2}. On-site white
pair forms 1D representation $\eta(g)$ for $G$ and is invariant under $T$.
Similar to the parity+on-site symmetry case, we can show that for any
$\omega$($\omega^2=1$) and 1D representation $\lambda(g)$ there exist
$D$-dimensional projective representation $R(g) \in \omega$ and matrix $M$ such
that $MM^*= I$ and $M^{-1}R(g)M=\lambda(g)R^*(g)$. Choose the inter-site black
pair to be composed of two $D$-dimensional spins. Define time reversal on this
pair to be $(M\otimes M)K$ and define on-site symmetry to be $R(g)\otimes
R(g)$. If the state of the black pair is chosen to be $(M\otimes I)
\sum_i|i\rangle\otimes |i\rangle$, $i=1,...,D$, it is easy to check that it is
invariant under time reversal and forms 1D representation $\lambda(g)$ for $G$
and contains projective representation $\omega$ at each end. Finally, define
the state of the inter-site grey pair to be $|0\>\otimes|1\>+\rho(T)|1\>\otimes|0\>$,
$\rho(T)=\pm 1$. Time reversal acts on it as
$(|0\>\<1|+\rho(T)|1\>\<0|)\otimes(|0\>\<1|+\rho(T)|1\>\<0|)K$. On-site $G$
acts trivially on it. The 1D spin state constructed as this is symmetric under
time reversal and on-site unitary $G$ and belongs to the SPT phase labeled by
$\beta(T)=\rho(T)$, $\alpha(g)=\eta(g)\lambda(g)$, $\omega$, and
$\gamma(g)=\lambda(g)$.
Applying the general classification result to specific cases we find
\begin{enumerate}
\item For Translation+$T+SO(3)$, there are $2\times 1\times 2\times 1=4$ types of phases
\item For Translation+$T+D_2$, there are $2\times 4\times 2\times 4=64$ types of phases
\end{enumerate}
If translation symmetry is not required,
different phases with time reversal and on-site $G$ symmetries are labeled by
\begin{enumerate}
\item $\beta(T)$, time reversal even/odd on the boundary spin
\item $\omega$, projective representation of $G$ on boundary spin, $\omega^2=1$.
\item $\gamma(g) \in \mathcal{G}/\mathcal{G}_2$,1D representation of $G$ related to commutation relation between time reversal and on-site $G$, where $\mathcal{G}$ is the group of 1D representation of $G$, $\mathcal{G}_2$ is the group of 1D representation squared of $G$.
\end{enumerate}
$\alpha(g)$ can no longer be used to distinguish
different phases. We find
\begin{enumerate}
\item For $T+SO(3)$, there are $2\times 2\times 1=4$ types of phases
\item For $T+D_2$, there are $2\times 2\times 4=16$ types of phases
\end{enumerate}
\subsection{Parity + Time reversal}
When parity is combined with time reversal, what SPT phases exist in 1D gapped spin systems? First we realize that due to parity and time reversal separately, different SPT phases exist labeled by different $\alpha(P)$, $\beta(P)$, $\beta(T)$ as defined in Eq.\ref{eqn:TR} and \ref{eqn:P}. $\alpha(P)$ labels parity even or odd, $\beta(P)$ labels parity even/odd between sites, and $\hat{T}^2=\beta(T) I$ on the boundary spin.
Does the commutation relation between parity and time reversal give more phases? The combined operation of parity first and time reversal next gives
\begin{equation}
\begin{array}{lll}
A_i \xrightarrow{P} A'_i &=& \sum_j w_{ij} A^T_j = \alpha(P)N^{-1}A_iN \\
A'_i \xrightarrow{T} A''_i &=& \sum_j v_{ij} (A')^*_j \\
&=& \alpha(P) (N^{-1})^*(\sum_j v_{ij}A^*_j)N^* \\
&=& \alpha(P) (N^{-1})^*M^{-1}A_iMN^*
\end{array}
\end{equation}
and the operation with time reversal first and parity next gives
\begin{equation}
\begin{array}{lll}
A_i \xrightarrow{T} A'_i &=& \sum_{j} v_{ij} A^*_j=M^{-1}A_iM \\
A'_i \xrightarrow{P} A''_i &=& \sum_j w_{ij} (A')^T_j \\
&=& M^T(\sum_j w_{ij} A^T_j)(M^T)^{-1} \\
&=& \alpha(P) M^TN^{-1}A_iN(M^T)^{-1}
\end{array}
\end{equation}
As parity and time reversal commute, $wv=vw^*$. The above
two operations should be equivalent.
\begin{equation}
(N^{-1})^*M^{-1}A_iMN^* = M^TN^{-1}A_iN(M^T)^{-1}
\end{equation}
As $A_i$ is injective, $MN^*M^TN^{-1} \propto I$. That is
\begin{equation}
MN^{\dagger}MN^{\dagger} = e^{i\theta} I
\end{equation}
But $e^{i\theta}$ can be set to be $1$ by changing the phase of $M$ or $N$, therefore, the commutation relation does not lead to more distinct phases.
There are hence eight SPT phases with both parity and time reversal symmetry, labeled by
\begin{enumerate}
\item $\alpha(P)$, parity even/odd;
\item $\beta(P)$, parity even/odd between sites;
\item $\beta(T)$, $\hat{T}^2=\pm I$ on boundary spins.
\end{enumerate}
The representative states of each phase can be given as in Fig.\ref{fig:rep_2}. Each pair of spins forms a 1D representation of parity and time reversal. The on-site white pair is in the state $|0\>\otimes|1\> +\eta(P)|1\>\otimes|0\>$ with $\eta(P)=\pm 1$. Parity on this pair is defined as exchange of spins and time reversal as $K$. Therefore, this pair has parity $\eta(P)$ and is invariant under $T$. The inter-site black pair is in the state $|0\>\otimes|1\> +\lambda|1\>\otimes|0\>$ with $\lambda=\pm 1$. Parity acts on it as exchange of spins and time reversal as $(|0\>\<1|+\lambda|1\>\<0|)\otimes (|0\>\<1|+\lambda|1\>\<0|) K$. This pair therefore has parity $\lambda(P)=\lambda$ and is invariant under time reversal. Time reversal on one of the spins square to $\lambda(T) I=\lambda I$. Finally, the inter-site grey pair is in state $|0\>\otimes|1\> +\rho(P)|1\>\otimes|0\>$ with $\rho(P)=\pm 1$. Parity on this pair is defined as exchange of spins and time reversal as $K$. Therefore this pair has parity $\rho(P)$ and is invariant under $T$. Time reversal on one of the spins square to $I$. This state is in the SPT phase labeled by $\alpha(P)=\eta(P)\lambda(P)\rho(P)$, $\beta(P)=\lambda(P)\rho(P)$, $\beta(T)=\lambda(T)$.
\subsection{Parity + Time reversal + On-site $G$}
Finally we put parity, time reversal and on-site unitary symmetry together and ask how many SPT phases exist if the ground state does not break any of the symmetries. From Eq. \ref{eqn:U}, \ref{eqn:TR}, \ref{eqn:P}, we know that due to the three symmetries separately, states with different $\alpha(g)$, $\omega$, $\alpha(P)$, $\beta(P)$, $\beta(T)$ belong to different SPT phases. $\alpha(g)$ labels the 1D representation the state forms under $G$, $\omega$ is the projective representation of $G$ on the boundary spin, $\alpha(P)$ labels parity even or odd, $\beta(P)$ labels parity even/odd between sites, and $\hat{T}^2=\beta(T) I$ on the boundary spin.
Moreover, the commutation relation between parity, time reversal and on-site $G$ yields further conditions.
The commutation relation between parity and on-site $G$ constrains $\omega^2=1$ and gives,
\begin{equation}
N^{-1}R(g)N=\gamma(g)R^*(g)
\end{equation}
$\gamma(g)$ is a 1D representation of $G$. $\gamma_1(g)$ and $\gamma_2(g)$ correspond to different SPT phases if and only if they are not related by the square of a third 1D representation.
The commutation relation between time reversal and on-site $G$ constrains $\alpha^2(g)=1$ and gives,
\begin{equation}
M^{-1}R(g)M=\gamma'(g)R^*(g)
\end{equation}
$\gamma'(g)$ is a 1D representation of $G$. $\gamma'_1(g)$ and $\gamma'_2(g)$ correspond to different SPT phases if and only if they are not related by the square of a third 1D representation.
The commutation relation between time reversal and parity gives,
\begin{equation}
MN^{\dagger}MN^{\dagger}\propto I
\end{equation}
which is equivalent to, because $N^T=\pm N$ and $M^T=\pm M$,
\begin{equation}
MN^*\propto NM^*
\end{equation}
Therefore, $MN^*$ and $NM^*$ conjugating $R(g)$ should give the same result
\begin{equation}
\begin{array}{lll}
(MN^*)R(g)(MN^*)^{-1} &=& \gamma(g)MR^*(g)M^{-1} \\
&=& \gamma(g)/\gamma'(g)R(g)
\end{array}
\end{equation}
On the other hand,
\begin{equation}
\begin{array}{lll}
(NM^*)R(g)(NM^*)^{-1} &=& \gamma'(g)NR^*(g)N^{-1} \\
&=& \gamma'(g)/\gamma(g)R(g)
\end{array}
\end{equation}
As $R(g)$ is nonzero, $\gamma'(g)=\pm \gamma(g)$.
$\gamma(g)$ and $\gamma'(g)$ are hence related by a 1D representation $\chi(g)$ which squares to $1$. As shown in Appendix \ref{append_C}, for fixed $\omega$, if $N$ and $R(g)$ exist that satisfy $N^{-1}R(g)N=\gamma(g)R^*(g)$, then any choice of $\gamma'(g)=\chi(g)\gamma(g)$($\chi^2(g)=1$) can be realized. The freedom in $\chi(g)$ is $\mathcal{G}/\mathcal{G}_2$. Considering the degree of freedom in choosing $\gamma(g)$, the total freedom in $\{\gamma(g),\gamma'(g)\}$ is $(\mathcal{G}/\mathcal{G}_2)\times (\mathcal{G}/\mathcal{G}_2)$.
The SPT phases with parity, time reversal and on-site unitary symmetries are labeled by
\begin{enumerate}
\item
$\alpha(g)$, 1D representation of $G$, $\alpha^2(g)=1$;
\item
$\omega$, projective representation of $G$ on boundary spin, $\omega^2=1$;
\item
$\alpha(P)$, parity even/odd;
\item
$\beta(P)$, parity even/odd between sites;
\item
$\beta(T)$, $\hat{T}^2=\pm I$ on the boundary spin;
\item
$\{\gamma(g),\gamma'{g}\} \in (\mathcal{G}/\mathcal{G}_2)\times (\mathcal{G}/\mathcal{G}_2)$, 1D representations of $G$, related to commutation relation between time reversal, parity and on-site $G$.
\end{enumerate}
Representative states can be constructed as in Fig.\ref{fig:rep_2}.
Each pair is invariant(up to phase) under $G$, $T$ and $P$.
White pair: forms a 1D representation $\eta(g)$ of $G$, $\eta(P)$ of $P$ and is invariant under time reversal.
Black pair: $G$ acts non-trivially on it as $R(g)\otimes R(g)$. $R(g)$ is a $D$-dimensional projective representation and belongs to class $\omega$. According to appendices \ref{append_A}, \ref{append_B} and \ref{append_C}, for any 1D representation $\lambda(g)$ and order $2$ 1D representation $\chi(g)$, we can find matrices $N$ and $M$ such that
$N^T= N$, $N^{-1}R(g)N = \lambda(g) R^*(g)$, $MM^*=I$, $M^{-1}R(g)M=\lambda'(g)R^*(g)=\chi(g)\lambda(g)R^*(g)$, $MN^*=NM^*$.
Now set the state of this pair to be $N\sum_i |i\rangle\otimes|i\rangle$, where $i=1...D$. Define parity as exchange of sites and time reversal as $(M\otimes M)K$. It can be checked that the state forms a 1D representation $\lambda(g)$ for $G$, has even parity and is invariant under time reversal. Time reversal squares to $I$ at each end.
Grey pair: $G$ acts trivially on it. The pair is in state $|0\>\otimes|1\> +\rho(P)|1\>\otimes|0\>$ with $\rho(P)=\pm 1$. Parity on this pair is defined as exchange of spins and time reversal as $(Y\otimes Y)^{(\rho(T)+1)/2}K$ with $\rho(T)=\pm 1$. Therefore this pair has parity $\rho(P)$ and is invariant under $T$. Time reversal on one of the spins square to $\rho(T)I$.
This state is representative of the SPT phase labeled by $\alpha(g)=\eta(g)\lambda(g)$, $\omega$, $\alpha(P)=\eta(P)\rho(P)$, $\beta(P)=\rho(P)$, $\beta(T)=\rho(T)$, $\gamma(g)=\lambda(g)$, $\gamma'(g)=\lambda'(g)$.
When $G=SO(3)$ or $G=D_2$, the classification result gives:
\begin{enumerate}
\item For translation+$T+P+SO(3)$, there are
$1\times 2\times 2\times 2\times 2 \times(1\times 1)=16$ types of phases
\item For translation+$T+P+D_2$, there are
$4\times 2\times 2\times 2\times 2\times (4\times 4)=1024$ types of phases
\end{enumerate}
\begin{table}[tb]
\centering
\begin{tabular}{ |c|c| }
\hline
Symmetry of Hamiltonian & Number of Different Phases \\
\hline
\hline
None & 1 \\
\hline
$SO(3)$& $2$ \\
\hline
$D_2$& $2$ \\
\hline
$T$ & 2 \\
\hline
$SO(3)+ T$& 4 \\
\hline
$D_2+T$& 16 \\
\hline
\hline
Trans. +$U(1)$& $\infty$ \\
\hline
Trans. +$SO(3)$& $2$ \\
\hline
Trans. +$D_2$& $4\times 2=8$ \\
\hline
\hline
Trans. + $P$ & 4 \\
\hline
Trans. + $T$ & 2 \\
\hline
Trans. + $P+T$ & 8 \\
\hline
\hline
Trans. +$SO(3)+ P$& 8 \\
\hline
Trans. +$D_2+P$& 128 \\
\hline
Trans. +$SO(3)+ T$& 4 \\
\hline
Trans. +$D_2+T$& 64 \\
\hline
Trans. +$SO(3)+ P+T $& 16 \\
\hline
Trans. +$D_2+P+T$ & 1024 \\
\hline
\end{tabular}
\caption{Numbers of different 1D gapped quantum phases
that do not break any symmetry. $T$ stands for time reversal, $P$ stands for parity, and Trans. stands for translational symmetry.
}
\label{table}
\end{table}
In table \ref{table}, we summarize the results obtained above and in
\Ref{CGW1107}.
\subsection{General classification for SPT phases}
Besides the cases discussed above, it is possible to have other types of symmetries in 1D spin systems. For example, there could be systems where time reversal and parity are not preserved individually but the combined action of them together defines a symmetry of the system. The general rule for classifying SPT phases under any symmetry is to classify all the projective representations of the total symmetry group, where on-site unitary symmetries should be represented with unitary matrices, on-site anti-unitary symmetries should be represented with anti-unitary matrices, and parity should be represented with anti-unitary matrices. Moreover, if translational symmetry is present, another independent label for SPT phases exists which corresponds to different 1D representations of the total symmetry group. In calculating this label, the representation of the total symmetry group is slightly different from the one for calculating projective representations. In particular, parity should be represented unitarily, i.e. as a complex number, while on-site unitary/anti-unitary symmetries should still be represented unitarily/anti-unitarily.
\section{Classification with symmetry breaking}
\label{SB}
In \Ref{CGW1107} and previous sections we have only considered 1D gapped phases whose ground state does not break any symmetry and hence is non-degenerate. These SPT phases correspond to one section(labeled `SF' in Fig.\ref{fig:phase}) in the phase diagram for short range entangled states. Of course apart from SPT phases, there are symmetry breaking phases. It is also possible to have phases where the symmetry is only partly broken and the non-broken symmetry protects non-trivial quantum order. In this section, we combine symmetry breaking with symmetry protection and complete the classification for gapped phases in 1D spin systems. We find that 1D gapped phases are labeled by (1) the unbroken symmetry subgroup (2) SPT order under the unbroken subgroup. This result is the same as that in \Ref{SPC1032}.
\subsection{Matrix product representation of symmetry breaking}
Before we try to classify, we need to identify the class of systems and their gapped ground states that are under consideration. As we briefly discussed in the introduction, while the meaning of symmetry breaking is straight forward in classical system, this concept is more subtle in the quantum setting. A classical system is in a symmetry breaking phase if each possible ground state has lower symmetry than the total system. For example, the classical Ising model has a spin flip symmetry between spin up $|\uparrow\>$ and spin down $|\downarrow\>$ which neither of its ground states $|\uparrow\uparrow...\uparrow\>$ and $|\downarrow\downarrow...\downarrow\>$ have. However, in quantum Ising model $H=\sum_{<i,j>} -\sigma^i_z\sigma^j_z$, the ground space contains not only these two states, but also any superposition of them, including the state $|\uparrow\uparrow...\uparrow\> +|\downarrow\downarrow...\downarrow\>$ which is symmetric under spin flip. This state is called the `cat' state or the GHZ state in quantum information literature. In fact, if we move away from the exactly solvable point by adding symmetry preserving perturbations(such as transverse field $B_x\sum_i \sigma^i_x$) and solve for the ground state at finite system size, we will always get a state symmetric under spin flip. Only in the thermodynamic limit does the ground space become two dimensional. How do we tell then whether the ground states of the system spontaneously break the symmetry?
With matrix product representation, the symmetry breaking pattern can be easily seen from the matrices.\cite{FNW9243,PVW0701,N9665} Suppose that we solved a system with certain symmetry at finite size and found a unique minimum energy state which has the same symmetry. To see whether the system is in symmetry breaking phase, we can write this minimum energy state in matrix product representation. The matrices in the representation can be put into a `canonical' form \cite{PVW0701} which is block diagonal
\begin{equation}
A_i = \begin{bmatrix} A^{(0)}_i & & \\ & A^{(1)}_i & \\ & & \ddots \end{bmatrix}
\label{canon_form}
\end{equation}
where the double tensor for each block $\mathbb{E}^{(k)} =\sum_i A_i^{(k)} \otimes (A_i^{(k)})^*$ has a non-degenerate largest eigenvalue $\lambda_i$. If in the thermodynamic limit, the canonical form contains only one block, this minimum energy state is short range correlated and the system is in a symmetric phase as discussed in \Ref{CGW1107} and the previous section. However, if the canonical form splits into more than one block with equal largest eigenvalue(set to be $1$) when system size goes to infinity, then we say the symmetry of the system is spontaneously broken in the ground states.
The symmetry breaking interpretation of block diagonalization of the canonical form can be understood as follows. Each block of the canonical form $A^{(k)}_i$ represents a short range correlated state $|\psi_k\>$. Note that here by correlation we always mean connected correlation $<O_1O_2>-<O_1><O_2>$. Therefore, the symmetry breaking states $|\uparrow\uparrow...\uparrow\>$ and $|\downarrow\downarrow...\downarrow\>$ both have short range correlation. Two different short range correlated states $|\psi_k\>$ and $|\psi_{k'}\>$ have zero overlap $\<\psi_{k'}|\psi_k\>=0$ and any local observable has zero matrix element between them $\<\psi_{k'}|O|\psi_k\>=0$. The ground state represented by $A_i$ is an equal weight superposition of them $|\psi\>=\sum_k |\psi_k\>$. Actually the totally mixed state $\rho=\sum_k |\psi_k\>\<\psi_k|$ has the same energy as $|\psi\>$ as $\<\psi_{k'}|H|\psi_k\>=0$ for $k' \neq k$. Therefore, the ground space is spanned by all $|\psi_k\>$'s. Consider the operation which permutes $|\psi_k\>$'s. This operation keeps ground space invariant and can be a symmetry of the system. However, each short range correlated ground state is changed under this operation. Therefore, we say that the ground states spontaneously break the symmetry of the system.
This interpretation allows us to study symmetry breaking in 1D gapped systems by studying the block diagonalized canonical form of matrix product states. Actually, it has been shown that for any such state a gapped Hamiltonian can be constructed having the space spanned by all $|\psi_k\>$'s as ground space\cite{N9665}. Therefore, we will focus on finite dimensional matrix product states in block diagonal canonical form for our classification of gapped phases involving symmetry breaking.
\subsection{Classification with combination of symmetry breaking and symmetry fractionalization}
We will consider class of systems with certain symmetry and classify possible phases. For simplicity of notation, we will focus on on-site unitary symmetry. With slight modification, our results also apply to parity and time reversal symmetry and their combination. Suppose that the system has on-site symmetry of group $G$ which acts as $u(g)\otimes u(g)...\otimes u(g)$. It is possible that this symmetry is not broken, totally broken or partly broken in the ground state. In general, suppose that there is a short range correlated ground state $|\psi_0\>$ that is invariant under only a subgroup $G'$ of $G$. Of course, different $G'$'s represent different symmetry breaking patterns and hence lead to different phases. Moreover, $|\psi_0\>$ could have different symmetry protected order under $G'$ which also leads to different phases. In the following we are going to show that these two sets of data: (1) the invariant subgroup $G'$ and (2) the SPT order under $G'$ describe all possible 1D gapped phases. Specifically we are going to show that if two systems symmetric under $G$ have short range correlated ground states $|\psi_0\>$ and $|\t{\psi}_0\>$ which are invariant under the same subgroup $G'$ and $|\psi_0\>$ and $|\t{\psi}_0\>$ have the same symmetry protected order under $G'$, then the two systems are in the same phase. We are going to construct explicitly a path connecting the ground space of the first system to that of the second system without closing gap and breaking the symmetry of the system.
Assume that $|\psi_0\>$ has a $D$-dimensional MPS representation $A^{(0)}_i$ which satisfies
\begin{equation}
\sum_{ij} u(g')_{ij} A^{(0)}_j = \alpha(g')M^{-1}(g') A^{(0)}_i M(g')
\end{equation}
where $g'\in G'$, $\alpha(g')$ is a 1D representation of $G'$ and $M(g')$ form a projective representation of $G'$.
Suppose that $\{h_0,h_1,...,h_m\}$ are representatives from each of the cosets of $G'$ in $G$, $h_0\in G'$. Define $|\psi_k\>=u(h_k)\otimes u(h_k) \otimes ...\otimes u(h_k)|\psi_0\>$, $k=1,...,m$. $|\psi_k\>$ are hence each short-range correlated and orthogonal to each other. The ground space of the system is spanned by $|\psi_k\>$. The MPS representation of $|\psi_k\>$ is then $A^{(k)}_i=\sum_j u(h_k)_{ij} A^{(0)}_j$, which satisfies similar symmetry conditions
\begin{equation}
\sum_{ij} u(h_kg'h_k^{-1})_{ij} A^{(k)}_j = \alpha(g')M^{-1}(g') A^{(k)}_i M(g')
\end{equation}
To represent the whole ground space, put all $A^{(k)}_i$ into a block diagonal form and define
\begin{equation}
A_i = \begin{bmatrix} A^{(0)}_i & & \\ & \ddots & \\ & & A^{(m)}_i \end{bmatrix}
\end{equation}
The state represented by $A_i$ is then the superposition of all $|\psi_k\>$, which is equivalent to the maximally mixed ground state with respect to any local observable.
Under any symmetry operation $g\in G$, $A_i$ changes as
\begin{equation}
\sum_j u(g)_{ij} A_j = P(g) \Theta(g)Q(g)A_iQ^{-1}(g)P^{-1}(g)
\label{A_symm}
\end{equation}
where
\begin{equation}
\Theta(g) = \begin{bmatrix} \alpha{(g'_{g,0})} & & \\ & \ddots & \\ & & \alpha{(g'_{g,m})} \end{bmatrix} \otimes I_n
\end{equation}
\begin{equation}
P(g) = p(g) \otimes I_n
\end{equation}
with $p(g)$ $m\times m$ permutation matrices and form a linear representation of $G$.
\begin{equation}
Q(g) = \begin{bmatrix} M(g'_{g,0}) & & \\ & \ddots & \\ & & M(g'_{g,m}) \end{bmatrix}
\label{Q(g)}
\end{equation}
To classify phases, we first deform the state into a simpler form by using the double tensor.
Define the double tensor for the whole state as
\begin{equation}
\mathbb{E} = \sum_i A_i\otimes A_i^*
\end{equation}
As $A_i=\oplus_k A_i^{(k)}$
\begin{equation}
\mathbb{E} = (\oplus_k \mathbb{E}^{(k)}) \oplus (\oplus_{k \neq k'} \mathbb{E}^{(kk')})
\end{equation}
where $\mathbb{E}^{(k)} =\sum_i A_i^{(k)} \otimes (A_i^{(k)})^*$, $\mathbb{E}^{(kk')} = \sum_i A_i^{(k)} \otimes (A_i^{(k')})^*$. As $A_i^{(k)}$ for different $k$ only differ by a local unitary on the physical index $i$, $\mathbb{E}^{(k)}$ all have the same form with a single non-degenerate largest eigenvalue. WLOG, we set it to be $1$ and denote the corresponding eigensector as $\mathbb{E}^{(k)}_0$. On the other hand, $\<\psi_k'|\psi_k\> = \lim_{n\to \infty} Tr \left(\mathbb{E}^{(kk')}\right)^n =0$, therefore, $\mathbb{E}^{(kk')}$ all have eigenvalues strictly less than $1$. Define
\begin{equation}
\mathbb{E}_0 = \oplus_k \mathbb{E}^{(k)}_0
\end{equation}
$\mathbb{E}_0$ is the eigenvalue $1$ sector of $\mathbb{E}$ and $\mathbb{E}_1=\mathbb{E}-\mathbb{E}_0$ has eigenvalues strictly less than $1$.
The symmetry condition on $A_i$ can be translated to $\mathbb{E}$ as
\begin{equation}
\mathbb{E} = \bar{P}(g) \bar{\Theta}(g) \bar{Q}(g) \mathbb{E} \bar{Q}^{-1}(g) \bar{P}^{-1}(g)
\label{E_symm1}
\end{equation}
where $\bar{P}(g)=P(g)\otimes P(g)$($P(g)$ is real), $\bar{\Theta}(g)=\Theta(g)\otimes \Theta^*(g)$, $\bar{Q}(g)=Q(g)\otimes Q^*(g)$.
Matching the eigenvalue $1$ sector on the two side of Eq. \ref{E_symm1}, it is clear to see that
\begin{equation}
\mathbb{E}_0 = \bar{P}(g) \bar{\Theta}(g) \bar{Q}(g) \mathbb{E}_0 \bar{Q}^{-1}(g) \bar{P}^{-1}(g)
\end{equation}
It follows that,
\begin{equation}
\mathbb{E}_ 1 = \bar{P}(g) \bar{\Theta}(g) \bar{Q}(g) \mathbb{E}_1 \bar{Q}^{-1}(g) \bar{P}^{-1}(g)
\end{equation}
That is, $\mathbb{E}_0$ and $\mathbb{E}_1$ satisfy the symmetry condition separately.
Now define the deformation path of the double tensor analogous to \Ref{CGW1107} as
\begin{equation}
\mathbb{E}(t) = \mathbb{E}_0 + \left(1-\frac{t}{T}\right)\mathbb{E}_1
\end{equation}
We will show that as $t$ increases from $0$ to $T$, this corresponds to a deformation of the ground space to a fixed point form while the system remains gapped and symmetric under $G$.
First, it can be checked that for $0\leq t \leq T$, $\mathbb{E}(t)$ remains a valid double tensor and satisfies symmetry condition Eq.\ref{E_symm1}. Decomposing $\mathbb{E}(t)$ back into matrices, $A_i(t)$ necessarily contains $m$ blocks each with finite correlation length. The fact that the total state remains gapped is proven by \Ref{N9665}. Because two equivalent double tensor can only differ by a unitary transformation on the physical index, the symmetry condition Eq.\ref{E_symm1} for $\mathbb{E}(t)$ gives that there exist unitary transformations $u(g)(t)$ such that $A_i(t)$ transform in the same way as $A_i$ in Eq.\ref{A_symm}. The symmetry operation can be defined continuously for all $t$.
At $t=T$, the state is brought to the fixed point form $\mathbb{E}(T) = \mathbb{E}_0 = \sum_k \mathbb{E}^{(k)}_0$. Each block $k$ represents a dimer state as in Fig.\ref{fig:rep_1} with entangled pairs between neighboring sites supported on dimension $1_k,...,D_k$, $|EP_k\> = \lambda_{i_k} |i_ki_k\>$. For different blocks $k$ and $k'$, $|i_k\> \perp |i_{k'}\>$. The total Hilbert space on one site is $(D\times m)^2$ dimensional. The symmetry operation on one site can then be defined as $u(g)=P(g)Q^*(g)\Theta(g)\otimes P(g)Q(g)$. With continuous deformation that does not close gap or violate symmetry, every gapped ground space is mapped to such a fixed point form. If we can show that two fixed point state with the same unbroken symmetry $G'$ and SPT order under $G'$ are in the same phase, we can complete the classification for combined symmetry breaking and SPT order.
Suppose that $|\psi_0\>$ and $|\t{\psi}_0\>$ are short range correlated fixed point ground states of two systems symmetric under $G$. Symmetry operations are defined as $u(g)$ and $\t u(g)$ respectively. $|\psi_0\>$ and $|\t{\psi}_0\>$ are symmetric under the same subgroup $G'$ and has the same SPT order. As shown in \Ref{CGW1107}, $|\psi_0\>$ and $|\t{\psi}_0\>$ can be mapped to each other with local unitary transformation $W_0$ that preserves $G'$ symmetry. The other short range correlated ground states can be obtained as $|\psi_k\>=u(h_k)|\psi_0\>$ and $|\t{\psi}_k\>=\t u(h_k)|\t{\psi}_0\>$. At fixed point, $|\psi_k\>$($|\t{\psi}_k\>$) are supported on orthogonal dimensions for different $k$ and $u(h_k)$ and $\t u(h_k)$ maps between these support spaces. Therefore, we can consistently define local unitary operations mapping between $|\psi_k\>$ and $|\t{\psi}_k\>$ as $W_k=\t u(h_k)W_0u^{\dagger}(h_k)$ and the total operation is $W=\oplus_k W_k$. $W$ as defined is a local unitary transformation symmetric under $G$ that maps between two fixed point gapped ground states with the same unbroken symmetry $G'$ and SPT order under $G'$. Combined with the mapping from a general state to its fixed point form, this completes our proof that 1D gapped phases are labeled by unbroken symmetry $G'$ and SPT order under $G'$.
\section{Application: 1D fermion SPT phases}
\label{fermion}
Although our previous discussions have been focused on spin systems, it actually also applies to fermion systems. Because in 1D, fermion systems and spin systems can be mapped to each other through Jordan Wigner transformation, we can classify fermionic phases by classifying corresponding spin phases. Specifically, for a class of fermion systems with certain symmetry we are going to 1. identify the corresponding class of spin systems by mapping the symmetry to spin 2. classify possible spin phases with this symmetry, including symmetry breaking and symmetry fractionalization 3. map the spin phases back to fermions and identify the fermionic order. In the following we are going to apply this strategy to 1D fermion systems in the following four cases respectively: no symmetry(other than fermion parity), time reversal symmetry for spinless fermions, time reversal symmetry for spin half integer fermions, and $U(1)$ symmetry for fermion number conservation. Our classification result is consistent with previous studies in \Ref{LK1103,TPB1102}. One special property of fermionic system is that it always has a fermionic parity symmetry. That is, the Hamiltonian is a sum of terms composed of even number of creation and annihilation operators. Therefore, the corresponding spin systems we classify always have an on-site $Z_2$ symmetry. Note that this approach can only be applied to systems defined on an open chain. For system with translation symmetry and periodic boundary condition, Jordan Wigner transformation could lead to non-local interactions in the spin system.
\subsection{Fermion Parity Symmetry Only}
For a 1D fermion system with only fermion parity symmetry, how many gapped phases exist?
To answer this question, first we do a Jordan-Wigner transformation and map the fermion system to a spin chain. The fermion parity operator $P_f = \prod (1-2a^{\dagger}_ia_i)$ is mapped to an on-site $Z_2$ operation. On the other hand, any 1D spin system with an on-site $Z_2$ symmetry can always be mapped back to a fermion system with fermion parity symmetry(expansion of local Hilbert space maybe necessary). As the spin Hamiltonian commute with the $Z_2$ symmetry, it can be mapped back to a proper physical fermion Hamiltonian. Therefore, the problem of classifying fermion chains with fermion parity is equivalent to the problem of classifying spin chains with $Z_2$ symmetry.
There are two possibilities in spin chains with $Z_2$ symmetry: (1) the ground state is symmetric under $Z_2$. As $Z_2$ does not have non-trivial projective representation, there is one symmetric phase. (If translational symmetry is required, systems with even number of fermions per site are in a different phase from those with odd number of fermions per site. This difference is somewhat trivial and we will ignore it.) (2) the ground state breaks the $Z_2$ symmetry. The ground state will be two-fold degenerate. Each short-range correlated ground state has no particular symmetry($G'=I$) and they are mapped to each other by the $Z_2$ operation. There is one such symmetry breaking phases. These are the two different phases in spin chains with $Z_2$ symmetry.
This tells us that there are two different phases in fermion chains with only fermion parity symmetry. But what are they? First of all, fermion states cannot break the fermion parity symmetry. All fermion states must have a well-defined parity. Does the spin symmetry breaking phase correspond to a real fermion phase?
The answer is yes and actually the spin symmetry breaking phase corresponds to a $Z_2$ symmetric fermion phase. Suppose that the spin system has two short-range correlated ground states $|\psi_0\>$ and $|\psi_1\>$. All connected correlations between spin operators decay exponentially on these two states. Mapped to fermion systems, $|\psi^f_0\>$ and $|\psi^f_1\>$ are not legitimate states but $|\t{\psi}^f_0\>=|\psi^f_0\> + |\psi^f_1\>$ and $|\t{\psi}^f_1\>=|\psi^f_0\> - |\psi^f_1\>$ are. They have even/odd parity respectively. In spin system, $|\t{\psi}_0\>$ and $|\t{\psi}_1\>$ are not short range correlated states but mapped to fermion system they are. To see this, note that any correlator between bosonic operators on the $|\t{\psi}^f_0\>$ and $|\t{\psi}^f_1\>$are the same as that on $|\psi_0\>$ and $|\psi_1\>$ and hence decay exponentially. Any correlator between fermionic operators on the $|\t{\psi}^f_0\>$ and $|\t{\psi}^f_1\>$ gets mapped to a string operator on the spin state, for example $a^{\dagger}_ia_j$ is mapped to $(X-iY))_iZ_{i+1}...Z_{j-1}(X-iY)_j$, which also decays with separation between $i$ and $j$. Therefore, the symmetry breaking phase in spin chain corresponds to a fermionic phase with symmetric short range correlated ground states.
The degeneracy can be understood as isolated Majorana modes at the two ends of the chain\cite{K0131,K0922}.
On the other hand, the short-range correlated ground state in spin symmetric phase still correspond to short-range correlated fermion state after JW transformation. Therefore, the symmetric and symmetry breaking phases in spin system both correspond to symmetric phases in fermion system. The two fermion phases cannot be connected under any physical fermion perturbation.
\subsection{Fermion Parity and $T^2=1$ Time Reversal}
Now consider the more complicated situation where aside from fermion parity, there is also a time reversal symmetry $\mathcal{T}$. $\mathcal{T}$ acts as an anti-unitary $T=UK$ on each site. In this section we consider the case where $T^2=1$(spinless fermion).
So now the total symmetry for the fermion system is the $Z_2$ fermion parity symmetry $P_f$ and $T^2=1$ time reversal symmetry. $T$ commutes with $P_f$. The on-site symmetry group is a $Z_2\times Z_2$ group and has four elements $G=\{I,T,P_f,TP_f\}$. Mapped to spin system, the symmetry group structure is kept.
The possible gapped phases for a spin system with on-site symmetry $G=\{I,T,P_f,TP_f\}$ include: (1) $G'=G$. Following discussion in section \ref{CS} we find that it has four different projective representations. Examples of the four representations are a.$\{I, K, Z, KZ\}$, b. $\{I,iYK, Z, iYKZ\}$, c. $\{I,iYKZ\otimes I, I\otimes Z, iYKZ\otimes Z\}$ d. $\{I, K, Y, KY\}$. There are hence four different symmetric phases. (If translational symmetry is required, the number is multiplied by $2$ due to $\alpha(Z_2)$) (2) $G'=\{I,P_f\}$ with no non-trivial projective representation, the time reversal symmetry is broken. There is one such phase. (If translational symmetry is required, there are two phases) (3) $G'=\{I,T\}$, with two different projective representations(time reversal squares to $\pm I$ on boundary spin). The $Z_2$ fermion parity is broken. There are two phases in this case. (4) $G'=\{I,TP_f\}$, with two different projective representations. The fermion parity symmetry is again broken. Two different phases. (5) $G'=I$, no projective representation, all symmetries are broken.
Mapped back to fermion systems, fermion parity symmetry is never broken. Instead, the $P_f$ symmetry breaking spin phases are mapped to fermion phases with Majorana boundary mode on the edge as discussed in the previous section. Therefore the above spin phases correspond in the fermion system to: (1) Four different symmetric phases (2) One time reversal symmetry breaking phase. (3) Two symmetric phases with Majorana boundary mode (4) Another two symmetric phases with Majorana boundary mode. (5) One time reversal symmetry breaking phase. (1)(3)(4) contains the eight symmetric phases for time reversal invariant fermion chain with $T^2=1$. This is consistent with previous studies in \Ref{LK1009,TPB1102}.
\subsection{Fermion Parity and $T^2 \neq I$ Time Reversal}
When $T^2 \neq I$, the situation is different. This happens when we take the fermion spin into consideration and for a single particle, time reversal is defined as $e^{i\pi\sigma_y}K$. With half integer spin, $\left(e^{i\pi\sigma_y}K\right)^2=-I$. Note that for every particle the square of time reversal is $-I$, however when we write the system in second quantization as creation and annihilation operator on each site, the time reversal operation defined on each site satisfies $T^2=P_f$. Therefore, the symmetry group on each site is a $Z_4$ group $G=\{I,T,P_f,TP_f\}$. To classify possible phases, we first map everything to spin.
The corresponding spin system has on-site symmetry $G=\{I,T,P_f,TP_f\}$. $T^2=P_f$, $P_f^2=I$. The possible phases are: (1) $G'=G$, with two possible projective representations, one with $T^4=I$, the other with $T^4=-I$. Example for the latter includes $T=(1/\sqrt{2})(X+Y)K$. Therefore, there are two possible symmetric phases. (If translational symmetry is required, there are four phases.) (2) $G'=\{I,P_f\}$, the time reversal symmetry is broken. One phase. (If translational symmetry is required, there are two phases.) (3) $G'=I$, all symmetries are broken. One phase.
Therefore, the fermion system has the following phases: (1) Two symmetric phases (2) One time reversal symmetry breaking phase. (3) One time reversal symmetry breaking phase with Majorana boundary mode. (1) contains the time reversal symmetry protected topological phase. Models in this phase can be constructed by first writing out the spin model in the corresponding spin phase and then mapping it to fermion system with Jordan-Wigner transformation.
\subsection{Fermion Number Conservation}
Consider the case of a gapped fermion system with fixed fermion number. This
corresponds to an on-site $U(1)$ symmetry, $e^{i\theta N}$. Mapped to spins,
the spin chain will have an on-site $U(1)$ symmetry. This symmetry cannot be
broken and $U(1)$ does not have a non-trivial projective representation. One
thing special about $U(1)$ symmetry though, is that it has an infinite family
of 1D representations. If translational symmetry is required, fermion number per
site is a good quantum number and labels different phases. Therefore, mapped back to fermions,
there is an infinite number of phases with different average number of fermions per site.
\section{Conclusion}
\label{conclude}
In this paper, we complete the classification of gapped phases in 1D spin
systems with various symmetries. Based on our classification of symmetry
protected topological phases with on-site unitary, parity or time reversal
symmetry in \Ref{CGW1107}, we give explicit results in this paper for the classification of SPT
phases with combined on-site unitary, parity and/or time reversal symmetry. A general rule is also given for the classification of SPT phases with any symmetry group. Moreover, we considered the classification of phases with possible (partial) symmetry breaking. We find that
1D gapped spin phases with symmetry of group $G$ are basically labeled
by (1) the unbroken symmetry subgroup $G'$, (2) projective representations of $G'$. Note that in calculating projective representations of $G'$, on-site unitary symmetries are represented unitarily while parity and on-site anti-unitary symmetries are represented anti-unitarily. We apply this classification result to interacting 1D fermion
systems, which can be mapped to spin systems with Jordan-Wigner transformation,
and classify possible gapped phases with no symmetry, time reversal symmetry
and also fermion number conservation.
We would like to thank Andreas Ludwig, and Zheng-Xin Liu for very helpful
discussions. This research is supported by NSF Grant No. DMR-1005541.
|
\section{Introduction}
The search for habitable planets has focused on stars similar to the
Sun as it is the sole example we have of a star with a habitable planet, and nuclear
burning provides a long-lived source of energy \citep{Kasting1993,Lunine2008}.
White dwarfs, which are as common as Sun-like stars, may also provide a
source of energy for planets for gigayear (Gyr) durations.
White dwarfs have typical masses 0.4---0.9 $M_\odot$ \citep{Provencal1998}, but have
radii only $\approx$1\% of the
Sun, about the same size as the Earth \citep{Hansen2004}.
The most common white dwarfs have surface temperatures of $\approx$5000 K
which are referred to as ``cool white dwarfs''
since hotter white dwarfs are easier to detect \citep{Hansen2004}.
Cool white dwarfs typically have luminosities of $10^{-4}L_\odot$, so a planet
must orbit at $\approx$0.01 AU to be at a temperature for liquid water to exist
on the surface, the so-called habitable zone \citep{Kasting1993}.
The small size of white dwarfs can cause large transit depths by
Earth-sized or even smaller bodies which could in principle be detectable with
ground-based telescopes \citep{DiStefano2010,Faedi2010,Drake2010}.
Prior to becoming a white dwarf, a Sun-like star expands to a red giant,
engulfing planets within $\approx$1 AU \citep[e.g.,][]{Nordhaus2010}.
Planets present in the white dwarf habitable zone (WDHZ)
must arrive after this phase. This may occur via several
paths \citep{Faedi2010}: planets can form out of gas near the white dwarf,
via the interaction or merger of binary stars \citep{Livio2005},
or by capture or migration from larger distances \citep{Debes2002}.
There are precedents for each of these processes: one neutron star has a
planetary system \citep{Wolszczan1992} which may have been formed from
a disk created after the supernova
\citep{Phinney1993}; pulsars show low-mass stellar
companions being whittled down to planet masses
\citep{Fruchter1988}; and white dwarfs show infrared emission and
atmospheric compositions indicative of close orbiting dust and/or bodies
\citep[e.g.,][]{Zuckerman2010}.
Consequently, short period planets around white dwarfs might plausibly
exist. I sidestep the question of formation, which
has had little theoretical attention, and instead address
the location and duration of habitable
zones around white dwarfs (Section \ref{wdhz}), the detection of planets in
the WDHZ, even if their frequency were much less than 1\%
(Sections \ref{detection} and
\ref{survey}), and how characterization of these planets might proceed
(Section \ref{characterization}).
\section{White dwarf habitable zone} \label{wdhz}
I compute the WDHZ boundary following the procedure in \citet{Selsis2007}.
I determine the white dwarf luminosity and effective temperature,
$T_{\rm eff}$, versus age from white dwarf cooling tracks \citep{Bergeron2001}.
With the luminosity and effective temperature, I compute the range of
distances within which an Earth-like planet could have liquid water
on the surface if it were placed there with an intact atmosphere. I use
computations of the limits of the habitable zone for stars of different
effective temperatures that are based on empirical limits
from our solar system combined with one-dimensional radiative---convective atmospheric
models for Earth-like planets that include water loss at the inner edge and
the maximum CO$_{\rm 2}$ greenhouse effect at the outer edge \citep{Kasting1993}.
Most white dwarfs have masses $M_{\rm WD}=0.6 M_\odot$ and CO interior
composition, for which I plot the WDHZ versus time in Figure 1 as
a blue shaded region. This region shrinks with time as the star
cools. A planet enters at the bottom of Figure 1 and moves vertically
up the figure as its white dwarf host ages, so it starts off too hot for
liquid water, passes through the WDHZ, and then becomes too cold.
The duration a planet spends within the WDHZ, $t_{\rm HZ}$, has a maximum of
8 Gyr at $\approx$0.01 AU. Based on the WDHZ limits, I next define the
``continuously habitable zone'' (CHZ)
as the range of planet orbital distances, $a$, that are habitable for
a minimum duration, $t_{\rm min}$ (Figure 2). I choose a minimum duration of
$t_{\rm min} = 3$ Gyr, which results in a CHZ within $a < 0.02$ AU: a planet
that orbits within this distance spends at least 3 Gyr within the WDHZ.
From 0.4 to 0.9 $M_\odot$ with $t_{\rm min}$=3 Gyr the outer boundary of the
CHZ always falls within 10\% of 0.02 AU for hydrogen and helium atmospheres
(Figure 2). To check the sensitivity to the white dwarf cooling computations,
I have also computed the WDHZ with BASTI models \citep{Salaris2010}, for
which I find a slightly longer $t_{\rm HZ}$. I have also computed
the CHZ for atmospheres with $t_{\rm min}=$ 1 Gyr and
5 Gyr which shifts the outer boundary of the CHZ outward/inward by a factor of
$\approx$1.5/0.7 (Figure 2).
\begin{figure}
\centerline{\psfig{figure=fig1.ps,width=\hsize}}
\caption{WDHZ for $M_{\rm WD}=0.6 M_\odot$ vs.\ white dwarf age and
planet orbital distance. Blue region denotes the WDHZ. Dashed line is
Roche limit for Earth-density planets. Planets to right of
dotted line are in the WDHZ for less than 3 Gyr.
Planet orbital period is indicated on the top axis; and white dwarf effective
temperature on the right axis. Luminosity of the white dwarf at different ages
are indicated on right.}\label{fig01}
\end{figure}
\begin{figure}
\centerline{\psfig{figure=fig2.ps,width=\hsize}}
\caption{CHZ vs.\ white dwarf mass and planet orbital distance.
Green region is the CHZ for $t_{\rm HZ} > t_{\rm min}$=3 Gyr, H-atmosphere.
Left solid line is Roche limit for Earth-density
planets. The other lines show how the CHZ outer boundary changes
for $t_{\rm min}$=1 Gyr (dotted line), for $t_{\rm min}$=5 Gyr (dash-dotted line), or
for an He atmosphere with $t_{\rm min}$= 3 Gyr (dashed line). Horizontal line
indicates the most common white dwarf mass of 0.6 $M_\odot$, plotted in Figure 1.}
\end{figure}
There are several important consequences of the WDHZ and CHZ.
First, the range of white dwarf temperatures in the portion of
the CHZ within the WDHZ is that of cool white dwarfs,
$\approx$3000--9000 K (right hand axis in Figure 1), similar
to the Sun. At the hotter end higher ultraviolet
flux might affect the retention of an atmosphere, these planets would
need to form a secondary atmosphere, as occurred on Earth. Excluding higher
temperature white dwarfs only slightly modifies the CHZ since they spend
little time at high temperature. Cool white dwarfs are photometrically stable
\citep{Fontaine2008},
which is critical for finding planets around them. Second, for white
dwarfs with temperatures $\ga$4500 K, the WDHZ is exterior to the Roche
limit (tidal disruption radius),
$a_{\rm R} = 0.0054 {\rm AU} (\rho_{\rm p}/\rho_\oplus)^{-1/3}
(M_{\rm WD}/0.6 M_\odot)^{1/3}$,
where $\rho_{\rm p,\oplus}$ are the mass densities of the planet and of Earth.
Consequently, the 3 Gyr CHZ lies between $a_{\rm R}<a<0.02$ AU,
indicated with the green region in Figure 2.
Third, the CHZ occurs at white dwarf luminosities of $10^{-4.5}$ to
$10^{-3} L_\odot$,
about 10 magnitudes fainter than the Sun, which sets the minimum telescope
size for detection. Finally, the orbital period of white dwarfs in the CHZ
is $\approx$4--32 hr.
At this period the timescales for tidal circularization and tidal locking
are $\approx$10--1000 years, so rocky planets will be synchronized and
circularized \citep{Heller2011}; the side of the planet near the star will have a permanent
day, while the far side will have a permanent night. The planet will
orbit stably as it cannot raise a tide on the compact white dwarf.
Radiation drag will not cause the orbit to decay,
but magnetic field drag might, depending
on the planet's conductivity \citep{Li1998}.
\section{Detection of planets in the white dwarf habitable zone} \label{detection}
I plot example light curves of planets in the WDHZ in Figure 3 for
sizes within a factor of $\approx$2 of Earth orbiting a white dwarf
near the peak of the white dwarf luminosity function.
The transits last $\approx$2 minutes, and have a maximum depth of
10\%---100\%; these events can be detected at a 100 pc distance with a 1 m
ground-based telescope.
To detect planets in the CHZ one must monitor a sample of white dwarfs for the
duration of the orbital period at 0.02 AU, and planets present in edge-on orbits will be
seen to transit their host stars.
The transit probability is
\begin{equation}
p_{\rm trans}= 1.0\% \left(\frac{R_{\rm p}/R_\oplus+R_{\rm WD}/0.013R_\odot} {a/0.01 {\rm AU}}\right),
\end{equation}
so for every 100 Earths orbiting white dwarfs at 0.01 AU with random orientations,
on average one will be seen to transit.
\begin{figure}
\centerline{\psfig{figure=fig3.ps,width=\hsize}}
\caption{Example light curves of habitable Earth-like planets transiting
a 0.6 $M_\odot$ white dwarf ($T_{\rm eff}=5200$K, $R_{\rm WD}=0.013 R_\odot$,
inclination $= 89^\circ.9$, 10\% linear limb darkening, and $a=0.013$ AU)
with masses of $0.1 M_\oplus$ (red, about Mars mass), $1.0
M_\oplus$ (black, Earth twin), and $10.0 M_\oplus$ (blue, super-Earth).
The ratios of the planet radii, $R_{\rm p}$, to white dwarf radius, $R_{\rm WD}$,
are indicated.}
\end{figure}
I define $\eta_\oplus$ to be the number of planets with $0.1 M_\oplus < M_{\rm p} <
10 M_\oplus$ in the 3 Gyr CHZ ($a<$ 0.02 AU). To measure $\eta_\oplus$ to
an accuracy of $\approx$33\% requires detecting $\approx$9 planets, so one must survey
$\approx$10$^3 \eta_\oplus^{-1}$ white dwarfs.
\section{Survey strategy} \label{survey}
The local density of white dwarfs is $(4.7\pm 0.5) \times 10^{-3}$ pc$^{-3}$
\citep{Harris2006}.
For a survey out to $D_{\rm max} < 200$ pc, the white dwarfs should be nearly
isotropically spaced on the sky at one per $\approx$2$(100 {\rm pc}/D_{\rm max})^{3}$ deg$^2$.
This exceeds the field of view of most telescopes, so each white dwarf must
be separately surveyed for the presence of transiting planets.
I have simulated an all-sky survey with a worldwide network of 1 m aperture
telescopes to monitor the white dwarf CHZ (typically 32 hr, during
which telescopes distributed in longitude follow a single star) following
\citet{Nutzman2008} to compute the telescope sensitivity, including
sky and read noise, and assuming an exposure time of 15 s;
I conservatively expanded the error bars an additional 50\%.
Each white dwarf is given a multi-planet system whose innermost planet is
chosen from a log-normal centered at $2a_{\rm R}$ with width of 0.5 dex, with
subsequent planets packed as closely
as allowed by dynamical stability on a timescale of $10^9$ yr \citep{Zhou2007},
leading to a uniform distribution in $\log{a}$, with a gradual cutoff within
$2a_{\rm R}$. The planet masses are drawn from $dn/dM_{\rm p} \propto M_{\rm p}^{-\alpha}$ from
$10^{-2}M_\oplus$ ($\approx$Moon) to $10^2M_\oplus$ ($\approx$Saturn),
with $\alpha=4/3$ to match
the observed slope measured with the {\it Kepler} satellite \citep{Borucki2011},
as well as that found in the solar system. I assumed Earth-composition
planets with the radius determined from the mass according to
\citet{Seager2007}. Two exposures within transit
with a signal-to-noise of at least 6 constitute a detection; this keeps the
false-positive level to $<$1\% for the entire survey. I then scale the
simulated detection rates with $\eta_\oplus$ to determine the expected
number of detected planets.
To detect 9$\pm$3 planets, for $\eta_\oplus$ = 50\% a survey of 2800 white
dwarfs within 52 pc is required for a total on-sky time of 10 years (for
a single telescope with 31 hours per white dwarf on average). A smaller
planet frequency of $\eta_\oplus$ = 10\% requires $\approx$20,000 white
dwarfs within 100 pc for a total 69 years of telescope time on sky.
For a network of twenty 1 m telescopes distributed
around the globe, such as the Las Cumbres Observatory Global Telescope
\citep[LCOGT;][]{Hidas2008}, or the Whole Earth Telescope
\citep[WET][]{Nather1990}, devoted
to observing white dwarfs at 25\% efficiency (50\% of time at night,
50\% weather loss), the total calendar time required would be 2 years
for a survey of 2800 white dwarfs and 14 years for a survey of 20,000.
If the CHZ is surveyed to only 0.01 AU, this would decrease
the required calendar time to 8.5 months and 5 years, respectively, but would
also decrease the planet yield.
Figure 4 shows the probability distribution for planets detected in 10$^4$ survey
simulations. Each one surveys 20,000 single white dwarfs out to 100 pc
for planets within 0.02 AU.
Of the detected planets, an average of 40\% will be currently within the WDHZ.
Remarkably, the detection probability peaks near planets of the size and
temperature of Earth due to the
coincidence in size of the Earth and white dwarfs, and the coincidence between
the WDHZ at the peak of the white dwarf luminosity function and $2a_{\rm R}$.
This leads to the following biases: (1) large planets have a higher transit
detection probability $\propto R_{\rm p} dn/dR_{\rm p}$, so the number
detected declines if $dn/dR_{\rm p}$ is steeper than $R_{\rm p}^{-1}$;
(2) small planets cause shallower transits for $R_{\rm p} \ll R_{\rm WD}
\approx R_\oplus$, so the detection rate scales as
$\propto R_{\rm p}^6 dn/dR_{\rm p}$ \citep{Pepper2003}, although
the range of luminosities of white dwarfs flattens this decline;
(3) cooler planets have a smaller probability of transit, so fewer are detected as
$\propto T_{\rm p}^4$ for small $T_{\rm p}$ and
a uniform distribution in $\log{a}$; and (4) hotter planets orbit hotter
stars, which are less numerous, $dn/dT_{\rm WD} \propto T_{\rm WD}^{-3.9}$
for large $T_{\rm WD}$ \citep{Hansen2004}. Although these trends should occur
for any volume-limited survey, the
break in the radius detection limit depends on the size of the
telescope and signal-to-noise cutoff that is chosen: larger telescopes
or smaller signal-to-noise cuts will be more sensitive to small radius planets.
Figure 4 is sensitive to the
properties of the planet population: if the inner cutoff, $a_{\rm in}$, is
further/closer, then the peak temperature moves to
cooler/hotter temperatures, $\propto a_{\rm in}^{-1/2}$, and the total number of
planets detected declines/increases as $a_{\rm in}^{-1}$ due to the lower/higher transit
probability. If the planet size distribution is steeper/flatter, then the detected
planet distribution peaks at smaller/larger sizes.
\begin{figure}
\centerline{\psfig{figure=fig4.ps,width=\hsize}}
\caption{Probability density, $d^2n/(dT_{\rm p}d\log{R_{\rm p}})$, of detected planets
vs.\ planet radius, $R_{\rm p}$ (log axis scale), and planet effective temperature, $T_{\rm p}$
(assuming the same albedo as Earth). The contours enclose 25\%, 50\%, and 75\% of all
detected planets; the contour levels are 29\%, 53\%, and 76\% of the peak density.
Earth ($\oplus$), Mars ($\mars$), and Venus ($\venus$) symbols indicate the radii
and effective temperatures of these solar system planets.}
\end{figure}
Prior to such a survey, a nearby sample of cool white dwarfs must be found using measurements
of the reduced proper motion. Ongoing and planned deep astrometric surveys, such as
90 Prime, Skymapper, Pan-STARRS, URAT, and GAIA, should find most cool
white dwarfs out to 100 pc ($V$$<$21) within the decade \citep{Henry2009,
Kalirai2009}. Some of these surveys might also find transits
if the requirement of three epochs in transit is relaxed.
For example, GAIA will observe 200,000 disk white dwarfs 50-100 times each
\citep{Perryman2001}, possibly detecting one epoch in eclipse for $\approx$10\%
of these stars with habitable transiting planets.
For values of $\eta_\oplus$$<$10\%, more stars must be observed to detect
planets, so a better strategy is to observe multiple white dwarfs simultaneously
with a wide-field imager with fast readout, such as the Large Synoptic Survey Telescope \citep[LSST][]{LSST2009}.
The LSST survey is expected to detect $\approx$10$^7$ white dwarfs over half
of the sky at $>$5$\sigma$ to $r$$<$24.5 with $\approx$1000 epochs each and two
15 s exposures per epoch over the duration of the 10 year survey. Since
LSST is a magnitude-limited survey, the white dwarf temperature distribution peaks
at $10^4$K. I have taken the simulated detected
distribution of white dwarfs for LSST \citep{Juric2008},
created simulated light curves for white dwarfs with planets, and added
noise \citep{LSST2009}. I find LSST can detect $>$9 CHZ planets if
$\eta_\oplus$$>$5$\times$10$^{-3}$, where detection requires
that at least three epochs fall within transit with two points each
detected at $>$7$\sigma$. The LSST survey will be biased toward detecting
shorter period ($\propto P^{-4/3}$) and large-size planets that have
yet to enter the WDHZ since their stars are hotter. This could be improved by
either continuously observing some fields for several nights, or taking more exposures
per field, resulting in detection of smaller, cooler planets, thus constraining smaller
$\eta_\oplus$. LSST will identify the white dwarfs with reduced proper motion
measurements as the survey is being carried out.
I estimate that $>$10$^3$ double white
dwarf eclipsing binaries with orbital periods similar to WDHZ planets will be found with
LSST using the BSE population synthesis model for binary stars \citep{Hurley2002} with
parameters taken from observed binaries \citep{Raghavan2010}. I find that about
2.5\% of white dwarfs will have a white dwarf companion with a period in the
range of 8---64 hr which might be mistaken for a transiting CHZ planet if
these are viewed edge-on. Follow-up of planet candidates will be required to
distinguish the two possibilities:
white dwarf binaries will show primary and secondary eclipses of different depths
(if the two white dwarfs differ in
temperature), offset secondary eclipses due to light travel time,
gravitational lensing \citep{Agol2002}, and Doppler modulation, and
an eclipse shape that differs from planetary transits if non-grazing. I
simulated light curves of white dwarf binaries including these effects and find
that in the worst case of two white dwarfs with identical temperatures
these distinguishing features may be detected with 10---100 m
ground-based telescopes for systems out to 50---100 pc, either photometrically
or with radial velocities.
Another concern is grazing eclipses from white dwarf/M dwarf binaries;
these can be identified by the eclipse shapes, spectral energy distribution, and differing secondary eclipse depths.
\section{Planet characterization} \label{characterization}
The parallax and spectrum of a white dwarf yield its mass, luminosity, atmospheric
composition, and radius; then the transit depth gives the planet radius.
The planet's mass cannot be measured from Doppler shifts
due to the featureless spectra of cool white dwarfs,
but may be bracketed by the range
of compositions for planets of a given size \citep{Seager2007}.
The mass might be measured
by observing wavelength dependent absorption, such as Rayleigh scattering
that varies as $8H/R_{\rm p} \ln{\lambda}$ \citep{LecavelierDesEtangs2008},
where $H$ is the atmospheric scale height, causing the transit depth
to vary by $\approx$few millimagnitude over 400---500 nm if Rayleigh scattering dominates
over other sources of opacity. If atmospheric molecular weight can be
estimated, so can the planet mass \citep{MillerRicci2009}.
If two planets transit a white dwarf (about 2.5\% of
the time for mutual inclinations within $5^\circ$ and in a
packed planet system), then transit timing variations might
constrain the planet masses if the orbital period ratio is nearly commensurate
\citep{Holman2005,Agol2005}.
Infrared phase variation of planets \citep{Knutson2007} in the CHZ is
$\approx$0.5\%---2\% at 15---20 $\mu$m; however, I estimate that the {\it James Webb
Space Telescope} ({\it JWST}) cannot detect this for cool white dwarfs due to
telescope noise.
For a hotter white dwarf of $10^4$ K at 100 pc with an Earth-like planet
at 0.01 AU with a day---night contrast of 30\%, the phase variation of 0.1\%
at 7.7 $\mu$m might be detectable at 9$\sigma$ with the MIRI imager on
{\it JWST} \citep{Swinyard2004}.
Several topics require further study. The global climate models for
determining the WDHZ should include fast synchronous rotation, magnetic fields,
varied planet atmospheric composition, radiogenic heating, and white dwarf
cooling. The WDHZ is a necessary but insufficient criteria for habitability.
For example, planets that start hot may not retain their atmospheres, as has
also been argued for planets orbiting M dwarfs \citep{Lissauer2007}; this may
require volatile delivery from more distant bodies in the system
\citep{Jura2010} or planetary outgassing. To retain an atmosphere
might require a larger planet escape velocity, possibly favoring super-Earths
for habitability.
Formation mechanisms must be modeled to help motivate future
surveys. For example, gravitational interactions
of a planet and star with a third companion body may be responsible for
creating hot Jupiters \citep{Fabrycky2007}, which is also promising
for moving
distant planets around white dwarfs to $2a_{\rm R} \approx 0.01$ AU, the tidal
circularization radius \citep{Ford2006}.
It is also possible that tidal
disruption of a planet or a companion star will result in the formation of
a disk which may cool and form planets \citep{Guillochon2010}, out of which
a second generation of planets might form \citep{Menou2001,
Perets2010,Hansen2009}.
The most common white dwarf has $T_{\rm eff}\approx$5000 K, close to that
of the Sun; consequently, inhabitants of a planet in the CHZ will see their
star as a similar angular size and color as we see our Sun.
The orbital and spin period of planets in the CHZ are similar
to a day, causing Coriolis and thermal forces similar to Earth.
The night sides of these planets
will be warmed by advection of heat from their day sides if
a cold-trap is avoided \citep{Merlis2010}.
Transit probabilities of habitable planets are similar for cool white
dwarfs and Sun-like stars, but the white dwarf planets can be found using
ground-based telescopes (e.g., LCOGT, WET, and LSST) at a much less expensive price
than space-based planet-survey telescopes.
\acknowledgments
I acknowledge NSF CAREER grant AST-0645416, and thank KITP
(NSF PHY05-51164) and the Whiteley Center for their hospitality.
I thank Fred Adams, Rory
Barnes, Pierre Bergeron, Tim Brown, Mark Claire, Nick Cowan, Dan
Fabrycky, Eric Gaidos, Rob Gibson, Brad Hansen, Mario Juric, Lisa
Kaltenegger, Piotr Kowalski, Jim Kasting, Vikki Meadows, Enric Palle,
Michael Perryman, David Spiegel, Giovanna Tinetti, and the anonymous
referee for help.
\noindent{\it Note added in proof.} Ren\'e Heller informed me that
\citet{Monteiro2010} discussed the white dwarf habitable zone for known
white dwarf stars, showing a boundary similar to that in Figure 1.
|
\section{Introduction}
\setcounter{equation}{0}\setcounter{theorem}{0}
The study of gradings of Lie algebras and the symmetries of those
gradings is an active research area in recent decades, which are
interesting to both mathematicians and physicians. In physics, Lie
algebras usually play the role as the algebra of infinitesimal
symmetries of a physical system. Knowledge about the gradings of a
Lie algebra will greatly help us to understand better the structure
of the Lie algebra. Study of the symmetries of those gradings offers
a very important tool for describing symmetries in the system of
nonlinear equations connected with contraction of a Lie algebra (see
e.g. \cite{jpt}).
Besides the
famous Cartan decomposition for semisimple Lie algebras, another
well-known example of grading is the grading of $sl(n,\mb C)$ by the
adjoint action of the Pauli group $\Pi_n$ generated by the $n\times
n$ generalized Pauli matrices, which decomposes $sl(n,\mb C)$ into
direct sum of $n^2-1$ one-dimensional subspaces, each of which
consists of semisimple elements.
Let $L$ be a complex simple Lie algebra. Let $\Aut(\L)$ and $\Int(L)$ be respectively the automorphism group and inner automorphism group of $L$, which are
both algebraic groups. A subgroup of $\Aut(\L)$ or $\Int(L)$ is
called \textit{diagonalizable} if it is abelian and consists of
semisimple elements. It is not hard to see that there is a natural
1-1 correspondence between gradings of $L$ and diagonalizable
subgroups of $\Aut(\L)$ (see Section 4). A grading is called
\textit{inner} if the respective diagonalizable subgroup is in
$\Int(\L)$. A grading (resp. inner grading) of $L$ is called fine if it could not be further refined by any other grading (resp. inner grading).
Among the gradings of a Lie algebra, fine (inner) gradings are especially
important. It was shown in \cite{pz} that the fine gradings of
simple Lie algebras correspond to maximal diagonalizable subgroups
(which were called MAD-groups in \cite{pz}) of $\Aut(\L)$. Then fine
inner grading of simple Lie algebras corresponds to maximal
diagonalizable subgroups of $\Int(\L)$. Given a fine (inner) grading
$\Ga$, one can define naturally its Weyl group (see Definition 2.3
of \cite{hg}) to describe its symmetry. Assume $K$ is the maximal
diagonalizable subgroup corresponding to $\Ga$, then one can show
that its Weyl group is isomorphic to the Weyl group of $K$. See Proposition 2.4 and Corollary 2.6 of \cite{hg}.
After many mathematicians and physicians' contribution, the
classification of fine gradings of all the simple Lie algebras are
almost done. For example, it can be found in \cite{e} the
classification of fine gradings of all the classical simple Lie
algebras over an algebraically closed field of characteristic 0.
People have also known a lot about the fine gradings for exceptional
simple Lie algebras, see \cite{k} for a survey of such results. For
the Weyl group of a fine inner grading of a simple Lie algebra $L$,
if the grading is Cartan decomposition (in which case the
corresponding maximal diagonalizable subgroup $K$ of $\Int(L)$ is
just the maximal torus), then it is well-known that the Weyl group
is a finite group generated by reflections; in other cases there is
no general result by far. The next step is to study the case $K$ is
discrete, and people have made some explorations in the case
$L=sl(n,\mb C)$ .
Recall that $PGL(n,\mb C)$ is the inner automorphism group of
$sl(n,\mb C)$. Let us first review the classification of maximal
diagonalizable subgroups of $PGL(n,\mb{C})$, which correspond to
fine inner gradings of $sl(n,\mb C)$. Let $\Pi_n$ be the Pauli group
of $GL(n,\mb{C})$ and $D_n$ be the subgroup of diagonal matrices of
$GL(n,\mb{C})$. Let
$\texttt{P}_n$ and $\texttt{D}_n$ be the respective images of $\Pi_n$ and $D_n$ in $PGL(n,\mb{C})$ under the adjoint
action on $M(n,\mb{C})$. Assume $n=k l_1\cdots l_t$ and each $l_i$
divides $l_{i-1}$. The group $D_k\ot\Pi_{l_1}\ot \cdots
\ot\Pi_{l_t}$ consists of all those elements $A_0\ot A_1\ot\cdots\ot
A_t$ with $A_0\in D_k$ and $A_i\in\Pi_{l_i}$ for $1\le i\le t$. The adjoint action of
$D_k\ot\Pi_{l_1}\ot \cdots \ot\Pi_{l_t}$ on
$$M(k,\mb{C})\ot M(l_1,\mb{C})\ot\cdots\ot M(l_t,\mb{C})\cong M(n,\mb{C})$$ induces the embedding
$$ \texttt{D}_k\times \texttt{P}_{l_1}\times \cdots \times \texttt{P}_{l_t}\hookrightarrow PGL(n,\mb{C}). $$ If we identify
$ \texttt{D}_k\times \texttt{P}_{l_1}\times \cdots \times
\texttt{P}_{l_t}$ with its image, then it was shown by Havlicek,
Patera and Pelantova in Theorem 3.2 of \cite{hpp} that any maximal
diagonalizable subgroup $K$ of $PGL(n,\mb{C})$ is conjugate
to one and only one of the $ \texttt{D}_k\times \texttt{P}_{l_1}\times \cdots \times \texttt{P}_{l_t}$.
Let $K$ be a discrete maximal diagonalizable subgroup of $PGL(n,\mb
C)$. Then $K\cong\texttt{P}_{l_1}\times\texttt{P}_{l_2}\times \cdots
\times \texttt{P}_{l_t}$ where $n=l_1 l_2\cdots l_t$ and each $l_i$
divides $l_{i-1}$. Then the fine grading induced by $K$ also
decomposes $sl(n,\mb C)$ into $n^2-1$ one-dimensional subspaces,
each of which consists of semisimple elements. We will show in
Section 5 that there
is a nonsingular anti-symmetric pairing $<,>$ on $K$,
such that
$(K,<>)$ is a nonsingular symplectic abelian group (see Definition \ref{d8}). Moreover the pairing $<,>$ is invariant under the Weyl group of
$K$. It is shown in Proposition \ref{d6} that there is a one-to-one correspondence between conjugacy classes
of finite maximal diagonalizable subgroups of $\PGL(n,\mathbb{C})$
and finite symplectic abelian groups of order $n^2$. The following
theorem about the structure of the isometry group of a finite
nonsingular symplectic abelian group is Theorem \ref{c3}. For the
definition of a transvection on a symplectic abelian group, see
Definition \ref{d4}.
\begin{theorem}\lb{cc3}
Let $(H,<,>)$ be a finite nonsingular symplectic abelian group. Then
its isometry group is generated by the set of transvections in it.
\end{theorem}
If $K=P_n$, then in the important paper \cite{jpt} the authors showed that the respective Weyl group is $\SL(2,\mb Z_n)$. If $m=p^2$ with $p$ a
prime and $K=\texttt{P}_p\times \texttt{P}_p$, then in \cite{pst}
the authors proved that the respective Weyl group is
$\Sp(4,\mathbb{Z}_p)$. Next in \cite{hg} we dealt with the case
$n=m^k$ ($m$ may not be a prime) and $K=\texttt{P}_m^{k}$ is the
$k$-fold direct product of $\texttt{P}_m$, and proved that the Weyl
group is isomorphic to $\Sp(2k,\mathbb{Z}_n)$ and is generated by
transvections. Then, in this paper we deal with the general case
that $K$ is an arbitrary discrete maximal diagonalizable subgroup,
and prove the following result in Theorem \ref{d7} generalizing the
previous result.
\begin{theorem}\lb{dd9}
Let $K$ be a finite maximal diagonalizable subgroup of
$G=\PGL(n,\mathbb{C})$ and $W_G(K)$ be its Weyl group. Then $W_G(K)$
equals the isometry group of $(K,<,>)$, and is generated by the set
of transvections in it.
\end{theorem}
The paper is organized as follows. The definition and classification of
finite nonsingular symplectic abelian groups will be reviewed in
Section 2. Then in Section 3 we will define transvections on a
finite symplectic abelian group and prove Theorem \ref{cc3}. Next in
section 4 we will review the definitions of the grading of a simple
Lie algebra and prove some important properties. Then in Section 5
we will define the anti-symmetric pairing on any finite maximal
diagonalizable subgroup of $\PGL(n,\mathbb{C})$ and prove the $1-1$
correspondence between conjugacy classes of finite maximal
diagonalizable subgroups of $\PGL(n,\mb C)$ and nonsingular
symplectic abelian groups of order $n^2$. In the last section,
Theorem \ref{dd9} will be proved.\bigskip
Finally we introduce some notations in the paper.
For a finite set $S$, we will use $|S|$ to denote its cardinality.
For any $n\in \mb Z_+$, let $\mb Z_n=\mb Z/n\mb
Z=\{\bar{0},\bar{1},\cdots,\overline{n-1}\}$. For simplicity
we will just use $i$ to denote $\bar{i}$ for $i=0,1,\cdots,n-1$.
Let $\omega_n=e^{2\pi i/n}$ and
$C_n=\{\omega_n^i|i=0,1,\cdots,n-1\}$ be the cyclic group of order
$n$ generated by $\omega_n$. Sometimes we will identify $\mb Z_n$
with $C_n$ by mapping $i$ to $\omega_n^i$.
\bigskip
\centerline{\textit{Acknowledgments}}
The research is supported by NSFC Grant No.10801116 and by 'the
Fundamental Research Funds for the Central Universities'. It is
finished during the author's visit at the department of mathematics
in MIT in 2011. He acknowledges the hospitality of MIT, and would
like to take this opportunity to heartily thank David Vogan for
drawing his attention to this problem and for Vogan's great
generosity in sharing his immense knowledge with him during the
research. Proposition \ref{d0} is due to him.
\section{Classification of finite nonsingular symplectic abelian groups
}
\setcounter{equation}{0}\setcounter{theorem}{0}
We will follow the definition of symplectic abelian groups in
\cite{ka}, which is defined with respect to any field. But for our
purpose we will always assume the field to be $\mb C$,
and we will write our abelian groups additively in Section 2 and 3.
Let $H$ be an abelian group. Recall that an abelian group is
automatically a $\mb Z$-module.
\begin{defi}\lb{d8}
A map $$<,>:H\times H\rt \mb C^{\times}$$ is called a pairing of $H$ into $\mb C^{\times}$ if
$<,>$ is $\mb Z$-bilinear. The pairing is called anti-symmetric if
for all $a,b\in H$, $$<a,b>=<b,a>^{-1}.$$ An anti-symmetric pairing
$<,>$ is called nonsingular if $<a,b>=1$ for any $b\in H$ implying
$a=0$.
\end{defi}
\begin{defi}
Assume $<,>$ is an anti-symmetric pairing of $H$ into $\mb C^{\times}$. Then $(H,<,>)$
is called a \textit{symplectic abelian group}. A symplectic abelian group $(H,<,>)$ is said to be
\textit{nonsingular} if $<,>$ is nonsingular.
\end{defi}
Now assume that $(H,<,>)$ is a nonsingular symplectic abelian group.
A subgroup $H_0$ of $H$ is called a \textit{nonsingular symplectic
abelian subgroup} if the restriction $<,>|H_0$ is nonsingular.
Two subgroups $H_1$ and $H_2$ of $H$ are said to be
\textit{orthogonal}, written $H_1\perp H_2$, if $<a,b>=1$ for any
$a\in H_1, b\in H_2$.
Two symplectic abelian
groups are said to be \textit{isometric} if there is a group isomorphism
between them preserving the respective pairings.
If
$H_1,H_2,\cdots,H_n$ is a family of nonsingular symplectic
abelian subgroups of $H$ such that
$$H=H_1\oplus H_2\oplus\cdots\oplus H_n$$ and $$H_i\perp H_j,\ i\neq j,$$
then we will say that $H$ is the orthogonal direct sum of symplectic
abelian subgroups $H_1,H_2,\cdots,H_n$.
Assume $n\in \mb Z^{+}$ and $n>1$. If a pair of elements $u,v\in H$
of order $n$ satisfying $<u,v>=\omega_n$, then we call $(u,v)$ a
\textit{hyperbolic pair} of order $n$ in $H$. Let \be \mb H_n=\mb
Z_n\times \mb Z_n\lb{d3}\ee and $<(i,j),(k,l)>=\omega_n^{il-jk}$ be
the pairing on $\mb H_n$, which is clearly nonsingular and
anti-symmetric. Then $(\mb H_n,<,>)$ (or just $\mb H_n$) is a
nonsingular symplectic abelian group, called the \textit{hyperbolic
group} of rank $n$. Note that in \cite{ka} the rank $n$
for a hyperbolic group is assumed to be a power of a prime, but we
will not have this restriction in this paper. Let $u_1=(1,0)\in \mb
H_n,v_1=(0,1)\in\mb H_n$. Then the hyperbolic pair $(u_1,v_1)\in\mb
H_n^2$ is called the \textit{standard hyperbolic pair} of $\mb H_n$.
\begin{lem}\lb{e}
Let $H$ be a finite symplectic abelian group. If $a,b\in H$ are both
of order $n$ and $<a,b>=\omega_n$, then $a$ and $b$ generate a
subgroup $K$ isometric to $\mb H_n$.
\end{lem}
\bp Assume for some $i,j\in \mb Z$, $ia+jb=0$. Then
$<a,ia+jb>=\omega_n^j=1$ thus $n|j$. Similarly $n|i$. So
$K=\{ia+jb|i,j\in \mb Z_n\}$. Then $$K\rt \mb H_n,\ ia+jb\mapsto
(i,j)$$ is clearly an isometry of symplectic abelian groups. \ep
\begin{lem}\lb{c}
If $m$ and $n$ are relatively prime, then $\mb H_{mn}\cong \mb H_m
\op \mb H_n$ as symplectic abelian groups.
\end{lem}
\bp Let $(u_1,v_1)\in \mb H_m^2$ (resp. $(u_2,v_2)\in \mb H_n^2$) be
the standard hyperbolic pair for $\mb H_m$ (resp. $\mb H_n$). Let
$$a=u_1+u_2\in \mb H_m \op \mb H_n,\ b=v_1+v_2\in \mb H_m
\op \mb H_n.$$ The order of $a$ and $b$ are both $mn$ as $m$ and $n$
are relatively prime.
One has $<a,b>=<u_1,v_1><u_2,v_2>=\omega_m
\omega_n=\omega_{mn}^{m+n}$. As $mn$ and $m+n$ are also relatively
prime, $<a,ib>=\omega_{mn}$ for some integer $i$. Clearly
$\ord(ib)$, the order of $ib$, is still $mn$. Then by Lemma \ref{e},
the subgroup of $\mb H_m \op \mb H_n$ generated by $a$ and $ib$ is
isometric to $\mb H_{mn}$. As $|\mb H_{mn}|=|\mb H_m \op \mb H_n|$,
$$\mb H_m \op \mb H_n\rt \mb H_{mn},\ ja+k(ib)\mapsto (j,k)$$ is an
isometry of symplectic abelian groups. \ep
\begin{lem}\lb{x}
Let $(H,<,>)$ be a symplectic abelian group. Let $a,b\in H$ with
$ord(a)=i,\ ord(b)=j$. Assume $<a,b>=x$.
(1) If $l$ is the minimal positive integer such that $x^l=1$, then
$l|i$ and $l|j$.
(2) If $i,j$ are relatively prime then $x=1$.
\end{lem}
\bp (1) As $x^i=<ia,b>=<0,b>=1$, one has $l|i$. Similarly $l|j$.
(2) Apply (1). \ep
For any prime $p$ dividing $|H|$, we will always denote the $p$-Sylow subgroup of $H$ by $H(p)$.
\begin{theorem}\lb{d1}[Lemma 1.6 and Theorem 1.8 of \cite{ka}]
Let $(H,<,>)$ be a finite nonsingular symplectic abelian group. Then
(1) $H$ is an orthogonal direct sum of all its Sylow subgroups.
(2) Assume that $H(p)$ is a $p$-Sylow subgroup of $H$. Then $$H(p)\cong \mb H_{p^{r_1}}\op \mb H_{p^{r_2}}\op\cdots \op \mb
H_{p^{r_s}}$$ for some positive integers $r_1,r_2,\cdots,r_s$ with
$r_i\geq r_{i+1}$.
(3) $H$ is an orthogonal direct sum of hyperbolic subgroups $\mb
H_n$, where each $n$ is a power of some prime.
\end{theorem}
\bp (1) It is proved in Lemma 1.6 of \cite{ka}. In fact it is an
easy consequence of Lemma \ref{x} (2).
(2) This is proved in Theorem 1.8 of
\cite{ka} implicitly.
(3) It is a consequence of (1) and (2).
It is also proved in Theorem 1.8 of
\cite{ka}.
\ep
\begin{coro}\lb{d9}
Let $(H,<,>)$ be a nonidentity finite nonsingular symplectic abelian
group. Then there exists positive integers $l_1,l_2,\cdots,l_k$ with
each $l_i|l_{i-1}$ such that \be H\cong \mb H_{l_1}\op \mb
H_{l_2}\op \cdots \op \mb H_{l_k}\lb{b}.\ee Such positive integers
are uniquely determined by $H$.
\end{coro}
\bp The existence of such positive integers and the isometry
(\ref{b}) follow directly from (1) and (2) of Theorem \ref{d1}.
According to the structure theorem of finite abelian groups, such
positive integers are uniquely determined by $H$. \ep
Let $(H,<,>)$ be a nonsingular symplectic abelian group and $H_0$ be a finite nonsingular symplectic subgroup. For any $a\in H$ of
order $n$ and any $j\in \mb
Z$, define
\be \omega_n^j\cdot a=_{def}ja.
\lb{t1}\ee
Assume that $H_0\cong \mb H_{l_1}\op \mb H_{l_2}\op \cdots \op
\mb H_{l_k}$ with $l_i|l_{i-1}$ for all $i$. For $i=1,\cdots,k$ let
$(a_i,b_i)$ be a hyperbolic pair of order $l_i$ for $\mb H_{l_i}$.
Let \be\pi:H\rt H_0,\ c\mapsto \sum_{i=1}^k(<c,b_i>\cdot
a_i-<c,a_i>\cdot b_i).\lb{p}\ee Let $H_0^{\perp}=\{a\in H|<a,b>=0,
\forall\ b\in H_0\}$.
\begin{prop}
(1)$H_0^{\perp}$ is a symplectic subgroup of $H$ and $H=H_0\op
H_0^{\perp}$ is a direct sum of nonsingular symplectic abelian
subgroups.
(2)The map $\pi$ is independent of the hyperbolic pairs $(a_i,b_i)$
chosen.
\end{prop}
\bp
(1) First it is clear that $H_0^{\perp}$ is a subgroup of $H$.
$H_0\cap H_0^{\perp}=0$ as \\ $<,>|H_0$ is nonsingular. Let $c\in
H$. Assume $<c,a_i>=\omega_{l_i}^{t_i}$ for $i=1,\cdots,k$, then
\bee\begin{split}<\pi(c),a_i>&=<-<c,a_i>b_i,a_i>\\&=<-t_i
b_i,a_i>\\&=<b_i,a_i>^{-t_i}=\omega_{l_i}^{t_i}.\end{split}\eee So
$$<c,a_i>=<\pi(c),a_i>$$ for
$i=1,\cdots,k$. Similarly $$<c,b_i>=<\pi(c),b_i>$$ for
$i=1,\cdots,k$. Thus one has $c-\pi(c)\in H_0^{\perp}$ and \be
c=\pi(c)+(c-\pi(c))\in H_0+H_0^{\perp}.\lb{q}\ee So $ H=H_0\op
H_0^{\perp}.$ It is clear that $<,>|H_0^{\perp}$ must also be
nonsingular as $<,>$ is nonsingular. Thus $H=H_0\op H_0^{\perp}$ is a direct sum of
nonsingular symplectic abelian subgroups.
(2) Let $\pi^{'}:H\rt
H_0$ be defined as in (\ref{p}) with respect to another choice of
hyperbolic pairs $(a_i^{'},b_i^{'})$ for each $\mb H_{l_i}$. Then
for any $c\in H$ one also has $$c=\pi^{'}(c)+(c-\pi^{'}(c))\in
H_0\op H_0^{\perp}.$$ Comparing to (\ref{q}), as $H=H_0\op
H_0^{\perp}$ is a direct sum, one must have \\ $\pi(c)=\pi^{'}(c)$
for any $c\in H$. \ep We call the map $\pi:H\rt H_0$ defined in
(\ref{p}) the \textit{projection} of $H$ onto $H_0$.
\section{Transvections and isometry groups of finite nonsingular symplectic abelian groups
}
\setcounter{equation}{0}\setcounter{theorem}{0}
Let $(H,<,>)$ be a finite nonsingular symplectic abelian group.
\begin{defi}
Let $\Sp(H)$ be the set of isometries of $H$ onto itself. Then $\Sp(H)$ is clearly a group, called the isometry group
of $H$.
\end{defi}
If $H=H_1\op H_2$ is a direct sum of nonsingular symplectic abelian
subgroups then it is clear that
$$\Sp(H_1)\times \Sp(H_2)\rt \Sp(H),\
(\phi,\nu)(a,b)=(\phi(a),\nu(b))$$ embeds $\Sp(H_1)\times \Sp(H_2)$
as a subgroup of $\Sp(H)$. For any symplectic subgroup $H_0$ of $H$,
as $H=H_0\op H_0^{\perp}$, we will always regard $\Sp(H_0)$ as a
subgroup of $\Sp(H)$ by the embedding
$$\Sp(H_0)\hookrightarrow \Sp(H_0)\times \Sp(H_0^{\perp})\st
\Sp(H),\ \phi\mapsto (\phi,1),$$ where 1 denotes the identity map on
$H_0^{\perp}$.
\begin{prop}
Let $H$ be a finite nonsingular symplectic abelian group. Then $\Sp(H)$ acts transitively on the set of hyperbolic pairs in $H$ with the same order.
\end{prop}
\bp Assume $(a_1,b_1)$ and $(a_2,b_2)$ are two hyperbolic pairs in
$H$ with order $n$. Let $H_i=Span (a_i,b_i)$ for $i=1,2$. Then $H_i$
is a symplectic subgroup of $H$ isometric to $\mb H_n$. Let
$\phi_1:H_1\rt H_2$ be the isometry such that $\phi_1(a_1)=a_2,\
\phi_1(b_1)=b_2.$
One has $H=H_1\op H_1^{\perp}$ and $H=H_2\op H_2^{\perp}$. By
Theorem \ref{d1} (3), $H_1^{\perp}\cong H_2^{\perp}$. Fix an
isometry $\phi_2:H_1^{\perp}\rt H_2^{\perp}$. Then
$$\phi=(\phi_1,\phi_2):H_1\op H_1^{\perp}\rt H_2\op H_2^{\perp},\
(c,d)\mapsto (\phi_1(c),\phi_2(d))$$ is in $\Sp(H)$ and maps
$(a_1,b_1)$ to $(a_2,b_2)$. Thus $\Sp(H)$ acts transitively on the
set of hyperbolic pairs in $H$ with the same order. \ep
Let $$\wh H=_{def}Hom(H,\mb C^{\times})$$ be the abelian group of characters of $H$.
For any $a\in H$, define
$$\ga_a:H\rt \mb C^{\times},\ b\mapsto <a,b>.$$
\begin{lem} The map \be \varphi:H\rt \wh H,\ a\mapsto \ga_a\lb{g}\ee is an isomorphism of abelian groups.\end{lem}
\bp
As $<,>$ is $\mb Z$-bilinear, $\va$ is $\mb Z$-linear. As $<,>$ is
nonsingular, $\va$ is one-to-one, thus is onto as $|H|=|\wh H|$.
\ep
\begin{defi}\lb{a5}
For any $\ga\in \wh H$, let $$\ga^*=_{def}\varphi^{-1}(\ga).$$
\end{defi}
Then for any $a\in H$, $\ga(a)=<\ga^*,a>$.
As (\ref{g}) is an isomorphism, $\ga$ and $\ga^*$ have the same
order, and for $\ga_1,\ga_2\in \wh H$,
\be (\ga_1+\ga_2)^*=\ga_1^*+\ga_2^*.\lb{b3}\ee
Recall that for a vector space $V$ over $\mb C$ (or other field), a linear map
$\phi\in \GL(V)$ is called a transvection if
$$\phi(v)=v+\lambda(v)u$$ for some $\lambda\in V^*, u\in V$ satisfying
$\lambda(u)=0$. If $V$ is a symplectic vector space with $<,>$ the
anti-symmetric pairing on it, then any transvection of $V$
preserving the form $<,>$ must be of the form
\be\phi(v)=v-k<u,v>u\lb{d5}\ee for some $u\in V,\ k\in \mb C$. One
knows that $\SL(V)$ and $\Sp(V)$ are both generated by their
transvections. One can define transvections for a symplectic abelian
group analogously.
For any $b\in H$ with $b\neq 0$, assume $\ord(b)=m$. For any $a\in
H$, $<b,a>=\ga_b(a)$ takes value in the cyclic group
$C_m=\{\omega_m^i|i=0,1,\cdots,m-1\}$. Recall the convention in
(\ref{t1}).
\begin{defi}\lb{d4} For any $b\in H$ with $b\neq 0$ and $k\in\mb Z$,
define a homomorphism \be s_{b,k}:H\rt H,\ a\mapsto a-k(<b,a>\cdot
b),\lb{g3}\ee and call it a transvection on $H$. Using the
identification $\varphi$ of $H$ and $\wh H$, for any $\ga\in \wh H$
define $ s_{\ga,k}= s_{\ga^*,k}$. Denote $s_{b,1}$ (resp.
$s_{\ga,1}$) by $s_{b}$ (resp. $s_{\ga}$) for simplicity.
\end{defi}
Then one
has \be s_\ga(a)=a-\ga(a)\ga^*.\lb{b9}\ee
\begin{lem} (1) For any $b\in H$ with $b\neq 0$ and any $k,j\in\mb Z$, one has
\be s_{b,0}=1,\lb{g0}\ee \be s_{b,k}
s_{b,j}= s_{b,k+j}, \lb{g1}\ee and
\be s_{b,k}^{-1}= s_{b,-k}\lb{g2}\ee
(2) If $\ord(b)=m$, them $\{ s_{b,k}|k\in\mb Z\}$ is a cyclic
group of order $m$ generated by $ s_{b,1}$.
\end{lem}
\bp (1) (\ref{g0}) follows from the definition (\ref{g3}). For any
$a\in H$, \bee\begin{split} s_{b,k}
s_{b,j}(a)&= s_{b,k}(a-j(<b,a>\cdot b))\\&=(a-j(<b,a>\cdot
b))-k(<b,a-j(<b,a>\cdot b)>\cdot b)\\&=a-(k+j)(<b,a>\cdot b)\\&=
s_{b,k+j}(a),\end{split}\eee So (\ref{g1}) holds. Then (\ref{g2})
follows from (\ref{g0}) and (\ref{g1}).
(2) It follows from (1). \ep
\begin{lem}
One has $ s_{b,k}\in \Sp(H)$, where $b\in H$ with $b\neq 0$ and $k\in\mb
Z$.
\end{lem}
\bp By (2) of last lemma one only need to show $ s_b\in \Sp(H)$. By
(\ref{g3}) and (\ref{g2}) it is clear that $ s_b$ is a $\mb
Z$-linear isomorphism of $H$, so we only need to prove that $ s_b$
preserves $<,>$. Assume $a,c\in H$ and $<b,a>=\omega_m^i$,
$<b,c>=\omega_m^j$. Then \bee
\begin{split}
< s_b(a), s_b(c)>&=<a-<b,a>b,c-<b,c>b>\\
&=<a-ib,c-jb>\\&=<a,c> <b,c>^{-i}<a,b>^{-j}\\
&=<a,c> \omega_m^{j(-i)}\omega_m^{(-i)(-j)}\\
&=<a,c>.
\end{split}
\eee
\ep
Let \be Q(H)=_{def}< s_{b,k}|0\neq b\in H,k\in \mb Z>\ee be the
subgroup of $\Sp(H)$ generated by all the transvections. It is clear
that $Q(H)$ is generated by those $ s_b$ with $b\in H$ and $b\neq
0$. For any element $b\neq 0$ in a nonsingular symplectic subgroup
$H_0$ of $H$, $ s_b\in Q(H_0)$ can be regarded as in $Q(H)$ since
$b\in H$. Thus $Q(H_0)$ can be naturally regarded as a subgroup of
$Q(H)$.
Let $\GL(2,\mathbb{Z}_n)$ be the group of $2\times 2$ invertible
matrices in $M(2,\mathbb{Z}_n)$. Let $$\SL(2,\mb Z_n)=\{A\in
\GL(2,\mathbb{Z}_n)|\det(A)=1\in\mb Z_n\}.$$ Let $$J=\left(
\begin{array}{cc}
0 & 1\\
-1 & 0 \\
\end{array}
\right)\in M(2,\mb Z_n)$$
and $$\Sp(2,\mb Z_n)=\{A\in \GL(2,\mathbb{Z}_n)|A^t J A=J\}.$$ It is
easily verified that $\SL(2,\mb Z_n)=\Sp(2,\mb Z_n)$.
\begin{lem}\lb{h}
One has $\Sp(\mb H_n)\cong \Sp(2,\mb Z_n)=\SL(2,\mb Z_n)$ and $\Sp(\mb
H_n)=Q(\mb H_n)$. In particular, $\Sp(\mb H_n)$ is generated by $
s_{u_1}$ and $ s_{v_1}$, where $(u_1,v_1)$ is the standard
hyperbolic pair of $\mb H_n$.
\end{lem}
\bp Note that $(u_1,v_1)$ is an (ordered) $\mb Z_n$-basis for $\mb
H_n$. For any $\varphi\in \Sp(\mb H_n)$, one has
$$\va(u_1)=a_{11}u_1 +a_{21}v_1,\ \va(v_1)=a_{12}u_1+a_{22}v_1$$ where $a_{ij}\in\mb Z_n$. Then with respect to the $\mb Z_n$-basis $(u_1,v_1)$, the matrix of $\va$ is defined to be \be C=\left(
\begin{array}{cc}
a_{11} & a_{12} \\
a_{21} & a_{22}\\
\end{array}
\right),\lb{}\ee which is in $\GL(2,\mathbb{Z}_n)$.
This defines a map $\Sp(\mb H_n)\rt \GL(2,\mb Z_n)$, which is an injective homomorphism. For any $a\in\mb H_n$, $a=i u_1+j v_1$
with $i,j\in\mb Z_n$. Then the coordinate $\wt a$ of $a$ is denoted $\wt a=(i,j)^t$, the transpose of $(i,j)$. It is clear that \be \wt{\va(a)}=C \wt a.\lb{n}\ee
If we identify $$C_n=\{\omega_n^i|i=0,1,\cdots,n-1\}\rt\mb Z_n,\
\omega_n^i\rt i,$$ then the matrix of the pairing $<,>$ in the $\mb
Z_n$-basis $(u_1,v_1)$ is $J$.
For any $a,b\in\mb H_n$, one has \be <a,b>=\wt a^t J \wt b\lb{m}.\ee
As $\va\in \Sp(\mb H_n)$, by (\ref{n}) and (\ref{m}) one has
$$C^t J C=J,$$ thus $C\in \Sp(2,\mb Z_n)$. So $\Sp(\mb H_n)\st \Sp(2,\mb Z_n)=\SL(2,\mb Z_n)$.
As $$ s_{u_1}(u_1)=u_1,\ s_{u_1}(v_1)=v_1-<u_1,v_1>u_1=v_1-u_1$$ so
the matrix of $ s_{u_1}$ is $$A=\left(
\begin{array}{cc}
1 & -1\\
0 & 1 \\
\end{array}
\right) .$$
As $$ s_{v_1}(u_1)=u_1-<v_1,u_1>v_1=u_1+v_1,\ s_{v_1}(v_1)=v_1$$ so
the matrix of $ s_{v_1}$ is $$B=\left(
\begin{array}{cc}
1 & 0\\
1 & 1 \\
\end{array}
\right) .$$
It is clear that $A,B$ generate $\SL(2,\mb Z_n)$, thus $\Sp(\mb
H_n)=Q(\mb H_n)= \SL(2,\mb Z_n)$. \ep
\begin{coro}\lb{l}
For any $a\in \mb H_n$, if $\ord(a)=n$ then $a$ is conjugate to $u_1$ under $\Sp(\mb H_n)$; otherwise $a$ is conjugate to $l u_1$ for
some $l\in\mb Z_n$.
In particular all the elements in $\mb H_n$ with order $n$ are
conjugate under $\Sp(\mb H_n)$.
\end{coro}
\bp Assume $a=(i,j)$. As $\mb Z_n$ is a principal ideal ring, the
ideal $I(i,j)$ in $\mb Z_n$ generated by $i,j$ must be generated by
some $l\in\mb Z_n$. So $I(i,j)=I(l)$.
If $\ord(a)=n$, $I(l)=\mb Z_n$. So there exists $k,m\in \mb Z_n$
such that $ik+jm=1$. Let $A=\left(
\begin{array}{cc}
k& m\\
-j & i \\
\end{array}
\right)\in \SL(2,\mb Z_n)$. Then $A(i,j)^t=(1,0)^t$, so $a$ is conjugate to $u_1$ by $A$.
If $\ord(a)<n$, then $(i,j)=l(i^{'},j^{'})$ for some
$a^{'}=(i^{'},j^{'})\in \mb H_n$ as $l$ is the greatest common
devisor of $i,j$. Then $\ord(a^{'})=n$ and there exists some $A\in
\SL(2,\mb Z_n)$ such that $A(a^{'})^t=(1,0)^t$ and $A a^t=(l,0)^t$.
Thus $a$ is conjugate to $l u_1$ by $A$. \ep
\begin{lem}\lb{o}
Let $H=\mb H_n\op \mb H_n$ and $\phi:H\rt H,\ (a,b)\mapsto (b,a)$. Then $\phi\in Q(H)$.
\end{lem}
\bp
Let $(u_1,v_1)$ (resp. $(u_2,v_2)$) be the standard hyperbolic pair
in $\mb H_n\op 0$ (resp. $0\op \mb H_n$). Then $\phi$ maps
$u_1$ to $u_2$ and $v_1$ to $v_2$. Let $x=v_1+v_2$, then $(u_1,x)$ and
$(u_2,x)$ are both hyperbolic pairs of order $n$. Let $H_1=Span
(u_1,x)$, then $H=H_1\op H_1^{\perp}$. By Corollary \ref{l}, there
exists $$\tau=(\tau^{'},1)\in \Sp(H_1)\times \Sp(H_1^{\perp})\st
\Sp(H)$$ such that $\tau(x)=u_1$. Similarly there exists $\va\st
\Sp(H)$ such that $\va(u_2)=x$. Then $\tau\va(u_2)=u_1$. Let
$v_2^{'}=\tau\va(v_2)$.
Assume $<v_2^{'},v_1>=\omega_n^i$. Then
$$ s_{u_1}^{i-1}(u_1)=u_1,$$
$$ s_{u_1}^{i-1}(v_2^{'})=v_2^{'}-(i-1)(<u_1,v_2^{'}>\cdot u_1)=v_2^{'}-(i-1)u_1.$$ Let $v_2^{''}=v_2^{'}-(i-1)u_1$. Then $<v_2^{''},v_1>=\omega_n$.
Let $q=v_2^{''}-v_1$. Note $\ord(q)=n$, then $$ s_q(u_1)=u_1-<q,u_1>\cdot q=u_1,$$
$$ s_q(v_2^{''})=v_2^{''}-<q,v_2^{''}>\cdot q=v_2^{''}-\omega_n\cdot q=v_2^{''}-q=v_1.$$ So the map $\nu= s_q s_{u_1}^{i-1}
\tau\va\in Q(H)$ maps $(u_2,v_2)$ to $(u_1,v_1)$. Then $\nu\phi$
fixes $(u_1,v_1)$ and maps its orthocomplement $0\op \mb H_n$
isometrically onto itself. Then there exists
$\theta=(1,\theta^{'})\in \Sp(\mb H_n)\times \Sp(\mb H_n)$ such that
$\theta\nu\phi=1$. So $\phi\in Q(H)$ as $\theta$ and $\nu$ are both
generated by transvections.
\ep
\begin{lem}
Assume that $H=H(p)$ for some prime $p$. Then by Theorem \ref{d1} (2), $H=\mb H_{p^{r_1}}\op \mb H_{p^{r_2}}\op\cdots \op \mb H_{p^{r_s}}$ for some positive integers
$r_1,r_2,\cdots,r_s$ with $r_i\geq r_{i+1}$. For any $a=(a_1,\cdots,a_s)\in H$ with order $p^{r_1}$, there exists $\phi\in Q(H)$ such that
$\phi(a)=b=(b_1,\cdots,b_s)$ with $\ord({b_1})=p^{r_1}$.
\end{lem}
\bp
In this case one has $\ord(a)=Max_{i=1}^s\{\ord(a_i)\}$. If $\ord(a_1)=p^{r_1}$ then we take $\phi=1$. If $\ord(a_i)=p^{r_1}$ for some $i>1$, then $r_i=r_1$.
Then as the subgroups $\mb H_{p^{r_1}}$ and $\mb H_{p^{r_i}}$ of $H$ are isometric, by Lemma \ref{o} there exists some
$\phi\in Q(\mb H_{p^{r_1}} \op \mb H_{p^{r_i}})\st Q(H)$
such that $$\phi(a)=\phi(a_1,\cdots,a_i,\cdots,a_s)=(a_i,\cdots,a_1,\cdots,a_s)=b. $$ Then $b_1=a_i$ and $\ord({b_1})=p^{r_1}$.
\ep
\begin{lem}
Assume $H=\mb H_{l_1}\op \mb H_{l_2}\op \cdots \op \mb H_{l_k}$ with $l_i|l_{i-1}$ for all $i$. Then for any $a=(a_1,\cdots,a_k)\in H$ with order $l_1$,
there exists $\phi\in Q(H)$ such that $\phi(a)=b=(b_1,\cdots,b_k)$ with $\ord(b_1)=l_1$.
\end{lem}
\bp Let $p_1,\cdots,p_s$ be the set of primes dividing $|H|$. Then
$H=\op_i H(p_i)$ and $H(p_i)=\mb H_{l_1}(p_i)\op \mb H_{l_2}(p_i)\op
\cdots \op \mb H_{l_k}(p_i)$. Let $\pi_i:H\rt H(p_i)$ be the
projection. Then $$\pi_i(a)=(a_{1i},a_{2i},\cdots,a_{ki})\in \mb
H_{l_1}(p_i)\op \mb H_{l_2}(p_i)\op \cdots \op \mb H_{l_k}(p_i).$$
One has $a=\sum_i \pi_i(a)$. By last lemma there exists $\phi_i\in
Q(H(p_i))\st Q(H)$ such that
$\phi_i(\pi_i(a))=(b_{1i},b_{2i},\cdots,b_{ki})$ with
$\ord(b_{1i})=\ord(\pi_i(a))$. Let $\phi=\Pi_i \phi_i$. Then
$\phi(a)=b=(b_1,b_2,\cdots,b_k)\in \mb H_{l_1}\op \mb H_{l_2}\op
\cdots \op \mb H_{l_k}$, where $b_1=\sum_i b_{1i}$ and
$\ord(b_{1})=\ord(b)=\ord(a)=l_1$.
\ep
\begin{lem}
Assume $H=\mb H_{l_1}\op \mb H_{l_2}\op \cdots \op \mb H_{l_k}$ with $l_i|l_{i-1}$ for all $i$. Then for any $a=(a_1,a_2,\cdots,a_k)\in H$ with order $l_1$,
there exists $\phi\in Q(H)$ such that $\phi(a)=b=(u_1,0,\cdots,0)$, where $u_1=(1,0)\in \mb H_{l_1}$.
\end{lem}
\bp By last lemma we can assume that $\ord({a_1})=l_1$. Then use
induction on the number $t$ of nonzero elements in
$\{a_1,\cdots,a_k\}$. The case $t=1$ follows from Corollary \ref{l}.
Assume there are $t=l\geq 2$ nonzero elements in
$\{a_1,a_2,\cdots,a_k\}$ and the result holds for $l-1$. Without
loss of generality we can assume $a_2\neq 0$. As $\ord({a_1})=l_1$,
there exists some $b_1\in \mb H_{l_1}\st H$ such that
$<b_1,a_1>=\omega_{l_1}$. Let $b=(b_1,a_2,0\cdots,0)$. Then
$\ord({b})=l_1$ and $<b,a>=\omega_{l_1}$, so \be \begin{split}
s_b(a)&=a-<b,a>\cdot b\\&=a-\omega_{l_1}\cdot b=a-b
\\&=(a_1-b_1,0,a_3,\cdots,a_k).
\end{split}\ee So $ s_b(a)=(a_1-b_1,0,a_3,\cdots,a_k)$ differs with $a$
only in the first and second position. As $ s_b(a)$ has order $l_1$
and has $l-1$ nonzero elements, by induction there exists
$\phi_1\in Q(H)$ such that $\phi_1( s_b(a))=(u_1,0,\cdots,0)$. Then
$\phi=\phi_1 s_b\in Q(H)$ has the desired property and the result
holds for $t=l$.
\ep
\begin{coro}\lb{j}
$Q(H)$ acts transitively on the set of elements in $H$ with maximal order.
\end{coro}
\begin{lem}
Assume $H=\mb H_{l_1}\op \mb H_{l_2}\op \cdots \op \mb H_{l_k}$ with $l_i|l_{i-1}$ for all $i$ and $G=\Sp(H)$. For $i=1,\cdots,k$ let $(u_i,v_i)$ be the standard hyperbolic pair in $\mb H_{l_i}$. Then (1) $G_{u_1}=Q(H)_{u_1}$. (2) $G=Q(H)$.
\end{lem}
\bp We will prove them by induction on $k$.
The case $k=1$ follows from Lemma \ref{h} as we proved there
$\Sp(H)=Q(H)$ if $H=\mb H_{l_1}$. Let $k>1$. Assume (1) and (2) holds for
$k-1$.
Assume $\si\in G_{u_1}$. As
$<u_1,v_1>=<\si(u_1),\si(v_1)>=<u_1,\si(v_1)>$,
$<u_1,\si(v_1)-v_1>=0$ so
$$\si(v_1)=v_1+j_1 u_1+\sum_{i=2}^k (p_i u_i+ q_i v_i)$$ for some
$j_1,p_i,q_i\in \mb Z$. By Corollary \ref{l} there exists $\phi_i\in
Q(\mb H_{l_i})\st Q(H)$ such that $\phi_i(p_i u_i+ q_i v_i)=j_i u_i$
for $i\geq 2$. Let $\phi=\Pi_{i=2}^k\ \phi_i$. Then $\phi\in
Q(H)_{u_1}$ and $$\phi \si(v_1)=v_1+j_1 u_1+\sum_{i=2}^k j_i u_i.$$
For any $i$ with $2\le i\le k$, $ s_{u_1+u_i}(u_t)=u_t$ for
$t=1,\cdots,k$. As $\ord(u_1+u_i)=l_1$, \be
\begin{split} s_{u_1+u_i}(v_1)&=v_1-<u_1+u_i,v_1>\cdot (u_1+u_i)\\&=v_1-\omega_{l_1}\cdot (u_1+u_i)\\&=v_1-(u_1+u_i).\end{split}\ee
Let $\tau=\Pi_{i=2}^k\ s_{u_1+u_i}^{j_i}$. Then $\tau\in
Q(H)_{u_1}$ and
$$ \tau(\phi\si(v_1)
)=v_1+(j_1-\sum_{i=2}^k j_i)u_1.$$
So $\tau\phi\si$ preserves $\mb H_{l_1}$ and also $\mb
H_{l_1}^{\perp}=\mb H_{l_2}\op \cdots \op \mb H_{l_k}$ thus
$$\tau\phi\si\in \Sp(\mb H_{l_1})\times \Sp(\mb H_{l_2}\op \cdots
\op \mb H_{l_k}).$$ By induction $$\Sp(\mb H_{l_1})\times \Sp(\mb
H_{l_2}\op \cdots \op \mb H_{l_k})=Q(\mb H_{l_1})\times Q(\mb
H_{l_2}\op \cdots \op \mb H_{l_k}),$$ so $$\tau\phi\si\in Q(\mb
H_{l_1})\times Q(\mb H_{l_2}\op \cdots \op \mb H_{l_k})\st Q(H).$$
As $\tau,\phi\in Q(H)$, one also has $\si\in Q(H)$. So $\si\in
Q(H)_{u_1}$ then $G_{u_1}\st Q(H)_{u_1}$. As $G\supset Q(H)$, one
must have $G_{u_1}= Q(H)_{u_1}$. So (1) holds for $k$.
By Corollary \ref{j}, $Q(H)$ acts transitively on the set of
elements in $H$ with maximal order, so does $G$. As
$$G_{u_1}=Q(H)_{u_1}\ \an \ |G/G_{u_1}|=|Q(H)/Q(H)_{u_1}|,$$ so
$|G|=|Q(H)|$ thus $G=Q(H)$. So (2) also holds for $k$. \ep
Now we have proved the following theorem.
\begin{theorem}\lb{c3}
Let $H$ be a finite nonsingular symplectic abelian group. Then
$\Sp(H)$ is generated by the set of transvections on $H$.
\end{theorem}
\begin{rem}
If $H=\mb H_n^k$, the $k$-fold direct sum of $\mb H_n$, then by
choosing some suitable $\mb Z_n$-basis of $\mb H_n$, it is easy to
see that
$$\Sp(\mb H_n^k)\cong\Sp(2k,\mb Z_n),$$ where
$$\Sp(2k,\mathbb{Z}_n)=\{A\in GL(2k,\mathbb{Z}_n)|A^t J_{2k}
A=J_{2k}\}$$ with $$J_{2k}=\left(
\begin{array}{ccccc}
0 & 1 & & &\\
-1 & 0 & & &\\
& & \ddots & &\\
& & & 0 & 1 \\
& & &-1 & 0 \\
\end{array}
\right)\in M(2k,\mathbb{Z}_n).$$ Thus this theorem implies
that in particular $\Sp(2k,\mb Z_n)$ is generated by the
transvections.
\end{rem}
\section{Fine gradings of Lie algebras
and their Weyl groups }
\setcounter{equation}{0}\setcounter{theorem}{0}
4.1. The details of this subsection can be found in Section 4 of \cite{hg}. We include it for completeness.
We always assume that $L$ is a complex simple Lie algebra. Let
$\Aut(L)$ be its automorphism group and $\Int(L)$ the identity
component of $\Aut(L)$, called the inner automorphism group
of $L$. It is clear that $\Aut(L)$ and $\Int(L)$ are both algebraic groups.
Let $\Lambda$ be a additive abelian group. A
$\Lambda$-\textit{grading} $\Gamma$ on $L$ is the decomposition of
$L$ into direct sum of subspaces $$\Gamma: L=\oplus_{\ga\in \Lambda}
\ L_\ga$$ such that
$$[L_\ga,L_\dt]\st L_{\ga+\dt},\ \forall\\ \ga,\dt\in \Lambda.$$ Let
$\ddt=\{\ga\in \Lambda|L_\ga\neq 0\}$. We will always
assume that $\Lambda$ is generated by $\ddt$, otherwise it could be
replaced by its subgroup generated by $\ddt$. So $\Lambda$ is always finitely generated.
Given a $\Lambda$-grading $\Gamma$ on $L$, let $$K=\widehat
\Lambda=_{def} \Hom(\Lambda, \mathbb{C}^{\times})$$ be the abelian
group of characters of $\Lambda$. Then $K$ acts on $L$ by
$$\si\cdot X=\si(\ga) X,\ \forall\\ X\in L_\ga,\ \forall\ \ga\in \Lambda,\
\forall\\si\in K.$$ This defines an injective homomorphism $K\rt
\Aut(L)$. So $K$ can be viewed as a subgroup of $\Aut(L)$. Recall
that an algebraic group is called \textit{diagonalizable} if it is
abelian and consists of semisimple elements. It is easy to see that
$K$ is a diagonalizable algebraic subgroup of $\Aut(L)$.
Conversely, given a diagonalizable algebraic subgroup $K$ of
$\Aut(L)$, let $$\Lambda=\widehat K=_{def} \Hom(K,
\mathbb{C}^{\times})$$ be the (additive) abelian group of
homomorphisms from $K$ to $\mb C^{\times}$ as algebraic groups. Then
one has a $\Lambda$-grading on $L$: $$\Ga:\ \ L=\oplus_{\gamma\in
\Lambda}\ L_\ga,$$ where $L_\ga=\{X\in L|\si \cdot X=\ga(\si) X,\
\forall\ g\in K\}.$ Let $$\ddt=\ddt(L,K)=_{def} \{\ga\in
\Lambda|L_{\ga}\neq 0\}.$$ We call $\ddt$ the set of \textit{roots}
of $K$ in $L$.
Thus there is a natural one-to-one correspondence between gradings of
$L$ by finitely generated abelian groups and diagonalizable
algebraic subgroups of $\Aut(L)$. A grading is called \textit{inner} if the respective diagonalizable subgroup is in
$\Int(\L)$. A grading (resp. inner grading) of $L$ is called fine if it could not be further refined by any other grading (resp. inner grading).
It is clear that the bigger the
diagonalizable algebraic subgroup $K$ is, the finer the
corresponding grading is. Thus a grading (resp. inner grading) $\Ga$ on $L$ is
fine if and only if the corresponding diagonalizable subgroup $K$
is a maximal diagonalizable subgroup of $\Aut(L)$ (resp. $\Int(L)$).
Let $G$ be either $\Aut(L)$ or $\Int(L)$. Let $K$ be a {maximal} diagonalizable subgroup of $G$,
$\Ga$ the grading on $L$ induced by the action of $K$. One could
define the Weyl group $W_G(\Ga)$ of the grading $\Ga$ with respect
to $G$, see Definition 2.3 of \cite{hg}, to describe the symmetry of
the grading $\Ga$.
One has the following result in \cite{hg}.
\begin{prop}[Corollary 2.6 of \cite{hg}]
Let $L$ be a simple Lie algebra and $G=\Int(L)$. Let $K$ be a maximal diagonalizable subgroup of $G$ and $\Ga$ be the corresponding grading
on $L$ induced by the action of $K$. Let $W_G(K)=N_G(K)/K$ be the Weyl group
of $K$ with respect to $G$. Then one has $W_G(\Ga)=W_G(K)$. \end{prop}
\bigskip
4.2. From now on we will always assume $K\st G=\Int(L)$ to be a {finite
maximal diagonalizable subgroup} and $\ddt=\ddt(L,K)$. Let $B$ be the Killing form on
$L$. Recall that a linear subspace $S$ of $L$ is called a toral
subalgebra if $[S,S]=0$ and the endomorphism $\ad_X$ is semisimple
for each $X\in S$.
As $L$ is simple, the adjoint map
$$\ad:L\rt \ad(L),\ X\mapsto \ad_X$$ is a $G$-equivariant isomorphism.
\begin{defi} For any
$\ga\in \ddt$, let
$$L_{[\ga]}=_{def}\oplus_{k\in \mathbb{Z}}\ L_{k\ga}.$$ \end{defi}
\begin{prop}\lb{d0}
(1) One has $L_0=0$, i.e., $0\notin \ddt$.
(2) Assume $\ga,\dt\in\ddt$ and $\ga+\dt\neq 0$. Then $B|L_\ga\times
L_\dt=0$. For any $X\in L_\ga$ with $X\neq 0$, there exists $Y\in
L_{-\ga}$ such that $B(X,Y)\neq 0$.
(3) For any $\ga\in \ddt$, $L^{\Ker \ga}=L_{[\ga]}$ and is a toral
subalgebra of $L$. One has $Lie\ Z(Ker \ga)_0=ad(L_{[\ga]})$ and
$Z(Ker \ga)_0$ is an algebraic torus (isomorphic to some
$(\mathbb{C}^\times)^i$).
\end{prop}
\bp (1) As $K$ is a maximal diagonalizable subgroup, $Z_G(K)=K$ by Lemma 2.2 of \cite{hg}. As
$K$ is finite, $$\ad(L^K)=\ad(L)^K=\Lie Z_G(K)=\Lie K=0.$$ So
$L_0=L^K=0$.
(2) Assume $\ga+\dt\neq 0$. For any $X\in L_\ga$ and $Y\in L_\dt$,
$\ad_X \ad_Y$ maps each $L_\zeta$ into $L_{\zeta+\ga+\dt}$ thus
$B(X,Y)=0$. Then $B|L_\ga\times L_\dt=0$. Because $B$ is nonsingular
on $L$, the second statement then follows from the first one.
(3) We first prove $L^{\Ker \ga}=L_{[\ga]}$. Choose some $\si\in K$
satisfying $\ga(\si)=\omega_m$, where $m$ is the order of $\ga$.
Then $\ga(\si)$ is a generator of the cyclic group $\ga(K)\cong
K/\Ker \ga$. $L^{\Ker \ga}$ is the direct sum of those
$L_\bt$ with $\bt$
being identity on $\Ker \ga$. Then $\bt(\si)=\omega_m^k$ for some
integer $k$ as $\si^m\in \Ker \ga $. Then
$\bt(\si)=\ga(\si)^{k}=(k\ga)(\si)$. As $\Ker \ga$ and $\si$
generate $K$, $\bt=k\ga$. Thus $L^{\Ker \ga}=\oplus_{k}
L_{k\ga}=L_{[\ga]}$.
One has $[L_0,L_{[\ga]}]=0$ as $L_0=0$. Then by Proposition 3.6 of
\cite{h}, $L_{[\ga]}$ is a toral subalgebra of $L$. As
$\Lie\Int(L)=\ad(L)$, it is clear that
$$\Lie\ Z(\Ker \ga)_0=\Lie\ Z(\Ker \ga)=\ad(L)^{\Ker \ga}=\ad(L^{\Ker
\ga})=\ad(L_{[\ga]}).$$Thus $Z(Ker \ga)_0$ is an algebraic torus.
\ep
\begin{rem}If $L$ is a semisimple Lie algebra then all the
results in this section still hold.
\end{rem}
\section{Finite maximal diagonalizable subgroups of $\PGL(n,\mb C)$ and anti-symmetric pairings on them}
\setcounter{equation}{0}\setcounter{theorem}{0}
Let $n\in \mathbb{Z}_+$. Recall $\omega_n=e^{2\pi i/n}$. Let
$$Q_n=diag(1,\omega_n,\omega_n^2,\cdots,\omega_n^{n-1})$$ and $$P_n=\left(
\begin{array}{ccccc}
0 & 1 & 0 &\cdots & 0 \\
0 & 0 & 1 &\cdots & 0 \\
\vdots& \vdots& \vdots & \ddots & \vdots \\
0 & 0 & 0 &\cdots & 1 \\
1 & 0 & 0 &\cdots & 0\\
\end{array}
\right).$$ Let $\Pi_n=\{\omega_n^j P_n^k Q_n^l\ |j,k,l=0,1,\cdots,n-1\}$. Note $P_n Q_n=\omega_n Q_n P_n$. This is a subgroup of $\GL(n,\mathbb{C})$,
called the \textit{Pauli group} of rank $n$.
Let $D_n$ be the subgroup of diagonal matrices
of $\GL(n,\mb{C})$. Let $\texttt{P}_n$ and $\texttt{D}_n$
be the respective images of $\Pi_n$ and $D_n$ under the adjoint action on $M(n,\mb C)$. One knows that
$$\texttt{P}_n=\{Ad_{P_n}^iAd_{Q_n}^j|i,j=0,\cdots,n-1\}\cong \mb
Z_n\times \mb Z_n,$$ and that
$\texttt{D}_n$ and $\texttt{P}_n$ are both maximal
diagonalizable subgroups of $\PGL(n,\mathbb{C})$.
Let $L=sl(n,\mathbb{C})$ and $G=\Int(L)\cong \PGL(n,\mathbb{C})$.
One has the standard isomorphism
$$M=M(t,\mathbb{C})\ot M(l_1,\mathbb{C})\ot \cdots \ot M(l_k,\mathbb{C})\rt M(n,\mathbb{C}),$$ where $n=t l_1\cdots l_k$ and $l_i|l_{i-1}$ for all $i$. It induces injective homomorphisms \be
D_t\ot\Pi_{l_1}\ot\cdots\ot \Pi_{l_k}\st S=\GL(t,\mathbb{C})\ot
\GL(l_1,\mb C)\ot \cdots \ot \GL(l_k,\mathbb{C})\rt
\GL(n,\mathbb{C}).\lb{a8}\ee
Let $A=A_0\ot A_1\ot\cdots \ot A_k\in S$. Then for $X=X_0\ot X_1\ot\cdots\ot X_k\in M$,
$$\Ad_A(X)=\Ad_{A_0}(X_0)\ot\Ad_{A_1}(X_1)\ot\cdots\ot \Ad_{A_k}(X_k).$$
Thus the adjoint action induces homomorphisms $$\GL(n,\mathbb{C})\rt \PGL(n,\mathbb{C}),$$
\bee \GL(t,\mathbb{C})\ot
\GL(l_1,\mb C)\ot \cdots \ot \GL(l_k,\mathbb{C})\rt \PGL(t,\mathbb{C})\times \PGL(l_1,\mb C)\times \cdots \times
\PGL(l_k,\mathbb{C}),\eee
\be\lb{a4} A=A_0\ot A_1\ot\cdots \ot A_k\mapsto Ad_A=(\Ad_{A_0},\Ad_{A_1},\cdots,\Ad_{A_k})\ee and by restriction
$$D_t\ot\Pi_{l_1}\ot\cdots\ot \Pi_{l_k}\rt \texttt{D}_t\times \texttt{P}_{l1}\times\cdots\times\texttt{P}_{l_k}.$$
Thus by (\ref{a8}) one has injective homomorphisms
\be \CD \texttt{D}_t\times
\texttt{P}_{l1}\times\cdots\times\texttt{P}_{l_k}\st
\PGL(t,\mathbb{C})\times \PGL(l_1,\mb C)\times \cdots \times
\PGL(l_k,\mathbb{C}) @>\phi>> \PGL(n,\mathbb{C}).
\endCD\lb{a9}
\ee
We will identify $\texttt{D}_t\times \texttt{P}_{l_1}\times \cdots
\times \texttt{P}_{l_k}$ with its image in $\PGL(n,\mb{C}).$
\begin{theorem}[Theorem 3.2 of \cite{hpp}] Any maximal diagonalizable subgroup of $\PGL(n,\mb{C})$ is conjugate to one and only one of the $\texttt{D}_t
\times \texttt{P}_{l_1}\times \cdots \times \texttt{P}_{l_k}$ with $n=t l_1\cdots l_k$ and each $l_i$ dividing $l_{i-1}$.
\end{theorem}
\begin{coro} \lb{b2} Any finite maximal diagonalizable subgroup of $\PGL(n,\mb{C})$ is conjugate to one and only one of the $\texttt{P}_{l_1}\times \cdots \times \texttt{P}_{l_k}$
with $n=l_1\cdots l_k$ and each $l_i$ dividing $l_{i-1}$.
\end{coro}
In the case $K=\texttt{P}_{l_1}\times \cdots \times
\texttt{P}_{l_k}$ is a finite maximal diagonalizable subgroup of
$\PGL(n,\mb{C})$, (\ref{a9}) becomes
\be \CD
\texttt{P}_{l1}\times\cdots\times\texttt{P}_{l_k}\st
\PGL(l_1,\mb C)\times \cdots \times
\PGL(l_k,\mathbb{C}) @>\phi>> \PGL(n,\mathbb{C})
\endCD\lb{b0}
\ee
\bigskip
Let $K\st \PGL(n,\mb{C})$ be a finite maximal diagonalizable
subgroup. Let
$$p:\GL(n,\mathbb{C})\rt \PGL(n,\mathbb{C})$$ be the projection.
\begin{defi} \lb{d} For
any $\si\in K$ fix some $\widetilde{\si}\in p^{-1}(\si)$. For any
$\si,\tau\in K$, $\widetilde{\si} \widetilde{\tau}
\widetilde{\si}^{-1} \widetilde{\tau}^{-1}=lI_n$ as
$p(\widetilde{\si} \widetilde{\tau} \widetilde{\si}^{-1}
\widetilde{\tau}^{-1})=1$. Clearly $l$ is independent of the
preimages $\widetilde{\si},\widetilde{\tau}$ chosen. Define
$<\si,\tau>= l$.
\end{defi}
\begin{lem}[Lemma 3.4 of \cite{hg}] The map $<,>:K\times K\rt \mathbb{C}^{\times}$ is an anti-symmetric pairing on $K$, which is invariant under $N_G(K)$.
\end{lem}
\begin{prop}\lb{b1}
Let $G=\PGL(n,\mb{C})$ and $K$ be a maximal diagonalizable subgroup
of $G$. If $K=\texttt{P}_{l_1}\times \cdots \times \texttt{P}_{l_k}$
with $n=l_1\cdots l_k$ and each $l_i$ dividing $l_{i-1}$, then $<,>$
is nonsingular on $K$. Thus $(K,<,>)$ is a nonsingular symplectic
abelian group isometric to $\mb H_{l_1}\op \cdots \op \mb H_{l_k}$.
\end{prop}
\bp For $i=1,\cdots,k$, let
$$\si_i=(1,\cdots,1,Ad_{P_{l_i}},1,\cdots,1)\in K,
\ \tau_i=(1,\cdots,1,Ad_{Q_{l_i}},1,\cdots,1)\in K$$ where
$Ad_{P_{l_i}}$ and $Ad_{Q_{l_i}}$ are in the $i$-th position. Then
$\{\si_i,\tau_i|i=1,\cdots,k\}$ is a set of generators of $K$ and
any element in $K$ can be written uniquely as
$\si_1^{i_1}\tau_1^{j_1}\cdots \si_k^{i_k}\tau_k^{j_k}$. By simple
computation one has for $i \neq j$,
$$<\si_i,\tau_i>=\omega_{l_i},\ <\si_i,\tau_j>=1,$$ $$<\si_i,\si_j>=1,\ <\tau_i,\tau_j>=1.$$
Thus $(\si_i,\tau_i)$ is a hyperbolic pair of order $l_i$ and spans
the symplectic subgroup $\texttt{P}_{l_i}$ isometric to $\mb
H_{l_i}$. It is clear that such subgroups $\texttt{P}_{l_i}$ are
mutually orthogonal to each other. The map
\be\lb{c4}\texttt{P}_{l_1}\times \cdots \times \texttt{P}_{l_k}\rt
\mb H_{l_1}\op \cdots \op \mb H_{l_k},\
\si_1^{i_1}\tau_1^{j_1}\cdots \si_k^{i_k}\tau_k^{j_k}\mapsto
((i_1,j_1),\cdots,(i_k,j_k))\ee is clearly an isometry of
nonsingular symplectic abelian groups. \ep
As a corollary of Corollary \ref{b2}, Proposition \ref{b1} and Corollary
\ref{d9} one has the following result.
\begin{prop}\lb{d6}
There is a one-to-one correspondence between conjugacy class of
finite maximal diagonalizable subgroups of $\PGL(n,\mathbb{C})$ and
nonsingular symplectic abelian groups of order $n^2$ .
\end{prop}
\section{Weyl groups of finite maximal diagonalizable subgroups of $\PGL(n,\mb C)$ }
\setcounter{equation}{0}\setcounter{theorem}{0}
Recall that $L=sl(n,\mb C)$ and $K$ is a finite maximal
diagonalizable subgroup of $G=\PGL(n,\mb{C})$. First we will
describe the grading of $sl(n,\mb C)$ and $gl(n,\mb C)$ induced by
the action of $K$.
At first let $K=\texttt{P}_n$. The character group $\wh{K}$, written
additively, is generated by $\bt_n$ and $\al_n$, which are dual to
$\Ad_{P_n},\Ad_{Q_n}$:
$$\bt_n(\Ad_{P_n})=\omega_n,\ \bt_n(\Ad_{Q_n})=1,$$
$$\al_n(\Ad_{P_n})=1,\ \al_n(\Ad_{Q_n})=\omega_n.$$
Thus $\wh{K}=\{i\bt_n+j\al_n|(i,j)\in \mathbb{Z}_n\times
\mathbb{Z}_n\}\cong \mathbb{Z}_n^2$.
One has
$$\bt_n(\Ad_{P_n}^{i}\Ad_{Q_n}^{j})=\omega_n^i=<\Ad_{Q_n}^{-1},\Ad_{P_n}^{i}\Ad_{Q_n}^{j}>$$
and
$$ \al_n(\Ad_{P_n}^{i}\Ad_{Q_n}^{j})=\omega_n^j=<\Ad_{P_n},\Ad_{P_n}^{i}\Ad_{Q_n}^{j}>,$$
so $$\bt_n^*=\Ad_{Q_n}^{-1},\ \al_n^*=\Ad_{P_n}.$$ As by (\ref{b3}) one has $(\ga+\dt)^*=\ga^*\dt^*$,
\be \lb{a1}(i\bt_n+j\al_n)^*=\Ad_{P_n}^{j}\Ad_{Q_n}^{-i}.\ee As
$$\Ad_{Q_n}(Q_n^i P_n^j)=\omega_n^{-j}Q_n^i P_n^j=(i\bt_n-j\al_n)(\Ad_{Q_n})Q_n^i P_n^j,$$
$$\Ad_{P_n}(Q_n^i P_n^j)=\omega_n^{i}Q_n^i P_n^j=(i\bt_n-j\al_n)(\Ad_{P_n})Q_n^i P_n^j,$$ one has
\be \lb{z} Q_n^i P_n^j\in L_{i\bt_n-j\al_n}.\ee In particular
$$P_n\in L_{-\al_n},\ Q_n\in L_{\bt_n}.$$ Note that $tr (Q_n^i P_n^j)=0$ unless
$(i,j)=(0,0)$. Let \be X_{i\bt_n-j\al_n}=Q_n^i P_n^{j}.\lb{a2}\ee
Then one has the following gradings
$$gl(n,\mathbb{C})=\oplus_{(i,j)}\ \mathbb{C} Q_n^i P_n^j=\op_{\ga\in \wh K}\ \mb C X_{\ga},$$
$$sl(n,\mathbb{C})=\oplus_{(i,j)\neq (0,0)}\ \mathbb{C} Q_n^i P_n^j=\op_{\ga\neq 0}\ \mb C X_{\ga}.$$ So
$\ddt(gl(n,\mathbb{C}),K)=\wh{K}$ and
$\ddt(sl(n,\mathbb{C}),K)=\wh{K}\setminus \{0\}$. Note that each
root space is one-dimensional, and for any $\ga\in\wh K$, by
(\ref{a1}) and (\ref{a2}) one has \be
\ga^*=(Ad_{X_\ga})^{-1}.\lb{a3}\ee The following result is
originally Theorem 10 of \cite{jpt}, for its proof we refer the
readers to Proposition 4.4 of \cite{hg}.
\begin{theorem}
Let $G=\PGL(n,\mb C)$ and $K=\texttt{P}_n$. One has $W_G(K)\cong
\SL(2,\mb Z_n)$ and is generated by $s_{\al_n}$ and $s_{\bt_n}$.
\end{theorem}
Next let $K=\texttt{P}_{l_1}\times \cdots \times \texttt{P}_{l_k}$
with $n=l_1\cdots l_k$ and each $l_i$ dividing $l_{i-1}$. As
$M(n,\mb C)=M(l_1,\mb C)\ot\cdots\ot M(l_k,\mb C)$, and $M(l_i,\mb
C)=\op_{\ga\in \wh{\texttt{P}_{l_1}}}\mb C X_\ga$, one has
$$ M(n,\mb C)=\op_{(\ga_1,\cdots,\ga_k)}\ \mb C X_{\ga_1}\ot\cdots\ot X_{\ga_k}.$$
Note that $\wh K=\wh{\texttt{P}_{l_1}}\times \cdots \times
\wh{\texttt{P}_{l_k}}$. Let $\ga=(\ga_1,\cdots,\ga_k)\in \wh K$ with
$\ga_i\in \wh{\texttt{P}_{l_i}}$. For any
$\si=(\si_1,\cdots,\si_k)\in K$, $\ga(\si)=\ga_1(\si_1)\cdots
\ga_k(\si_k)$ and one has \bee
\begin{split} \si\cdot X_{\ga_1}\ot\cdots\ot
X_{\ga_k}&=\si_1\cdot X_{\ga_1}\ot\cdots\ot \si_k\cdot
X_{\ga_k}\\&=\ga_1(\si_1)X_{\ga_1}\ot\cdots\ot
\ga_k(\si_k)X_{\ga_k}\\&=\ga(\si) X_{\ga_1}\ot\cdots\ot
X_{\ga_k}.\end{split}\eee
So $X_{\ga_1}\ot\cdots\ot
X_{\ga_k}\in L_\ga$. Note that $tr (X_{\ga_1}\ot\cdots\ot
X_{\ga_k})=\Pi_i tr(X_{\ga_i})$, which is nonzero if and only if
$\ga_1=\cdots=\ga_k=0$. Let
$$Y_\ga=X_{\ga_1}\ot\cdots\ot X_{\ga_k}.$$ Then
$$gl(n,\mathbb{C})=\op_{\ga\in \wh K}\ \mb C Y_\ga$$ and
$$sl(n,\mathbb{C})=\op_{\ga\neq 0}\ \mb C Y_\ga.$$
So $\ddt(gl(n,\mb C),K)=\wh K$ and $\ddt(sl(n,\mb C),K)=\wh
K\setminus \{0\}$. Note that each root space is also one-dimensional
and consists of semisimple elements.
\begin{lem}For any $\ga\in \wh K$, $\ga^*=(\Ad_{Y_\ga})^{-1}$. \end{lem}
\bp For any $Ad_X\in K$, as $Y_\ga$ is invertible, $$Y_\ga^{-1} X
Y_\ga X^{-1}=Y_\ga^{-1} (\ga(Ad_X)Y_\ga)=\ga(Ad_X) I,$$ so
$<(\Ad_{Y_\ga})^{-1},Ad_X>=\ga(Ad_X)$. Thus
$\ga^*=(\Ad_{Y_\ga})^{-1}$.
\ep
Recall in (\ref{b0}) one has the
embedding \bee \CD
\texttt{P}_{l1}\times\cdots\times\texttt{P}_{l_k}\st \PGL(l_1,\mb
C)\times \cdots \times \PGL(l_k,\mathbb{C}) @>\phi>>
\PGL(n,\mathbb{C}).
\endCD
\eee
Let $N(\texttt{P}_{l_i})$ be
the normalizer of $\texttt{P}_{l_i}$ in $\PGL(l_i,\mathbb{C})$, then
clearly $\phi$ restricts to \bee \CD
N(\texttt{P}_{l_1})\times\cdots\times N(\texttt{P}_{l_k}) @>\phi>>
\PGL(n,\mathbb{C}).
\endCD
\eee The left hand side is in $N_G(K)$. As
$$N(\texttt{P}_{l_i})/\texttt{P}_{l_i}\cong \SL(2,\mathbb{Z}_{l_i}),$$ one has
\be\lb{s} \SL(2,\mathbb{Z}_{l_1})\times\cdots\times
\SL(2,\mathbb{Z}_{l_k})\st W_G(K).\ee
Let $\ga\in \ddt(sl(n,\mb C),K)$ and $G=\PGL(n,\mb C)$. Assume the order of $\ga$ is $m$ and choose $\si\in K$ satisfying $\ga(\si)=\omega_m$. As $\si\in Z(Ker \ga)$,
$\Ad_\si$ maps $Z(Ker \ga)_0$ into $Z(Ker \ga)_0$. Let $f_\si:G\rt
G,\ \eta\mapsto\si\eta\si^{-1}\eta^{-1}$. Then $f_\si(Z(Ker
\ga)_0)\st Z(Ker \ga)_0$. Denote
$Z(\Ker \ga)_0$ by $Z_0$.
\begin{lem}\lb{a6}
(1) The map $f_\si:Z_0\rt Z_0,\ \eta\mapsto\si\eta\si^{-1}\eta^{-1}$ is a
continuous epimorphism.
(2) Assume $\ga^*\in Z_0$. Then there exists $\zeta\in Z_0$ with
$f_\si(\zeta)=\ga^*$. One has $\zeta\in N_G(K)$ and
$Ad_\zeta:K\rt K,\ \tau\mapsto \zeta\tau\zeta^{-1}$ is just the transvection $$ s_\ga:K\rt K,\ \tau\mapsto \tau(\ga^*)^{-\ga(\tau)}$$ as in
(\ref{b9}). (Note that as a subgroup of G,
$K$ is a multiplicative abelian group.) Thus $ s_\ga\in W_G(K)$.
\end{lem}
\bp As it was shown in Proposition \ref{d0} (3) that $L_{[\ga]}$ is
a toral subalgebra of $L$, the lemma follows from Lemma 3.7 of
\cite{hg}. \ep
\begin{lem}\lb{c1} For each $\ga\in\ddt(sl(n,C),K)$, $ s_\ga\in W_G(K)$. \end{lem}
\bp Assume
$\ga=(a_1\bt_{l_1}+b_1\al_{l_1},\cdots,a_k\bt_{l_k}+b_k\al_{l_k})$,
then $$Y_\ga =Q_{l_1}^{a_1}P_{l_1}^{-b_1}\ot\cdots\ot
Q_{l_k}^{a_k}P_{l_k}^{-b_k}.$$
By Corollary \ref{l} and (\ref{s}) there exists some
$Y_\dt=Q_{l_1}^{c_1}\ot\cdots\ot Q_{l_k}^{c_k}$ such that
$Ad_{Y_\ga}$ is conjugate to $Ad_{Y_\dt}$ under $N_G(K)$. Assume $
Y_\dt=Q_{l_1}^{c_1}\ot\cdots\ot Q_{l_k}^{c_k}$ as an element of
$\GL(n,\mathbb{C})$ has order $m$. Then $L_{i\dt}=\mb C Y_{\dt}^i$
for $i=1,\cdots,m-1$ and $L_{[\dt]}=\oplus_{i=1}^{m-1} \mb C
Y_\dt^i$ is an abelian Lie algebra consisting of semisimple
elements. We will show $\Ad_{Y_\dt}\in Z(\Ker \dt)_0$, then
$\Ad_{Y_\ga}\in Z(\Ker \ga)_0$ as $Ad_{Y_\ga}$ and $Ad_{Y_\dt}$ are
conjugate. Thus $\ga^*=(\Ad_{Y_\ga})^{-1}\in Z(\Ker \ga)_0$ and $
s_\ga\in W_G(K)$ by Lemma \ref{a6}.
The set $D_i$ of
eigenvalues of $Q_{l_i}^{c_i}$ is a cyclic group for each $i$. Let
$D$ be the set of eigenvalues of $Y_\dt$. For any $a,b\in D$,
$a=a_1\cdots a_k,\ b=b_1\cdots b_k$ with $a_i,b_i\in D_i$. Then
$ab^{-1}=(a_1 b_1^{-1})\cdots (a_k b_k^{-1})\in D$. So $D$ is also a
group. As the order of $Y_\dt$ is $m$, $D$ is a subgroup of the
cyclic group $C_m$. Then $D=C_m$ as the order of $Y_\dt$ is $m$.
Let $\omega=\omega_m$ then in some suitable basis of
$\mathbb{C}^{n}$,
$$Y_\dt=diag(1,\cdots,1,\omega,\cdots,\omega,\cdots\cdots,\omega^{m-1},\cdots,\omega^{m-1}),$$
where for $j=0,1,\cdots,m-1,$ there are $t_j$ copies of $\omega^j$
on the diagonal with each $t_j>0$. Let $s=\frac{2\pi i}{m}$ and
$$A=diag(0,\cdots,0,s,\cdots,s,\cdots\cdots,(m-1)s,\cdots,(m-1)s),$$ where for $j=0,1,\cdots,m-1$ there are $t_j$ copies of $js$ on the
diagonal. Then $exp(A)=Y_\dt$.
Let $D=(d_{i,j})_{m\times m}$ where $d_{i,j}=w^{ij}$ for
$i,j=0,1,\cdots,m-1$. As $D$ is invertible, there are unique complex
numbers $c_0,c_1,\cdots,c_{m-1}$ satisfying
$$D\cdot (c_0,c_1,\cdots,c_{m-1})^{t}=(0,s,\cdots,(m-1)s)^{t}.$$
Then
$\sum_{i=0}^{m-1}c_i Y_\dt^i=A$ and $exp(\sum_{i=0}^{m-1}c_i
Y_\dt^i)=Y_\dt.$ Then as $$\sum_{i=1}^{m-1}c_i\ {Y_\dt^i}\in
L_{[\dt]}\ \an \ \Lie Z(\Ker \dt)_0=\ad L_{[\dt]}$$ one has
$$\Ad_{Y_\dt}=exp(\ad(\sum_{i=1}^{m-1}c_i\ {Y_\dt^i}))\in Z(\Ker
\dt)_0.$$ \ep
Let $K$ be a finite maximal diagonalizable subgroup of
$G=\PGL(n,\mathbb{C})$. Recall that $K$ has a $W_G(K)$-invariant
anti-symmetric pairing $<,>$ and $(K,<,>)$ is a nonsingular
symplectic abelian group by Proposition \ref{b1}. Thus
$W_G(K)\st\Sp(K)$, where $\Sp(K)$ is the isometry group of
$(K,<,>)$.
\begin{theorem}\lb{d7}
The Weyl group $W_G(K)$ equals $\Sp(K)$, and is generated by
the set of transvections $ s_\ga$ with $\ga\in\ddt(sl(n,\mb C),K)$.
\end{theorem}
\bp By Lemma \ref{c1} one has $s_\ga\in W_G(K)$ for each
$\ga\in\ddt(sl(n,C),K)$. As $\ddt(sl(n,C),K)=\wh K\setminus \{0\}$,
$W_G(K)$ contains all $ s_\si$ with $\si$ a nonidentity element in
$K$. By Theorem \ref{c3} all such $ s_\si$ generate $\Sp(K)$, thus
$W_G(K)=\Sp(K)$.
\ep
|
\section{Simple facts about the unitary gas}
\label{sec:sf}
\subsection{What is the unitary gas ?}
First, the unitary gas is $\ldots$ a gas. As opposed to a liquid,
it is a dilute system with respect to the interaction range $b$: its
mean number density $\rho$ satisfies the constraint
\begin{equation}
\label{eq:gaz}
\rho b^3 \ll 1.
\end{equation}
For a rapidly decreasing interaction potential $V(r)$, $b$ is the spatial
width of $V(r)$. In atomic physics, where $V(r)$ may be viewed
as a strongly repulsive core and a Van der Waals attractive tail $-C_6/r^6$,
one usually assimilates $b$ to the Van der Waals length
$(m C_6/\hbar^2)^{1/4}$.
The intuitive picture of a gas is that the particles mainly experience
binary scattering, the probability that more than two particles
are within a volume $b^3$ being negligible. As a consequence,
what should really matter is the knowledge of the scattering amplitude
$f_k$ of two particles, where $k$ is the relative momentum,
rather than the $r$ dependence of the interaction potential $V(r)$.
This expectation has
guided essentially all many-body works on the BEC-BCS crossover:
One uses convenient models for $V(r)$ that are very different
from the true atomic interaction potential, but that reproduce
correctly the momentum dependence of $f_k$ at the relevant low values
of $k$, such as the Fermi momentum or the inverse thermal de Broglie wavelength,
these relevant low values of $k$ having to satisfy $k b \ll 1$ for this
modelization to be acceptable.
Second, the unitary gas is such that, for the relevant values of
the relative momentum $k$, the modulus of $f_k$ reaches the maximal
value allowed by quantum mechanics, the so-called unitary limit
\cite{livre_collisions}. Here, we consider $s$-wave scattering
between two opposite-spin fermions, so that $f_k$ depends
only on the modulus of the relative momentum. The optical
theorem, a consequence of the unitarity of the quantum evolution operator
\cite{livre_collisions}, then implies
\begin{equation}
\mathrm{Im}\, f_k = k |f_k|^2.
\end{equation}
Dividing by $|f_k|^2$, and using $f_k/|f_k|^2=1/f_k^*$,
one sees that this fixes the value of the imaginary part of
$1/f_k$, so that it is strictly equivalent to the requirement that
there exists a real function $u(k)$ such that
\begin{equation}
\label{eq:u(k)}
f_k = - \frac{1}{ik+u(k)}
\end{equation}
for all values of $k$. We then obtain the upper bound $|f_k| \leq 1/k$.
Ideally, the unitary gas saturates this inequality for all values
of $k$:
\begin{equation}
\label{eq:fku}
f_k^{\rm unitary} = -\frac{1}{ik}.
\end{equation}
In reality, Eq.(\ref{eq:fku}) cannot hold for all $k$.
It is thus important to understand over which range of $k$
Eq.(\ref{eq:fku}) should hold to have a unitary gas, and to estimate
the deviations from Eq.(\ref{eq:fku}) in that range in a real
experiment.
To this end, we use the usual low-$k$ expansion of the denominator
of the scattering amplitude \cite{livre_collisions}, under validity
conditions specified in \cite{math_re}:
\begin{equation}
\label{eq:lke}
u(k) = \frac{1}{a} - \frac{1}{2} r_e k^2 + \ldots
\end{equation}
The length $a$ is the scattering length, the length $r_e$ is the effective
range of the interaction. Both $a$ and $r_e$ can be of arbitrary sign.
Even for $1/a=0$, even for an everywhere non-positive interaction potential,
$r_e$ can be of arbitrary sign. As this last property seems to contradict
a statement in the solution of problem 1
in \S 131 of \cite{Landau},
we have constructed an explicit example depicted in Fig.~\ref{fig:contre},
which even shows that the effective range may be very
different in absolute value from the true potential range $b$, i.e.\
$r_e/b$ for $a^{-1}=0$
may be in principle an arbitrarily large and negative number.
Let us assume that the $\ldots$ in Eq.(\ref{eq:lke}) are negligible if
$k b \ll 1$, an assumption that will be revisited in \S\ref{subsubsec:nccm}.
Noting $k_{\rm typ}$ a typical relative momentum in the gas, we thus
see that the unitary gas is in practice obtained as a double limit,
a {\sl zero range} limit
\begin{equation}
\label{eq:zrl}
k_{\rm typ} b \ll 1, k_{\rm typ} |r_e| \ll 1
\end{equation}
and an {\sl infinite scattering length} limit:
\begin{equation}
\label{eq:isll}
k_{\rm typ} |a| \gg 1.
\end{equation}
\begin{figure}[b]
\centerline{\includegraphics[width=9cm,clip=]{contre.eps}}
\caption{A class of non-positive
potentials (of compact support of radius $b$)
that may lead to a negative effective range in the resonant
case $a^{-1}=0$. The resonant case is achieved when the three parameters
$\alpha,\beta$ and $\epsilon$ satisfy
$\tan[(1-\epsilon)\alpha] \tan(\epsilon \beta) = \alpha/\beta.$
Then from Smorodinskii's formula, see Problem 1 in \S 131 of \cite{Landau},
one sees that $r_e/b \leq 2$. One also finds that
$r_e/b\sim -\cos^2\theta/(\pi\epsilon)^2
\to -\infty$ when $\epsilon\to 0$ with
$\alpha=\pi$, $\beta\epsilon \to \theta$, where $\theta=2.798386\ldots$
solves $1+\theta\tan\theta=0$.
}
\label{fig:contre}
\end{figure}
At zero temperature, we assume that $k_{\rm typ}= k_F$, where
the Fermi momentum is conventionally defined in terms of the gas
total density $\rho$ as for the ideal spin-1/2 Fermi gas:
\begin{equation}
\label{eq:kf}
k_F = (3\pi^2\rho)^{1/3}.
\end{equation}
In a trap, $\rho$ and thus $k_F$ are position dependent.
Condition~(\ref{eq:isll}) is well satisfied experimentally, thanks
to the Feshbach resonance. The condition $k_F b \ll 1$
is also well satisfied at the per cent level, because
$b \approx$ the Van der Waals length is in the nanometer range.
Up to now, there is no experimental tuning of the effective range $r_e$, and there
are cases where $k_F |r_e|$ is not small. However, to study
the BEC-BCS crossover, one uses in practice
the so-called broad Feshbach resonances, which do not require
a too stringent control of the spatial homogeneity of the
magnetic field, and where $|r_e|\sim b$; then Eq.(\ref{eq:zrl}) is also
satisfied.
We note that the assumption $k_{\rm typ}=k_F$, although quite intuitive,
is not automatically correct. For example, for bosons, as shown
by Efimov \cite{Efimov70}, an effective three-body attraction takes place,
leading to the occurrence of the Efimov trimers; this attraction leads
to the so-called problem of {\sl fall to the center} \cite{Landau},
and one has $1/k_{\rm typ}$ of the order of the largest of the
two ranges $b$ and $|r_e|$. Eq.(\ref{eq:zrl}) is then violated,
and an extra length scale, the three-body parameter, has to be introduced,
breaking the scale invariance of the unitary gas.
Fortunately, for three fermions,
there is no Efimov attraction, except for the case
of different masses for the two spin components:
If two fermions of mass $m_\uparrow$ interact with a lighter particle
of mass $m_\downarrow$, the Efimov effect takes place
for $m_\uparrow / m _\downarrow$ larger than $\simeq 13.607$
\cite{Efimov73,Petrov}.
If a third fermion of mass $m_\uparrow$ is added, a four-body Efimov effect appears at a
slightly lower mass ratio $m_\uparrow / m _\downarrow\simeq13.384$~\cite{CMP}.
In what follows we consider the case of equal masses, unless specified otherwise.
At non-zero temperature $T>0$,
another length scale appears in the unitary gas properties,
the thermal de Broglie wavelength $\lambda_{\rm dB}$, defined as
\begin{equation}
\label{eq:ldb}
\lambda_{\rm dB}^2 = \frac{2\pi \hbar^2}{m k_B T}.
\end{equation}
At temperatures larger than the Fermi temperature $T_F
= \hbar^2 k_F^2/(2mk_B)$, one has to take $k_{\rm typ}\sim 1/\lambda_{\rm dB}$
in the conditions~(\ref{eq:zrl},\ref{eq:isll}).
In practice, the most interesting regime is however the degenerate
regime $T<T_F$, where the non-zero temperature does not bring
new conditions for unitarity.
\subsection{Some simple properties of the unitary gas}
\label{subsec:ssfp}
As is apparent in the expression of the two-body
scattering amplitude Eq.(\ref{eq:fku}), there is no parameter or
length scales issuing from the interaction.
As a consequence, for a gas in the trapping potential $U(\mathbf{r})$, the
eigenenergies $E_i$ of the $N$-body problem only depend on $\hbar^2/m$
and on the spatial dependence of $U(\mathbf{r})$:
the length scale required to get an energy out
of $\hbar^2/m$ is obtained from the shape of the container.
This is best formalized in terms of a spatial scale invariance.
Qualitatively, if one changes the volume of the container, even if the gas
becomes arbitrarily dilute, it remains
at unitarity and strongly interacting. This is of course
not true for a finite value of the scattering length $a$:
If one reduces the gas density, $\rho^{1/3} a$ drops eventually to small
values, and the gas becomes weakly interacting.
Quantitatively, if one applies to the container
a similarity factor $\lambda$ in all directions, which
changes its volume from $V$ to $\lambda^3 V$, we expect
that each eigenenergy scales as
\begin{equation}
\label{eq:sca_ener}
E_i \rightarrow \frac{E_i}{\lambda^2}
\end{equation}
and each eigenwavefunction scales as
\begin{equation}
\label{eq:sca_psi}
\psi_i(\mathbf{X}) \rightarrow \frac{\psi_i(\mathbf{X}/\lambda)}{\lambda^{3N/2}}.
\end{equation}
Here $\mathbf{X}=(\mathbf{r}_1,\ldots,\mathbf{r}_N)$ is the set of all coordinates of the particles,
and the $\lambda$-dependent factor ensures that the wavefunction
remains normalized. The properties~(\ref{eq:sca_ener},\ref{eq:sca_psi}),
which are at the heart of what the unitary gas really is,
will be put on mathematical grounds
in section \ref{sec:vm} by replacing the interaction
with contact conditions on $\psi$.
Simple consequences may be obtained from these scaling properties, as we now discuss.
In a harmonic isotropic trap, where a single particle has an
oscillation angular frequency $\omega$, taking as the scaling factor
the harmonic oscillator length $a_{\rm ho}=[\hbar/(m\omega)]^{1/2}$,
one finds that
\begin{equation}
\label{eq:simple}
\frac{E_i}{\hbar\omega} = \mathcal{F}_i(N)
\end{equation}
where the functions $\mathcal{F}_i$ are universal functions, ideally independent
of the fact that one uses lithium 6 or potassium 40 atoms, and depending
only on the particle number.
In free space, the unitary gas cannot have a $N$-body bound state (an eigenstate
of negative energy), whatever the value of $N\geq 2$.
If there was such a bound state, which corresponds to a square integrable
eigenwavefunction of the relative (Jacobi) coordinates of the particles,
one could generate a continuum of such square integrable eigenstates
using Eqs.(\ref{eq:sca_ener},\ref{eq:sca_psi}). This would violate
a fundamental property of self-adjoint Hamiltonians \cite{analyse_spectrale}.
Another argument is that the energy of a discrete universal bound state would depend only on $\hbar$ and $m$,
which is impossible by dimensional analysis.
At thermal equilibrium in the canonical ensemble in a box,
say a cubic box of volume $V=L^3$ with periodic boundary conditions,
several relations may be obtained if one takes the thermodynamic
limit $N\to +\infty$, $L^3\to +\infty$ with a fixed density $\rho$
and temperature $T$, and if one assumes that the free energy $F$
is an extensive quantity. Let us consider for simplicity
the case of equal population of the two spin states,
$N_\uparrow=N_\downarrow$. Then, in the thermodynamic
limit, the free energy per
particle $F/N\equiv f$ is a function of the density $\rho$ and
temperature $T$. If one applies a similarity of factor $\lambda$
and if one change $T$ to $T/\lambda^2$ so as to keep a constant
ratio $E_i/(k_B T)$, that is a constant occupation probability
for each eigenstate, one obtains from Eq.(\ref{eq:sca_ener}) that
\begin{equation}
\label{eq:sca_f}
f(\rho/\lambda^3, T/\lambda^2) = f(\rho, T)/\lambda^2.
\end{equation}
At zero temperature, $f$ reduces to the ground state energy per
particle $e_0(\rho)$. From Eq.(\ref{eq:sca_f}) it appears that
$e_0(\rho)$ scales as $\rho^{2/3}$, exactly as the ground state energy
of the ideal Fermi gas. One thus simply has
\begin{equation}
\label{eq:e0}
e_0(\rho) = \xi \, e_0^{\rm ideal}(\rho) = \frac{3\xi}{5}
\frac{\hbar^2 k_F^2}{2m}
\end{equation}
where $k_F$ is defined by Eq.(\ref{eq:kf}) and
$\xi$ is a universal number. This is also a simple consequence
of dimensional analysis \cite{Ho}. Taking the derivative with respect
to $N$ or to the volume,
this shows that the same type of relation holds for the zero temperature
chemical potential, $\mu_0(\rho) = \xi \mu_0^{\rm ideal}(\rho)$,
and for the zero temperature total pressure,
$P_0(\rho) = \xi P_0^{\rm ideal}(\rho)$,
so that
\begin{eqnarray}
\label{eq:mu0}
\mu_0(\rho) &=& \xi \frac{\hbar^2 k_F^2}{2m}\\
P_0(\rho) &=& \frac{2\xi}{5} \rho \frac{\hbar^2 k_F^2}{2m}.
\label{eq:P0}
\end{eqnarray}
At non-zero temperature, taking the derivative of Eq.(\ref{eq:sca_f})
with respect to $\lambda$ in $\lambda=1$, and using $F=E-TS$, where
$E$ is the mean energy and $S=-\partial_T F$ is the entropy,
as well as $\mu=\partial_N F$, one obtains
\begin{equation}
\frac{5}{3} E - \mu N = T S.
\end{equation}
From the Gibbs-Duhem relation, the grand potential
$\Omega=F-\mu N$ is equal to $-P V$, where $P$ is the pressure
of the gas. This gives finally the useful relation
\begin{equation}
\label{eq:pv}
PV = \frac{2}{3} E,
\end{equation}
that can also be obtained from dimensional analysis \cite{Ho}, and that
of course also holds at zero temperature (see above).
All these properties actually also apply to the ideal Fermi gas,
which is obviously scaling invariant. The relation (\ref{eq:pv})
for example was established for the ideal gas in \cite{ideal}.
Let us finally describe at a macroscopic level, i.e.\ in a hydrodynamic picture,
the effect of the similarity Eq.(\ref{eq:sca_psi}) on the quantum state of a unitary gas,
assuming that it was initially at thermal equilibrium in a trap.
In the initial state of the gas, consider a small (but still macroscopic) element, enclosed in a volume $dV$ around point $\mathbf{r}$.
It is convenient to assume that $dV$ is a fictitious cavity with periodic boundary conditions.
In the hydrodynamic picture, this small element is assumed to be at local thermal equilibrium
with a temperature $T$. Then one performs the spatial scaling transform Eq.(\ref{eq:sca_ener})
on each many-body eigenstate $\psi$ of the thermal statistical mixture,
which does not change the statistical weigths.
How will the relevant physical quantities be transformed in the hydrodynamic approach~?
The previously considered small element is now at position $\lambda \mathbf{r}$, and occupies a volume $\lambda^3 dV$,
with the same number of particles. The hydrodynamic mean density profile after rescaling,
$\rho_\lambda$, is thus related to the mean density profile $\rho$ before scaling as
\begin{equation}
\label{eq:resca_macro_dens}
\rho_\lambda(\lambda \mathbf{r}) = \rho(\mathbf{r})/\lambda^3.
\end{equation}
Second, is the small element still at (local) thermal equilibrium after scaling~?
Each eigenstate of energy $E_{\rm loc}$ of the locally homogeneous unitary gas within the initial cavity of
volume $dV$ is transformed by the scaling
into an eigenstate within the cavity of volume $\lambda^3 dV$, with the eigenenergy $E_{\rm loc}/\lambda^2$.
Since the occupation probabilities of each local eigenstate
are not changed, the local statistical mixture remains thermal provided that one rescales
the temperature as
\begin{equation}
\label{eq:resca_macro_temp}
T_\lambda = T/\lambda^2.
\end{equation}
A direct consequence is that the entropy of the small element of the gas is unchanged by the scaling,
so that the local entropy {\sl per particle} $s$ in the hydrodynamic approach obeys
\begin{equation}
\label{eq:resca_macro_entropie_par_particule}
s_\lambda(\lambda \mathbf{r}) = s(\mathbf{r}).
\end{equation}
Also, since the mean energy of the small element is reduced by the factor $\lambda^2$ due to the scaling,
and the volume of the small element is multiplied by $\lambda^3$, the equilibrium relation Eq.(\ref{eq:pv}) imposes
that the local pressure is transformed by the scaling as
\begin{equation}
\label{eq:resca_macro_pression}
p_\lambda(\lambda\mathbf{r}) = p(\mathbf{r})/\lambda^5.
\end{equation}
\subsection{Application: Inequalities on $\xi$ and finite-temperature
quantities}
Using the previous constraints imposed by scale invariance of the unitary
gas on thermodynamic quantities, in addition to standard
thermodynamic inequalities, we show that one can produce constraints
involving both the zero-temperature quantity $\xi$ and finite-temperature
quantities of the gas.
Imagine that, at some temperature $T$, the energy $E$ and the chemical
potential $\mu$ of the non-polarized unitary Fermi gas have been
obtained, in the thermodynamic limit.
If one introduces the
Fermi momentum
Eq.(\ref{eq:kf}) and the corresponding Fermi energy
$E_F=\hbar^2 k_F^2/(2m)$, this means that on has at hand the two
dimensionless quantities
\begin{eqnarray}
A &\equiv& \frac{E}{N E_F} \\
B &\equiv& \frac{\mu}{E_F}.
\end{eqnarray}
As a consequence of Eq.(\ref{eq:pv}), one also has access to the
pressure $P$.
We now show that the following inequalities hold at any temperature $T$:
\begin{equation}
\label{eq:xib}
\left(\frac{3}{5A}\right)^{2/3} B^{5/3} \leq
\xi \leq \frac{5A}{3}.
\end{equation}
In the canonical ensemble, the
mean energy
$E(N,T,V)$ is an increasing function of temperature for fixed
volume $V$ and atom number $N$.
Indeed one has the well-known relation
$k_B T^2 \partial_T E(N,T,V) = \mathrm{Var}\, H$,
and the variance of the Hamiltonian is non-negative.
As a consequence, for any temperature $T$:
\begin{equation}
E(N,T,V) \ge E(N,0,V).
\end{equation}
From Eq.(\ref{eq:e0}) we then reach the upper bound on $\xi$ given
in Eq.(\ref{eq:xib}).
In the grand canonical ensemble, the
pressure $P(\mu,T)$ is
an increasing function of temperature for a fixed chemical potential.
This results from the Gibbs-Duhem relation
$\Omega(\mu,T,V)=-V P(\mu,T)$ where $\Omega$ is the grand potential
and $V$ the volume, and from the differential relation
$\partial_T \Omega(\mu,T)=-S$ where $S\geq 0$ is the
entropy.
As a consequence, for any temperature $T$:
\begin{equation}
P(\mu,T) \geq P(\mu,0).
\end{equation}
For the unitary gas,
the left hand side can be expressed in terms of $A$ using (\ref{eq:pv}).
Eliminating the density between Eq.(\ref{eq:mu0}) and Eq.(\ref{eq:P0})
we obtain the zero temperature pressure
\begin{equation}
P(\mu,0) = \frac{1}{15\pi^2\xi^{3/2}} \frac{\hbar^2}{m}
\left(\frac{2m\mu}{\hbar^2}\right)^{5/2}.
\end{equation}
This leads to the lower bound on $\xi$ given in Eq.(\ref{eq:xib}).
Let us apply Eq.(\ref{eq:xib}) to the Quantum Monte Carlo results
of \cite{Burovski}: At the critical temperature $T=T_c$,
$A=0.310(10)$ and $B=0.493(14)$, so that
\begin{equation}
0.48(3) \leq \xi_{\cite{Burovski}} \leq 0.52(2).
\end{equation}
This deviates by two standard deviations from the fixed node
result $\xi \leq 0.40(1)$ \cite{Carlson}.
The Quantum Monte Carlo results of \cite{Bulgac},
if one takes a temperature equal to the critical temperature of \cite{Burovski},
give
$A=0.45(1)$ and $B=0.43(1)$; these values, in clear disagreement
with \cite{Burovski}, lead to the non-restrictive bracketing
$0.30(2)\leq \xi_{\cite{Bulgac}}\leq 0.75(2)$.
The more recent work~\cite{Goulko} finds
$k_BT_c/E_F=0.171(5)$
and at this critical temperature,
$A=0.276(14)$ and $B=0.429(9)$,
leading to
\begin{equation}
0.41(3) \leq \xi_{\cite{Goulko}} \leq 0.46(2).
\end{equation}
Another, more graphical application of our simple bounds
is to assume some reasonable value of $\xi$, and then to use
Eq.(\ref{eq:xib}) to construct a zone in the energy-chemical potential
plane that is forbidden at all temperatures. In Fig.\ref{fig:exclue}, we took $\xi=0.41$, inspired
by the fixed node upper bound on the exact value of $\xi$ \cite{Carlson}:
The shaded area is the resulting forbidden zone, and the filled disks with error bars represent the in principle
exact Quantum Monte Carlo results of various groups at $T=T_c$. The prediction
of \cite{Burovski} lies within the forbidden zone.
The prediction of \cite{Bulgac} is well within the allowed zone, whereas the most recent
prediction of \cite{Goulko} is close to the boundary between the forbidden and the allowed zones.
If one takes a smaller value for $\xi$,
the boundaries of the forbidden zone will shift as indicated by the arrows on the figure.
\begin{comment}
The discrepancies between these three data points, as well as the fact that the result of~\cite{Burovski} violates thermodynamic equalities, has to be taken with a grain of salt: All these calculations used lattice models with not-very-small filling factors (i.e. not-very-small values of $k_F b$), because in the employed Monte-Carlo methods the computational cost increses when the filling factor decreases.
The data of~\cite{Burovski} and~\cite{Goulko} actually seem consistent at the finite fillings where the simulations were done; but fitting the dependence of these data on filling factor and extrapolating them to zero filling amplifies the rather small differences, and the error bar for the extrapolated result was difficult to estimate a priori since it has to includes the systematic uncertainty coming from the extrapolation procedure. In~\cite{Bulgac}, this extrapolation was not done.
\end{comment}
All this shows that simple reasonings may be useful to test and guide
numerical studies of the unitary gas.
\begin{figure}[b]
\centerline{\includegraphics[width=9cm,clip=]{exclue.eps}}
\caption{For the spin balanced uniform unitary gas at thermal equilibrium: Assuming $\xi=0.41$ in Eq.(\ref{eq:xib})
defines a zone (shaded in gray) in the plane energy-chemical potential that is forbidden
at all temperatures.
The black disks correspond to the unbiased Quantum Monte Carlo results
of Burovski et al. \cite{Burovski}, of Bulgac et al. \cite{Bulgac},
and of Goulko et al. \cite{Goulko} at the critical temperature. Taking the unknown exact value of $\xi$, which is below the
fixed node upper bound 0.41 \cite{Carlson}, will shift the forbidden zone boundaries
as indicated by the arrows.
}
\label{fig:exclue}
\end{figure}
\subsection{Is the unitary gas attractive or repulsive ?}
\label{subsec:aor}
According to a common saying, a weakly interacting Fermi gas
($k_F |a|\ll 1$)
experiences an effective repulsion for a positive scattering length $a>0$,
and an effective attraction for a negative scattering length $a<0$.
Another common fact is that, in the unitary limit
$|a|\to +\infty$, the gas properties do not depend on the sign
of $a$. As the unitary limit may be apparently equivalently obtained
by taking the limit $a\to +\infty$ or the limit $a\to -\infty$,
one reaches a paradox, considering the fact that the unitary
gas does not have the same ground state energy than the ideal gas
and cannot be at the same time an attractive and repulsive state of matter.
This paradox may be resolved by considering the case of two
particles in an isotropic harmonic trap.
After elimination of the center of mass motion, and restriction
to a zero relative angular momentum to have $s$-wave interaction,
one obtains the radial Schr\"odinger equation
\begin{equation}
\label{eq:schr3d}
-\frac{\hbar^2}{2\mu} \left[\psi''(r) + \frac{2}{r} \psi'(r)\right]
+\frac{1}{2} \mu \omega^2 r^2 \psi(r) = E_{\rm rel} \psi(r),
\end{equation}
with the relative mass $\mu=m/2$. The interactions
are included in the zero range limit by the $r=0$ boundary conditions,
the so-called Wigner-Bethe-Peierls contact conditions described in section
\ref{sec:vm}:
\begin{equation}
\label{eq:bp}
\psi(r) = A [r^{-1}-a^{-1}] + O(r)
\end{equation}
that correctly reproduce the free space scattering amplitude
\begin{equation}
\label{eq:fkzr}
f_k^{\rm zero\ range} = -\frac{1}{a^{-1} + ik}.
\end{equation}
The general solution of Eq.(\ref{eq:schr3d}) may be expressed in
terms of Whittaker $M$ et $W$ functions. For an energy $E_{\rm rel}$
not belonging to the non-interacting spectrum $\{(\frac{3}{2}+2n)\hbar
\omega,n\in \mathbb{N}\}$, the Whittaker function $M$ diverges
exponentially for large $r$ and has to be disregarded. The small $r$
behavior of the Whittaker function $W$, together with the Wigner-Bethe-Peierls
contact condition, leads to the implicit equation for the relative
energy, in accordance with \cite{Wilkens}:
\begin{equation}
\label{eq:impl}
\frac{\Gamma(\frac{3}{4}- \frac{E_{\rm rel}}{2\hbar\omega})}
{\Gamma(\frac{1}{4}- \frac{E_{\rm rel}}{2\hbar\omega})}
= \frac{a_{\rm ho}^{\rm rel}}{2a}
\end{equation}
with the harmonic oscillator length of the relative motion,
$a_{\rm ho}^{\rm rel} = [\hbar/(\mu\omega)]^{1/2}$.
The function $\Gamma(x)$ is different from zero $\forall x\in \mathbb{R}$
and diverges on each non-positive integers.
Thus Eq.(\ref{eq:impl}) immediately leads in the unitary case to the spectrum
$E_{\rm rel}\in \{(2n+1/2)\hbar \omega,n\in\mathbb{N}\}$.
This can be readily obtained by setting in Eq.(\ref{eq:schr3d})
$\psi(r)=f(r)/r$, so that $f$ obeys Schr\"odinger's equation
for a 1D harmonic oscillator, with the constraint issuing
from Eq.(\ref{eq:bp}) that $f(r=0)\neq 0$, which selects the even 1D
states.
The graphical solution of Eq.(\ref{eq:impl}), see Fig.~\ref{fig:spec3d},
allows to resolve the paradox about the attractive or repulsive
nature of the unitary gas. E.g. starting with the ground state
wavefunction of the ideal gas case, of relative energy $E_{\rm rel}
=\frac{3}{2}\hbar \omega$, it appears that the two adiabatic followings
(i) $a=0^+\rightarrow a= +\infty$ and (ii) $a=0^-\rightarrow -\infty$
lead to {\sl different} final eigenstates of the unitary case,
to an excited state $E_{\rm rel}=\frac{5}{2}\hbar\omega$ for the procedure
(i), and to the ground state $E_{\rm rel}=\frac{1}{2}\hbar\omega$
for procedure (ii).
\begin{figure}[b]
\sidecaption
\includegraphics[width=7cm,clip=]{spec3d.eps}
\caption{For the graphical solution of Eq.(\ref{eq:impl}),
which gives the spectrum for two particles in a three-dimensional
isotropic harmonic trap, plot of the function
$f_{3D}(x) =
\Gamma(\frac{3}{4}- \frac{x}{2})/
\Gamma(\frac{1}{4}- \frac{x}{2})$, where $x$
stands for $E_{\rm rel}/(\hbar\omega)$.
}
\label{fig:spec3d}
\end{figure}
The same explanation holds for the many-body case: The interacting
gas has indeed several energy branches in the BEC-BCS crossover,
as suggested by the toy model
\footnote{This toy model replaces the many-body
problem with the one of a matterwave interacting with a single scatterer
in a hard wall cavity of radius $\propto 1/k_F$.}
of \cite{PricoupenkoToy}, see Fig.~\ref{fig:toy}.
Starting from the weakly attractive Fermi gas
and ramping the scattering length down to $-\infty$ one explores
a part of the ground energy branch, where the unitary gas is attractive;
this ground branch continuously evolves into a weakly repulsive
condensate of dimers \cite{PetrovShlyapSalomon}
if $1/a$ further moves from $0^-$ to $0^+$ and then to $+\infty$.
The attractive nature of the unitary gas on the ground energy branch
will become apparent in the lattice model of section \ref{sec:vm}.
On the other hand, starting from the weakly repulsive Fermi gas and
ramping the scattering up to $+\infty$, one explores an effectively
repulsive excited branch.
In the first experiments on the BEC-BCS crossover,
the ground branch was explored by adiabatic
variations of the scattering length and was found to be stable.
The first excited energy branch
was also investigated in the early work~\cite{Bourdel_Eint}, and more recently in~ \cite{Ketterle_excited} looking for a Stoner demixing instability
of the strongly repulsive two-component Fermi gas.
A difficulty for the study of this excited branch is its metastable
character: Three-body collisions gradually transfer
the gas to the ground branch, leading e.g. to the formation
of dimers if $0 < k_F a \lesssim 1$.
\begin{figure}[b]
\sidecaption
\includegraphics[width=7cm,clip=]{toy.eps}
\caption{
In the toy model of \cite{PricoupenkoToy},
for the homogeneous two-component unpolarized Fermi gas,
energy per particle on the ground branch and the first excited branch
as a function of the inverse scattering length.
The Fermi wavevector is defined in Eq.(\ref{eq:kf}), $E_F=\hbar^2 k_F^2/(2m)$
is the Fermi energy, and $a$ is the scattering length.
}
\label{fig:toy}
\end{figure}
\subsection{Other partial waves, other dimensions}
We have previously considered the two-body scattering amplitude
in the $s$-wave channel. What happens for example in the $p$-wave
channel~? This channel is relevant for the interaction between
fermions in the same internal state, where a Feshbach resonance
technique is also available \cite{Salomon_p,Jin_p}. Can one
also reach the unitarity limit Eq.(\ref{eq:fku}) in the $p$-wave channel~?
Actually the optical theorem shows that
relation Eq.(\ref{eq:u(k)}) also holds for the $p$-wave scattering
amplitude $f_k$. What differs is the low-$k$ expansion of $u(k)$,
that is now given by
\begin{equation}
\label{eq:ukp}
u(k)= \frac{1}{k^2\mathcal{V}_s} + \alpha + \ldots,
\end{equation}
where $\mathcal{V}_s$ is the scattering volume (of arbitrary sign)
and $\alpha$ has the dimension of the inverse of
a length.
The unitary limit would require $u(k)$ negligible as compared
to $k$. One can in principle tune $\mathcal{V}_s$ to infinity
with a Feshbach resonance. Can one then have a small value of $\alpha$
at resonance~? A theorem for a compact support interaction
potential of radius $b$ shows however that \cite{LudoPRL_ondeP,Jona}
\begin{equation}
\lim_{|\mathcal{V}_s|\to +\infty} \alpha \geq 1/b.
\end{equation}
A similar conclusion holds using two-channel models of the Feshbach
resonance \cite{Jona,Chevy}. $\alpha$ thus assumes a huge positive
value on resonance, which breaks the scale invariance and precludes
the existence of a $p$-wave unitary gas.
This does not prevent however to reach the unitary
limit in the {\sl vicinity} of a particular value of $k$.
For $\mathcal{V}_s$ large and negative, neglecting
the $\ldots$ in Eq.(\ref{eq:ukp}) under the condition $k b \ll 1$,
one indeed has $|u(k)|\ll k$,
so that $f_k \simeq -1/(ik)$, in a vicinity of
\begin{equation}
k_0 = \frac{1}{(\alpha |\mathcal{V}_s|)^{1/2}}.
\end{equation}
Turning back to the interaction in the $s$-wave channel, an interesting
question is whether the unitary gas exists in reduced dimensions.
In a one-dimensional system the zero range interaction may be
modeled by a Dirac potential $V(x)=g\delta(x)$. If $g$ is finite,
it introduces a length scale $\hbar^2/(m g)$ that breaks the scaling
invariance. Two cases are thus scaling invariant, the ideal gas
$g=0$ and the impenetrable case $1/g=0$. The impenetrable case
however is mappable to an ideal gas in one dimension, it has in particular
the same energy spectrum and thermodynamic properties \cite{Gaudin}.
In a two-dimensional system, the scattering amplitude for a zero
range interaction potential is given by \cite{Olshanii2D}
\begin{equation}
\label{eq:fk2D}
f_k^{2D} = \frac{1}{\gamma+\ln(k a_{2D}/2) -i\pi/2}
\end{equation}
where $\gamma=0.57721566\ldots$
is Euler's constant and $a_{2D}$ is the scattering length.
For a finite value of $a_{2D}$, there is no scale invariance.
The case $a_{2D}\to 0$ corresponds to the ideal gas limit. At first sight,
the opposite limit $a_{2D}\to +\infty$ is a good candidate for a
two-dimensional unitary gas; however this limit also corresponds
to an ideal gas. This appears in the 2D version of the lattice
model of section \ref{sec:vm} \cite{Tonini}.
This can also be checked for two particles in an isotropic
harmonic trap. Separating out the center of mass motion,
and taking a zero angular momentum state for the relative motion,
to have interaction in the $s$-wave channel, one has to solve
the radial Schr\"odinger equation:
\begin{equation}
-\frac{\hbar^2}{2\mu} [\psi''(r) + \frac{1}{r} \psi'(r)]
+\frac{1}{2}\mu \omega^2 r^2 \psi(r) = E_{\rm rel} \psi(r)
\label{eq:rad_schr}
\end{equation}
where $\mu=m/2$ is the reduced mass of the two particles,
$E_{\rm rel}$ is an eigenenergy of the relative motion,
and $\omega$ is the single particle angular oscillation frequency.
The interactions are included by the boundary condition in $r=0$:
\begin{equation}
\psi(r) = A\ln(r/a_{2D}) + O(r),
\end{equation}
which is constructed to reproduce the expression of the
scattering amplitude Eq.(\ref{eq:fk2D}) for the free space problem.
The general solution of Eq.(\ref{eq:rad_schr}) may be expressed
in terms of Whittaker functions $M$ and $W$.
Assuming that $E_{\rm rel}$ does not belong to the ideal gas spectrum
$\{(2n+1)\hbar \omega,n\in \mathbb{N}\}$, one finds that the
$M$ solution has to be disregarded because it diverges
exponentially for $r\to +\infty$. From the small $r$ behavior
of the $W$ solution, one obtains the implicit equation
\begin{equation}
\label{eq:impl2d}
\frac{1}{2} \psi\left(\frac{\hbar\omega-E_{\rm rel}}{2\hbar\omega}\right)
+\gamma = - \ln(a_{2D}/a_{\rm ho}^{\rm rel})
\end{equation}
where the relative harmonic oscillator length is $a_{\rm ho}^{\rm rel}
=[\hbar/(\mu\omega)]^{1/2}$ and the digamma function $\psi$
is the logarithmic derivative of the $\Gamma$ function.
If $a_{2D}\to +\infty$, one then finds that $E_{\rm rel}$
tends to the ideal gas spectrum $\{(2n+1)\hbar \omega,n\in \mathbb{N}\}$
from below, see Fig.~\ref{fig:spec2d}, in agreement with
the lattice model result that the 2D gas with a large and finite
$a_{2D}$ is a weakly attractive gas.
\begin{figure}[b]
\sidecaption
\includegraphics[width=7cm,clip=]{spec2d.eps}
\caption{For the graphical solution of Eq.(\ref{eq:impl2d}),
which gives the spectrum for two interacting particles in a two-dimensional
isotropic harmonic trap, plot of the function
$f_{2D}(x) = \frac{1}{2}
\psi[(1-x)/2] +\gamma$
where $x$ stands for $E_{\rm rel}/(\hbar\omega)$ and the special
function $\psi$
is the logarithmic derivative of the $\Gamma$ function.
}
\label{fig:spec2d}
\end{figure}
\section{Various models and general relations}
\label{sec:vm}
There are basically two approaches to model the interaction
between particles for the unitary gas (and more generally for
the BEC-BCS crossover).
In the first approach, see subsections \ref{subsec:lm} and \ref{subsec:tcm},
one takes a model with a finite range $b$
and a fixed (e.g. infinite) scattering length $a$. This model
may be in continuous space or on a lattice, with one or several channels.
Then one tries to calculate the eigenenergies, the thermodynamic properties from
the thermal density operator $\propto \exp(-\beta H)$,
etc, and the zero range limit $b\to 0$ should be taken at the end of the
calculation. Typically, this approach is followed in numerical
many-body methods, such as the approximate fixed node Monte Carlo
method \cite{Carlson,Panda,Giorgini} or unbiased Quantum Monte Carlo
methods \cite{Burovski,Bulgac,Juillet}.
A non-trivial question however is whether each eigenstate of the model
is universal in the zero range limit, that is if the eigenenergy
$E_i$ and the corresponding wavefunction $\psi_i$ converge
for $b\to 0$. In short, the challenge is to prove
that the ground state energy of the system does not tend
to $-\infty$ when $b\to 0$.
In the second approach, see subsection \ref{subsec:bp},
one directly considers the zero range limit,
and one replaces the interaction by the so-called Wigner-Bethe-Peierls
contact conditions on the $N$-body
wavefunction. This constitutes what we shall call the {\it zero-range model}.
The advantage is that only the scattering length
appears in the problem, without unnecessary details on the interaction,
which simplifies the problem and allows to obtain analytical results.
E.g. the scale invariance of the unitary gas becomes clear.
A non-trivial question however is to know whether the zero-range model
leads to a self-adjoint Hamiltonian, with a spectrum then necessarily
bounded from below for the unitary gas (see Section \ref{subsec:ssfp}), without having to add
extra boundary conditions.
For $N=3$ bosons,
due to the Efimov effect,
the Wigner-Bethe-Peierls or zero-range model becomes self-adjoint
only if one adds an extra three-body contact condition,
involving a
so-called three-body parameter.
In an isotropic
harmonic trap, at unitarity, there exists however a non-complete
family of bosonic universal states, independent from
the three-body parameter
and to which the restriction of the Wigner-Bethe-Peierls
model is hermitian \cite{Jonsell,WernerPRL}.
For equal mass two-component fermions, it is hoped in the physics
literature that the zero-range model is self-adjoint
for an arbitrary number of particles $N$.
Surprisingly, there exist works in mathematical physics
predicting
that this is
{\sl not} the case when $N$ is large enough \cite{Teta,Minlos};
however
the critical mass ratio for the appearance of an Efimov effect in the unequal-mass $3+1$ body problem given without proof in~\cite{Minlos}
was not confirmed by the numerical study~\cite{CMP},
and
the variational ansatz used in~\cite{Teta}
to show that the energy is unbounded below does not have the proper fermionic exchange symmetry.
This mathematical problem thus remains open.
\subsection{Lattice models and
general relations}
\label{subsec:lm}
\subsubsection{The lattice models}
The model that we consider here assumes that the spatial positions are discretized on a cubic lattice,
of lattice constant that we call $b$ as the interaction range.
It is quite appealing in its simplicity and generality.
It naturally allows to consider a contact interaction potential, opposite spin fermions
interacting only when they are on the same lattice site. Formally, this constitutes
a separable potential for the interaction (see subsection \ref{subsec:tcm} for a reminder),
a feature known to simplify diagrammatic calculations \cite{NSR}.
Physically, it belongs to the same class as the Hubbard model, so that it may truly
be realized with ultracold atoms in optical lattices \cite{BlochMott}, and it allows
to recover the rich lattice physics of condensed matter physics and the
corresponding theoretical tools such as Quantum Monte Carlo methods
\cite{Burovski,Juillet}.
The spatial coordinates $\mathbf{r}$ of the particles are thus discretized on a cubic
grid of step $b$. As a consequence, the components of the wavevector of a particle have
a meaning modulo $2\pi/b$ only, since the plane wave function $\mathbf{r}\rightarrow \exp(i\mathbf{k}\cdot\mathbf{r}\,)$
defined on the grid is not changed if a component of $\mathbf{k}$ is shifted by an integer
multiple of $2\pi/b$. We shall therefore restrict the wavevectors to the first Brillouin
zone of the lattice:
\begin{equation}
\mathbf{k} \in {\mathcal D}\equiv \left[-\frac{\pi}{b},\frac{\pi}{b}\right[^3.
\end{equation}
This shows that the lattice structure in real space automatically provides a cut-off in
momentum space.
In the absence of interaction and of confining potential, eigenmodes of the system are plane waves
with a dispersion relation $\mathbf{k}\to \epsilon_{\mathbf{k}}$, supposed to be an even and non-negative function of $\mathbf{k}$.
We assume that this dispersion relation is independent of the spin state, which is a natural choice
since the $\uparrow$ and $\downarrow$ particles have the same mass. To recover
the correct continuous space physics in the zero lattice spacing limit $b\to 0$, we further impose
that it reproduces the free space dispersion relation in that limit, so that
\begin{equation}
\label{eq:rdavdz}
\epsilon_{\mathbf{k}} \sim \frac{\hbar^2 k^2}{2m} \ \ \ \mbox{for}\ \ \ kb\to 0.
\end{equation}
The interaction between opposite spin particles takes place when two particles are
on the same lattice site, as in the Hubbard model. In first quantized form,
it is represented by a discrete delta potential:
\begin{equation}
V= \frac{g_0}{b^3} \delta_{\mathbf{r}_1,\mathbf{r}_2}.
\end{equation}
The factor $1/b^3$ is introduced because $b^{-3} \delta_{\mathbf{r},\mathbf{0}}$ is
equivalent to the Dirac distribution $\delta(\mathbf{r})$ in the continuous space limit.
To summarize, the lattice Hamiltonian in second quantized form in the general trapped case is
\begin{eqnarray}
H &=& \sum_{\sigma=\uparrow,\downarrow}
\int_{\mathcal{D}} \frac{d^3k}{(2\pi)^3}\,\epsilon_{\mathbf{k}} c_\sigma^\dagger(\mathbf{k})c_\sigma(\mathbf{k}) +
\sum_{\sigma=\uparrow,\downarrow}
\sum_{\mathbf{r}}
b^3 U(\mathbf{r}) (\psi_\sigma^\dagger \psi_\sigma)(\mathbf{r}) \nonumber \\
&& + g_0 \sum_\mathbf{r} b^3 (\psi^\dagger_\uparrow\psi^\dagger_\downarrow\psi_\downarrow\psi_\uparrow)(\mathbf{r}).
\end{eqnarray}
The plane wave annihilation operators $c_\sigma(\mathbf{k})$ in spin state $\sigma$
obey the usual continuous space anticommutation relations $\{c_\sigma(\mathbf{k}),c_{\sigma'}^\dagger(\mathbf{k}')\}=(2\pi)^3
\delta(\mathbf{k}-\mathbf{k}')\delta_{\sigma\sigma'}$ if $\mathbf{k}$ and $\mathbf{k}'$ are in the first Brillouin zone
\footnote{In the general case, $\delta(\mathbf{k}-\mathbf{k}')$ has to be replaced with $\sum_{\mathbf{K}} \delta(\mathbf{k}-\mathbf{k}'-\mathbf{K})$
where $\mathbf{K}\in (2\pi/b)\mathbb{Z}^3$ is any vector in the reciprocal lattice.},
and the field operators $\psi_{\sigma}(\mathbf{r})$ obey the usual discrete space
anticommutation relations $\{\psi_{\sigma}(\mathbf{r}),\psi_{\sigma'}^\dagger(\mathbf{r}')\}=
b^{-3}\delta_{\mathbf{r}\rr'} \delta_{\sigma\sigma'}$.
In the absence of trapping potential, in a cubic box with size $L$ integer multiple of $b$,
with periodic boundary conditions, the integral in the kinetic energy term is replaced by the sum
$\sum_{\mathbf{k}\in\mathcal{D}} \epsilon_{\mathbf{k}} \tilde{c}_{\mathbf{k}\sigma}^\dagger \tilde{c}_{\mathbf{k}\sigma}$
where the annihilation operators
then obey the discrete anticommutation relations $\{\tilde{c}_{\mathbf{k}\sigma},\tilde{c}^\dagger_{\mathbf{k}'\sigma'}\}=
\delta_{\mathbf{k}\kk'} \delta_{\sigma\sigma'}$ for $\mathbf{k},\mathbf{k}'\in \mathcal{D}$.
The coupling constant $g_0$ is a function of the grid spacing $b$. It is adjusted
to reproduce the scattering length of the true interaction.
The scattering amplitude of two atoms on the lattice with vanishing total momentum,
that is with incoming particles of opposite spin and opposite momenta $\pm\mathbf{k}_0$,
reads
\begin{equation}
\label{eq:fklm}
f_{k_0} = -\frac{m}{4\pi\hbar^2}
\left[g_0^{-1}-\int_{\mathcal D}\frac{d^3k}{(2\pi)^3}
\frac{1}{E+i0^+-2\epsilon_{\mathbf{k}}}\right]^{-1}
\end{equation}
as derived in details in \cite{Houches03} for a quadratic dispersion relation and in~\cite{Tangen} for a general dispersion relation.
Here the scattering state energy
$E=2\epsilon_{\mathbf{k}_0}$ actually introduces
a dependence of the scattering amplitude
on the direction of $\mathbf{k}_0$ when the dispersion relation $\epsilon_\mathbf{k}$
is not parabolic.
If one is only interested in the expansion of $1/f_{k_0}$
up to second order in $k_0$, e.g.\ for an effective range calculation,
one may conveniently use the isotropic approximation $E=\hbar^2 k_0^2/m$
thanks to~(\ref{eq:rdavdz}).
Adjusting $g_0$ to recover the correct scattering length gives
from Eq.(\ref{eq:fklm}) for $k_0\to 0$:
\begin{equation}
\frac{1}{g_0}= \frac{1}{g} -\int_{\mathcal D} \frac{d^3k}{(2\pi)^3} \,
\frac{1}{2 \epsilon_{\mathbf{k}}},
\label{eq:g0_gen}
\end{equation}
with $g=4\pi\hbar^2 a/m$. The above formula Eq.(\ref{eq:g0_gen}) is reminiscent of the technique of renormalization of the coupling
constant \cite{Randeria,Randeria2}.
A natural case to consider is the one of the usual parabolic dispersion relation,
\begin{equation}
\label{eq:pdr}
\epsilon_{\mathbf{k}}=\frac{\hbar^2 k^2}{2m}.
\end{equation}
A more explicit form of Eq.(\ref{eq:g0_gen}) is then \cite{Mora,LudoVerif}:
\begin{equation}
g_0= \frac{4\pi\hbar^2 a/m}{1-Ka/b}
\label{eq:g0_exp}
\end{equation}
with a numerical constant given by
\begin{equation}
K=\frac{12}{\pi}\int_{0}^{\pi/4}
d\theta\, \ln(1+1/\cos^{2}\theta)
= 2.442\ 749\ 607\ 806\ 335\ldots,
\end{equation}
and that may be expressed analytically in terms of the dilog special function.
\subsubsection{Simple variational upper bounds}
The relation Eq.(\ref{eq:g0_exp}) is quite instructive in the zero range limit $b\to 0$, for fixed non-zero
scattering length $a$ and atom numbers $N_\sigma$: In this limit, the lattice filling factor tends to zero, and the lattice model
is expected to converge to the continuous space zero-range model, that is to the Wigner-Bethe-Peierls
model described in subsection \ref{subsec:bp}. For each of the eigenenergies this means that
\begin{equation}
\label{eq:lat_to_bp}
\lim_{b\to 0} E_i(b) = E_i,
\end{equation}
where in the right hand side the set of $E_i$'s are the energy spectrum of the zero range model.
On the other hand, for a small enough value of $b$, the denominator in the
right-hand side of Eq.(\ref{eq:g0_exp}) is dominated by the term $-Ka/b$,
the lattice coupling constant $g_0$ is clearly negative,
and the lattice model is attractive, as already pointed out in \cite{kitp}.
By the usual variational argument,
this shows that the ground state energy of the zero range interacting gas is below
the one of the ideal gas, for the same trapping potential and atom numbers $N_\sigma$:
\begin{equation}
\label{eq:bounde}
E_0 \leq E_0^{\rm ideal}.
\end{equation}
Similarly, at thermal equilibrium in the canonical ensemble, the free energy of the interacting gas is below
the one of the ideal gas:
\begin{equation}
\label{eq:boundf}
F \leq F^{\rm ideal}.
\end{equation}
As in \cite{Blaizot} one indeed introduces the free-energy functional of the (here lattice model) interacting gas, $\mathcal{F}[\hat{\rho}]=
\mathrm{Tr}[H\hat{\rho}] +k_B T \mathrm{Tr}[\hat{\rho}\ln\hat{\rho}]$,
where $\hat{\rho}$ is any unit trace system density operator. Then
\begin{equation}
\label{eq:ident}
\mathcal{F}[\hat{\rho}_{\rm th}^{\rm ideal}] = F^{\rm ideal}(b) + \mathrm{Tr}[\hat{\rho}_{\rm th}^{\rm ideal}V],
\end{equation}
where $\hat{\rho}_{\rm th}^{\rm ideal}$ is the thermal equilibrium density operator
of the ideal gas in the lattice model, and $V$
is the interaction contribution to the $N$-body Hamiltonian.
Since the minimal value of $\mathcal{F}[\hat{\rho}]$ over $\hat{\rho}$
is equal to the interacting gas lattice model free energy $F(b)$,
the left hand side of Eq.(\ref{eq:ident}) is larger than $F(b)$.
Since the operator $V$ is negative for small $b$, because $g_0 <0$, the right hand side of Eq.(\ref{eq:ident}) is smaller than $F^{\rm ideal}(b)$.
Finally taking the limit $b\to 0$, one obtains the desired inequality.
The same reasoning can be performed in the grand canonical ensemble, showing
that the interacting gas grand potential is below the one of the ideal gas, for the same temperature and
chemical potentials $\mu_\sigma$:
\begin{equation}
\Omega \leq \Omega^{\rm ideal}.
\end{equation}
In \cite{ChevyNature}, for the unpolarized unitary gas, this last inequality was checked to be obeyed by the experimental
results, but it was shown, surprisingly, to be violated by some of the Quantum Monte Carlo results of
\cite{Burovski}.
For the particular case of the spatially homogeneous unitary gas, the above reasonings
imply that $\xi\leq 1$ in Eq.(\ref{eq:e0}), so that the unitary gas is attractive (in the ground branch,
see subsection \ref{subsec:aor}). Using the BCS variational ansatz in the lattice model
\footnote{One may check, e.g.\ in the sector $N_\uparrow=N_\downarrow=2$,
that the BCS variational wavefunction, which is a condensate of pairs in some pair wavefunction,
does not obey the Wigner-Bethe-Peierls boundary conditions
even if the pair wavefunction does, so it looses its variational character in the zero-range
model.} \cite{Varenna06}
one obtains the more stringent upper bound \cite{Randeria2}:
\begin{equation}
\xi \leq \xi_{\rm BCS} = 0.5906\ldots
\end{equation}
\subsubsection{Finite-range corrections}
For the parabolic dispersion relation, the expectation Eq.(\ref{eq:lat_to_bp}) was checked analytically for
two opposite spin particles: For $b\to 0$, in free space the scattering amplitude (\ref{eq:fklm}),
and in a box the lattice energy spectrum, converge to the predictions of the zero-range model \cite{LudoVerif}.
It was also checked numerically for $N=3$ particles in a box, with two $\uparrow$ particles
and one $\downarrow$ particle: As shown in Fig.~\ref{fig:3piab}, for the first low energy eigenstates
with zero total momentum, a convergence of the lattice eigenenergies to the Wigner-Bethe-Peierls ones is observed,
in a way that is eventually linear in $b$ for small enough values of $b$.
As discussed in \cite{Tangen},
this asymptotic linear dependence in $b$ is expected
for Galilean invariant continuous space models,
and the first order deviations of the eigenergies
from their zero range values are linear in the effective range $r_e$
of the interaction potential,
as defined in Eq.(\ref{eq:lke}), with model-independent coefficients:
\begin{equation}
\label{eq:deriv_univ}
\frac{dE_i}{d r_e}(b\to 0) \ \ \ \mbox{is model-independent}.
\end{equation}
However, for lattice models, Galilean invariance is broken and the scattering between two particles
depends on their center-of-mass momentum; this leads to a breakdown of
the universal relation~(\ref{eq:deriv_univ}),
while preserving the linear dependence of the energy with $b$
at low $b$~\cite{BurovskiNJP}.
\begin{figure}[b]
\sidecaption
\includegraphics[width=7cm,clip=]{3piab.eps}
\caption{
Diamonds: The first low eigenenergies for three $(\uparrow\uparrow\downarrow)$
fermions in a cubic box with a lattice model, as
functions of the lattice constant $b$ \cite{LudoVerif}.
The box size is $L$, with periodic boundary conditions, the scattering length is
infinite, the dispersion relation is parabolic Eq.(\ref{eq:pdr}).
The unit of energy is $E_0=(2\pi\hbar)^2/2mL^2$.
Straight lines: Linear fits performed on the data over the range ${b/L\leq 1/15}$,
except for the energy branch $E\simeq 2.89 E_0$ which is linear on a smaller range.
Stars in $b=0$: Eigenenergies predicted by the zero-range model.
}
\label{fig:3piab}
\end{figure}
A procedure to calculate $r_e$ in the lattice model
for a general dispersion relation
$\epsilon_{\mathbf{k}}$ in presented in Appendix~1. For the parabolic dispersion relation Eq.(\ref{eq:pdr}),
its value was given in \cite{Varenna06} in numerical form. With the technique exposed in Appendix~1,
we have now the analytical value:
\begin{equation}
\label{eq:rep}
r_e^{\rm parab} = b\, \frac{12\sqrt{2}}{\pi^3} \arcsin\frac{1}{\sqrt{3}} =
0.336\ 868 \ 47\ldots b.
\end{equation}
The usual Hubbard model, whose rich many-body physics is reviewed in \cite{AntoineVarenna},
was also considered in \cite{Varenna06}:
It is defined in terms of the tunneling
amplitude between neighboring lattice sites, here $t=-\hbar^2/(2mb^2)<0$, and of the on-site
interaction $U=g_0/b^3$. The dispersion relation is then
\begin{equation}
\epsilon_{\mathbf{k}} = \frac{\hbar^2}{mb^2} \sum_{\alpha=x,y,z} \left[1-\cos(k_\alpha b)\right]
\label{eq:disphub}
\end{equation}
where the summation is over the three dimensions of space. It reproduces the free space
dispersion relation only in a vicinity of $\mathbf{k}=\mathbf{0}$.
The explicit version of Eq.(\ref{eq:g0_gen}) is
obtained from Eq.(\ref{eq:g0_exp}) by replacing the numerical constant
$K$ by $K^{\rm Hub}=3.175 911 \ldots$. In the zero range limit this leads for $a\neq 0$ to
$U/|t| \to -7.913 552\ldots$, corresponding as expected to an {\sl attractive} Hubbard model,
lending itself to a Quantum Monte Carlo analysis for equal
spin populations with no sign problem \cite{Burovski,Bulgac}.
The effective range of the Hubbard model, calculated as in Appendix~1, remarkably is negative \cite{Varenna06}:
\begin{equation}
\label{eq:reh}
r_e^{\rm Hub} \simeq -0.305\, 718 b.
\end{equation}
It becomes thus apparent that an {\sl ad hoc} tuning of the dispersion relation $\epsilon_{\mathbf{k}}$
may lead to a lattice model with a zero effective range.
As an example, we consider a dispersion relation
\begin{equation}
\label{eq:mixte}
\epsilon_{\mathbf{k}} = \frac{\hbar^2 k^2}{2m} [ 1- C (kb/\pi)^2],
\end{equation}
where $C$ is a numerical constant less than $1/3$.
From Appendix~1 we then find that
\begin{equation}
\label{eq:magic}
r_e = 0 \ \ \ \ \mbox{for} \ \ \ \ C=0.2570224\ldots
\end{equation}
The corresponding value of $g_0$ is given by Eq.(\ref{eq:g0_exp}) with $K=2.899952\ldots$.
As pointed out in \cite{BurovskiNJP},
additionally fine-tuning the dispersion relation to cancel not only $r_e$ but also another coefficient (denoted by $B$ in \cite{BurovskiNJP})
may have some practical interest for Quantum Monte Carlo calculations that are performed
with a non-zero $b$, by canceling the undesired linear dependence of thermodynamical quantities and of the critical temperature
$T_c$ on $b$.
\subsubsection{Energy functional,
tail of the momentum distribution
and pair correlation function at short distances}
\label{subsubsec:tail}
A quite ubiquitous quantity in the short-range or large-momentum
physics of gases with zero range interactions
is the so-called ``contact'', which, restricting here for simplicity to thermal equilibrium in the canonical ensemble,
can be defined by
\begin{equation}
C\equiv\frac{4\pi m}{\hbar^2}\left(\frac{dE}{d(-1/a)}\right)_{\!S}
=\frac{4\pi m}{\hbar^2}\left(\frac{dF}{d(-1/a)}\right)_{\!T}.
\label{eq:defC}
\end{equation}
For zero-range interactions, this quantity $C$ determines the
large-$k$ tail of the momentum distribution
\begin{equation}
n_\sigma(\mathbf{k})\underset{k\to\infty}{\sim}\frac{C}{k^4}
\label{eq:C_nk}
\end{equation}
as well as the
short-distance behavior of the
pair distribution function
\begin{equation}
\int d^3R\ g^{(2)}_{\uparrow\downarrow}\left(\mathbf{R}+\frac{\mathbf{r}}{2},\mathbf{R}-\frac{r}{2}\right)\underset{r\to0}{\sim}\frac{C}{(4\pi r)^2}.
\label{eq:C_g2}
\end{equation}
Here the spin-$\sigma$ momentum distribution $n_\sigma(\mathbf{k})$ is
normalised as $\int \frac{d^3k}{(2\pi)^3}n_\sigma(\mathbf{k})=N_\sigma$.
The
relations~(\ref{eq:defC},\ref{eq:C_nk},\ref{eq:C_g2}) were obtained in
\cite{Tan1,Tan2}.
Historically, analogous relations were first established
for one-dimensional bosonic systems \cite{Lieb,Olshanii} with techniques
that may be straightforwardly extended to two dimensions and three
dimensions \cite{Tangen}.
Another relation derived in~\cite{Tan1} for the zero-range model
expresses the energy as a functional of the one-body density matrix:
\begin{equation}
E=
\sum_{\sigma=\uparrow,\downarrow}\int\frac{d^3k}{(2\pi)^3} \frac{\hbar^2 k^2}{2m}\left[n_\sigma(\mathbf{k})-\frac{C}{k^4}\right]
+\frac{\hbar^2 C}{4\pi m a}
+\sum_{\sigma=\uparrow,\downarrow}\int d^3r\,U(\mathbf{r}) \rho_\sigma(\mathbf{r})
\label{eq:fonctionelle}
\end{equation}
where $\rho_\sigma(\mathbf{r})$ is the spatial number density.
One usually uses (\ref{eq:C_nk}) to define $C$, and then derives (\ref{eq:defC}).
Here we rather take (\ref{eq:defC}) as the definition of $C$. This choice is convenient both for the two-channel model discussed in Section~\ref{subsec:tcm} and for the rederivation of
(\ref{eq:C_nk},\ref{eq:C_g2},\ref{eq:fonctionelle}) that we shall now present, where we use a lattice model before taking the zero-range limit.
From the Hellmann-Feynman theorem (that was already put forward in
\cite{Lieb}), the interaction energy $E_{\rm int}$ is equal to $g_0 (dE/dg_0)_{\!S}$. Since we have $d(1/g_0)/d(1/g)=1$ [see the relation (\ref{eq:g0_gen}) between $g_0$ and $g$], this can be rewritten as
\begin{equation}
E_{\rm int}=\frac{\hbar^4}{m^2} \frac{C}{g_0}.
\label{eq:E_int}
\end{equation}
Expressing $1/g_0$ in terms of $1/g$ using once again (\ref{eq:g0_gen}), adding the kinetic energy, and taking the zero-range limit, we immediately get the relation (\ref{eq:fonctionelle}). For the integral over momentum to be convergent, (\ref{eq:C_nk}) must hold (in the absence of mathematical pathologies).
To derive (\ref{eq:C_g2}), we again use (\ref{eq:E_int}), which implies that the relation
\begin{equation}
\sum_\mathbf{R} b^3\,g^{(2)}_{\uparrow\downarrow}\left(\mathbf{R}+\mathbf{r}/2,\mathbf{R}-\mathbf{r}/2 \right)=\frac{C}{(4\pi)^2}|\phi(\mathbf{r})|^2
\label{eq:g2_phi2}
\end{equation}
holds for $\mathbf{r}=\mathbf{0}$, were $\phi(\mathbf{r})$ is the zero-energy two-body scattering wavefunction, normalised in such a way that
\begin{equation}
\phi(\mathbf{r})\simeq \frac{1}{r}-\frac{1}{a} {\rm\ \ for\ \ } r\gg b
\label{eq:phi_norm}
\end{equation}
[see~\cite{Tangen} for the straightforward calculation of $\phi(\mathbf{0})$]. Moreover, in the regime where $r$ is much smaller than the typical interatomic distances and than the thermal de~Broglie wavelength
(but not necessarily smaller than $b$), it is generally expected that the $\mathbf{r}$-dependence of $g^{(2)}_{\uparrow\downarrow}(\mathbf{R}+\mathbf{r}/2,\mathbf{R}-\mathbf{r}/2)$ is proportional to $|\phi(\mathbf{r})|^2$, so that (\ref{eq:g2_phi2}) remains asymptotically valid. Taking the limits $b\to0$ and then $r\to0$ gives the desired (\ref{eq:C_g2}).
Alternatively, the link (\ref{eq:C_nk},\ref{eq:C_g2}) between short-range pair correlations and large-$k$ tail of the momentum distribution can be directly deduced from the
short-distance singularity of the wavefunction
coming from the contact condition~(\ref{eq:bpN})
and the corresponding tail in Fourier space~\cite{Tangen}, similarly to the original derivation in 1D~\cite{Olshanii}.
Thus this link remains true for a generic out-of-equilibrium statistical mixture of states satisfying the contact condition~\cite{Tan1,Tangen}.
\subsubsection{Absence of simple collapse}
To conclude this subsection on lattice models, we try to address the question of the advantage of lattice models
as compared to the standard continuous space model with a binary interaction potential $V(\mathbf{r})$ between
opposite spin fermions. Apart from practical advantages, due to the separable nature of the interaction
in analytical calculations, or to the absence of sign problem in the Quantum Monte Carlo
methods, is there a true physical advantage in using lattice models~?
One may argue for example that everywhere non-positive interaction potentials may be used
in continuous space, such as a square well potential, with a range dependent depth $V_0(b)$ adjusted
to have a fixed non-zero scattering and no two-body bound states. E.g.\ for a square well
potential $V(\mathbf{r})=-V_0 \theta(b-r)$, where $\theta(x)$ is the Heaviside function, one simply
has to take
\begin{equation}
V_0 = \frac{\hbar^2}{m b^2} \left(\frac{\pi}{2}\right)^2
\end{equation}
to have an infinite scattering length.
For such an attractive interaction, it seems then that one can easily reproduce the reasonings leading
to the bounds Eqs.(\ref{eq:bounde},\ref{eq:boundf}). It is known however that there exists
a number of particles $N$, in the unpolarized case $N_\uparrow=N_\downarrow$,
such that this model in free space has a $N$-body bound state, necessarily of energy $\propto -\hbar^2/(m b^2)$
\cite{Blatt,Panda,Baym}.
In the thermodynamic limit, the unitary gas is thus not the ground phase of the system, it is at most a metastable
phase, and this prevents a derivation of the bounds Eqs.(\ref{eq:bounde},\ref{eq:boundf}).
This catastrophe is easy to predict variationally, taking as a trial wavefunction the ground state
of the ideal Fermi gas enclosed in a fictitious cubic hard wall
cavity of size $b/\sqrt{3}$ \cite{theseFelix}.
In the large $N$ limit, the kinetic energy in the trial wavefunction is then $(3N/5) \hbar^2 k_F^2/(2m)$,
see Eq.(\ref{eq:e0}), where the Fermi wavevector is given by Eq.(\ref{eq:kf}) with
a density $\rho=N/(b/\sqrt{3})^3$, so that
\begin{equation}
E_{\rm kin} \propto N^{5/3}\frac{\hbar^2}{m b^2}.
\end{equation}
Since all particles are separated by a distance less than $b$,
the interaction energy is exactly
\begin{equation}
E_{\rm int} = -V_0 (N/2)^2
\end{equation}
and wins over the kinetic energy for $N$ large enough, $2800 \lesssim N$ for the considered ansatz.
Obviously, a similar reasoning leads to the same conclusion for an everywhere negative, non-necessarily square
well interaction potential \footnote{In fixed node calculations, an everywhere negative interaction
potential is used \cite{Carlson,Panda,Giorgini}. It is unknown if $N$ in these simulations
exceeds the minimal value required to have a bound state. Note that the imposed nodal wavefunction in
the fixed node method, usually the one of the Hartree-Fock or BCS state, would be however quite different from
the one of the bound state.}.
One could imagine to suppress this problem by introducing a hard core repulsion, in which case
however the purely attractive nature of $V$ would be lost, ruining our simple derivation
of Eqs.(\ref{eq:bounde},\ref{eq:boundf}).
The lattice models
are immune to this catastrophic variational
argument, since one cannot put more than two spin $1/2$ fermions ``inside" the interaction potential, that is
on the same lattice site. Still they preserve the purely attractive nature of the interaction.
This does not prove however that their spectrum is bounded from below in the zero range limit, as pointed
out in the introduction of this section.
\subsection{Zero-range model, scale invariance and virial theorem}
\label{subsec:bp}
\subsubsection{The zero-range model} \label{subsubsec:ZRM}
The interactions are here replaced with contact conditions on the $N$-body
wavefunction. In the two-body case, the model, introduced
already by Eq.(\ref{eq:bp}), is discussed in details in the literature,
see e.g.\ \cite{HouchesCastin99} in free space
where the scattering amplitude $f_k$
is calculated and the existence for $a>0$ of a dimer of energy
$-\hbar^2/(2\mu a^2)$ and wavefunction $\phi_0(r) =
(4\pi a)^{-1/2} \exp(-r/a)/r$ is discussed, $\mu$ being the reduced mass
of the two particles.
The two-body trapped case, solved in \cite{Wilkens}, was already
presented in subsection \ref{subsec:aor}. Here we present the model
for an arbitrary value of $N$.
For simplicity, we consider in first quantized form
the case of a fixed number $N_\uparrow$ of
fermions in spin state $\uparrow$ and a fixed number $N_\downarrow$ of
fermions in spin state $\downarrow$, assuming that the Hamiltonian
cannot change the spin state. We project the $N$-body
state vector $|\Psi\rangle$ onto the non-symmetrized spin state with the
$N_\uparrow$ first particles in spin state $\uparrow$
and the $N_\downarrow$ remaining particles in spin state $\downarrow$,
to define a scalar $N$-body wavefunction:
\begin{equation}
\label{eq:def_psi}
\psi(\mathbf{X}) \equiv \left(\frac{N!}{N_\uparrow! N_\downarrow!}\right)^{1/2}
\langle \uparrow, \mathbf{r}_1| \otimes
\ldots \langle \uparrow, \mathbf{r}_{N_\uparrow}|
\langle \downarrow, \mathbf{r}_{N_\uparrow+1}| \otimes \ldots
\langle \downarrow, \mathbf{r}_{N}| \Psi\rangle
\end{equation}
where $\mathbf{X}=(\mathbf{r}_1,\ldots,\mathbf{r}_N)$ is the set of all coordinates,
and the normalization factor ensures that $\psi$ is normalized to unity
\footnote{
The inverse formula giving the full state vector
in terms of $\psi(\mathbf{X})$ is
$|\Psi\rangle = \left(\frac{N!}{N_\uparrow!N_{\downarrow}!}\right)^{1/2}
A |\uparrow\rangle^{N_\uparrow}
|\downarrow\rangle^{N_\downarrow} |\psi\rangle$,
where the projector $A$ is the usual antisymmetrizing operator
$A=(1/N!) \sum_{\sigma\in S_N} \epsilon(\sigma) P_\sigma$.
}.
The fermionic symmetry of the state vector allows to express the wavefunction
on another spin state (with any different order of $\uparrow$
and $\downarrow$ factors) in terms of $\psi$. For the considered
spin state, this fermionic symmetry imposes that $\psi$ is odd
under any permutation of the first $N_{\uparrow}$ positions
$\mathbf{r}_1,\ldots,\mathbf{r}_{N_\uparrow}$,
and also odd under any permutation of the last $N_{\downarrow}$
positions $\mathbf{r}_{N_\uparrow+1},\ldots,\mathbf{r}_{N}$.
In the Wigner-Bethe-Peierls model, that we also call zero-range model,
the Hamiltonian for the wavefunction $\psi$ is simply represented
by the same partial differential operator as for the ideal gas case:
\begin{equation}
H =\sum_{i=1}^{N} \left[ -\frac{\hbar^2}{2m} \Delta_{\mathbf{r}_i}
+U(\mathbf{r}_i)\right],
\label{eq:hamil}
\end{equation}
where $U$ is the external trapping potential supposed for
simplicity to be spin state independent.
As is however well emphasized in the mathematics of operators
on Hilbert spaces \cite{analyse_spectrale}, an operator is defined
not only by a partial differential operator, but also by the choice
of its so-called {\sl domain} $D(H)$. A naive presentation
of this concept of domain is given in the Appendix~2.
Here the domain does not coincide with the ideal gas one.
It includes the following Wigner-Bethe-Peierls contact conditions:
For any pair of particles $i,j$, when $r_{ij}\equiv |\mathbf{r}_i-\mathbf{r}_j|\to 0$
for a fixed position of their centroid $\mathbf{R}_{ij}=(\mathbf{r}_i+\mathbf{r}_j)/2$,
there exists a function $A_{ij}$ such that
\begin{equation}
\label{eq:bpN}
\psi(\mathbf{X}) = A_{ij} (\mathbf{R}_{ij}; (\mathbf{r}_k)_{k\neq i,j})
(r_{ij}^{-1}-a^{-1}) + O(r_{ij}).
\end{equation}
These conditions are imposed for all values of $\mathbf{R}_{ij}$ different
from the positions of the other particles $\mathbf{r}_k$, $k$ different from
$i$ and $j$. If the fermionic particles $i$ and $j$ are in the same
spin state, the fermionic symmetry imposes $\psi(\ldots, \mathbf{r}_i=\mathbf{r}_j,\ldots)=0$
and one has simply $A_{ij}\equiv 0$.
For $i$ and $j$ in different spin states,
the unknown functions $A_{ij}$ have to be determined
from Schr\"odinger's equation, e.g.\ together with the
energy $E$ from the eigenvalue problem
\begin{equation}
\label{eq:ep}
H \psi = E \psi.
\end{equation}
Note that in Eq.(\ref{eq:ep}) we have excluded the values of $\mathbf{X}$
where two particle positions coincide. Since
$\Delta_{\mathbf{r}_i} r_{ij}^{-1} = - 4 \pi \delta(\mathbf{r}_i-\mathbf{r}_j)$,
including these values
would require a calculation with distributions rather than with functions,
with regularized delta interaction pseudo-potential, which is
a compact and sometimes useful reformulation of the Wigner-Bethe-Peierls
contact conditions \cite{Petrov,HouchesCastin99,Huang,CRASCastin}.
As already pointed out below Eq.(\ref{eq:bpN}), $A_{ij}\equiv 0$
if $i$ and $j$ are fermions in the same spin state. One may wonder
if solutions exist such that $A_{ij}\equiv 0$ even if $i$ and $j$
are in different spin states, in which case $\psi$ would
simply vanish when $r_{ij}\to 0$.
These solutions would then be common eigenstates
to the interacting gas and to the ideal gas. They would correspond
in a real experiment to long lived eigenstates, protected from three-body
losses by the fact that $\psi$ vanishes when two particles or more
approach each other.
In a harmonic trap, one can easily construct such ``non-interacting"
solutions, as for example
the famous Laughlin wavefunction of the Fractional Quantum Hall Effect.
``Non-interacting" solutions also exists for spinless bosons.
These non-interacting states actually dominate the ideal gas density of states at high energy
\cite{WernerPRL,theseFelix}.
\subsubsection{What is the kinetic energy?}
The fact that the Hamiltonian is the same as the ideal gas, apart from
the domain, may lead physically to some puzzles. E.g.\ the absence of
interaction term may give the impression that the energy $E$ is the
sum of trapping potential energy and kinetic energy only.
This is actually not so. The correct definition of the mean kinetic energy,
valid for general boundary conditions on the wavefunction, is
\begin{equation}
E_{\rm kin} = \int d^{3N}X\, \frac{\hbar^2}{2m}
|\partial_\mathbf{X} \psi|^2.
\end{equation}
This expression in particular guaranties that $E_{\rm kin}\geq 0$.
If $A_{ij}\neq 0$ in Eq.(\ref{eq:bpN}), one then sees that, although
$\psi$ is square integrable in a vicinity of $r_{ij}=0$
thanks to the Jacobian $\propto r_{ij}^2$ coming from three-dimensional
integration, the gradient of $\psi$ diverges as $1/r_{ij}^2$
and cannot be square integrable. Within the zero-range model
one then obtains an infinite kinetic energy
\begin{equation}
E_{\rm kin}^{\rm WBP}=+\infty.
\end{equation}
Multiplying Eq.(\ref{eq:ep}) by $\psi$ and integrating over $\mathbf{X}$,
one realizes that the total energy is split as the trapping
potential energy,
\begin{equation}
E_{\rm trap} = \int d^{3N}X\, |\psi(\mathbf{X})|^2 \sum_{i=1}^{N} U(\mathbf{r}_i)
\end{equation}
and as the sum of kinetic plus interaction energy:
\begin{equation}
E_{\rm kin} + E_{\rm int} = -\int d^{3N}X\, \frac{\hbar^2}{2m}
\psi^* \Delta_\mathbf{X} \psi.
\end{equation}
This means that the interaction energy is $-\infty$ in the Wigner-Bethe-Peierls
model. All this means is that, in reality,
when the interaction has a non-zero range,
both the kinetic energy and the interaction energy
of interacting particles depend on the interaction range $b$,
and diverge for $b\to 0$, in such a way however that the sum
$E_{\rm kin} + E_{\rm int}$ has a finite limit given by the Wigner-Bethe-Peierls
model.
We have seen more precisely how this happens for lattice models in section \ref{subsubsec:tail},
see~the expression~(\ref{eq:E_int}) of $E_{\rm int}$ and the subsequent derivation of~(\ref{eq:fonctionelle}).\footnote{For a continuous-space model with an interaction potential $V(r)$,
we have~\cite{ZhangLeggett,Tangen} $E_{\rm int}=\frac{C}{(4\pi)^2}\int d^3r\,V(r) |\phi(r)|^2$ where $C$ is still defined by~(\ref{eq:defC}) and $\phi(r)$ still denotes the zero-energy two-body scattering state normalised according to~(\ref{eq:phi_norm}).}
\subsubsection{Scale invariance and virial theorem}
In the case of the unitary gas, the scattering length is infinite,
so that one sets $1/a=0$ in Eq.~(\ref{eq:bpN}). The domain of the Hamiltonian
is then imposed to be invariant by any isotropic rescaling Eq.(\ref{eq:sca_psi})
of the particle positions. To be precise, we define
for any scaling factor $\lambda >0$:
\begin{equation}
\label{eq:sctr}
\psi_\lambda(\mathbf{X}) \equiv \frac{\psi(\mathbf{X}/\lambda)}{\lambda^{3N/2}},
\end{equation}
and we impose that $\psi_\lambda\in D(H)$ for all $\psi\in D(H)$.
This is the precise mathematical definition of the scale invariance
loosely introduced in subsection \ref{subsec:ssfp}.
In particular, it is apparent in Eq.(\ref{eq:bpN})
that, for $1/a=0$, $\psi_\lambda$ obeys the Wigner-Bethe-Peierls contact
conditions if $\psi$ does. On the contrary, if $\psi$ obeys
the contact conditions for a finite scattering length $a$, $\psi_\lambda$ obeys
the contact condition for a different, fictitious
scattering length $a_\lambda =\lambda a \neq a$
and $D(H)$ cannot be scaling invariant.
There are several consequences of the scale invariance of the domain of the Hamiltonian $D(H)$
for the unitary gas. Some of them were presented in subsection
\ref{subsec:ssfp}, other ones will be derived in section \ref{sec:ds}.
Here we present another application, the derivation of a virial
theorem for the unitary gas. This is a first step towards
the introduction of a SO(2,1) Lie algebra in section \ref{sec:ds}.
To this end, we introduce the infinitesimal generator $D$
of the scaling transform Eq.(\ref{eq:sctr}), such that
\footnote{
The scaling transform (\ref{eq:sctr}) defines
a unitary operator $T(\lambda)$ such that $\psi_\lambda = T(\lambda) \psi$.
One has $T(\lambda_1) T(\lambda_2) = T(\lambda_1 \lambda_2)$.
To recover the usual additive structure as for the group of spatial
translations, one sets $\lambda=\exp\theta$, so that $T(\theta_1)
T(\theta_2)=T(\theta_1+\theta_2)$ and $T(\theta)=\exp(-i\theta D)$
where $D$ is the generator. This is why $\ln\lambda$ appears in
Eq.(\ref{eq:itod}).
}
\begin{equation}
\label{eq:itod}
\psi_\lambda(\mathbf{X}) = e^{-i D \ln \lambda} \psi(\mathbf{X}).
\end{equation}
Taking the derivative of Eq.(\ref{eq:sctr}) with respect to
$\lambda$ in $\lambda=1$, one obtains the hermitian operator
\begin{equation}
\label{eq:defD}
D = \frac{1}{2i} (\mathbf{X}\cdot \partial_\mathbf{X} + \partial_\mathbf{X} \cdot \mathbf{X})=
\frac{3N}{2i} -i \mathbf{X}\cdot \partial_\mathbf{X}.
\end{equation}
The commutator of $D$ with the Hamiltonian is readily obtained.
From the relation
$\Delta_{\mathbf{X}} \psi_\lambda(\mathbf{X}) = \lambda^{-2}
(\Delta \psi)(\mathbf{X}/\lambda)$, one has
\begin{equation}
e^{iD\ln \lambda} (H-H_{\rm trap}) e^{-iD\ln \lambda} =
\frac{1}{\lambda^2} (H-H_{\rm trap})
\end{equation}
where $H_{\rm trap}=\sum_{i=1}^{N} U(\mathbf{r}_i)$ is the trapping potential
part of the Hamiltonian. It remains to take the derivative in $\lambda=1$ to
obtain
\begin{equation}
\label{eq:comm1}
i [D, H-H_{\rm trap}] = - 2 (H-H_{\rm trap}).
\end{equation}
The commutator of $D$ with the trapping potential is evaluated directly from
Eq.(\ref{eq:defD}):
\begin{equation}
\label{eq:comm2}
i[D, H_{\rm trap}] = \sum_{i=1}^{N} \mathbf{r}_i\cdot \partial_{\mathbf{r}_i}U(\mathbf{r}_i).
\end{equation}
This gives finally
\begin{equation}
\label{eq:comm}
i [D,H] = - 2 (H-H_{\rm trap}) + \sum_{i=1}^{N} \mathbf{r}_i\cdot \partial_{\mathbf{r}_i}U(\mathbf{r}_i).
\end{equation}
The standard way to derive the virial theorem in quantum mechanics
\cite{virial}, in a direct generalization of the one of classical
mechanics, is then to take the expectation value of $[D,H]$ in an eigenstate
$\psi$ of $H$ of eigenenergy $E$.
This works here for the unitary gas because the domain $D(H)$
is preserved by the action of $D$.
On one side, by having $H$ acting on $\psi$ from the right
or from the left, one trivially has $\langle [D,H]\rangle_\psi=0$.
On the other side, one has Eq.(\ref{eq:comm}), so that
\begin{equation}
\label{eq:virth}
E = \sum_{i=1}^{N} \langle U(\mathbf{r}_i) +
\frac{1}{2} \mathbf{r}_i\cdot \partial_{\mathbf{r}_i}U(\mathbf{r}_i) \rangle_\psi.
\end{equation}
This relation was obtained with alternative derivations in the literature
(see \cite{Thomas_viriel_demo} and references therein).
One of its practical interests is that it gives access to the energy
from the gas density distribution \cite{ThomasVirial}.
As already mentioned, the scale invariance of the domain of $H$
is crucial to obtain this result. If $1/a$ is non zero,
a generalization of the virial relation can however
be obtained, that involves $dE/d(1/a)$, see \cite{TanVirial,WernerVirial}.
\subsection{Two-channel model and closed-channel fraction}
\label{subsec:tcm}
\subsubsection{The two-channel model}
\label{subsubsec:tcmdes}
The lattice models or the zero-range model are of course dramatic simplifications
of the real interaction between two alkali atoms.
At large interatomic distances, much larger than the radius of the electronic orbitals,
one may hope to realistically represent this interaction by a function $V(r)$ of the interatomic
distance, with a van der Waals attractive tail $V(r) \simeq -C_6/r^6$,
a simple formula that actually neglects retardation effects and long-range magnetic dipole-dipole
interactions.
As discussed below the gas phase condition Eq.(\ref{eq:gaz}), this allows to estimate
$b$ with the so-called van der Waals length, usually in the range of 1-10 nm.
At short interatomic distances, this simple picture of a scalar interaction potential $V(r)$
has to be abandoned. Following quantum chemistry or molecular physics methods, one has to
introduce the various Born-Oppenheimer potential curves obtained from the solution of the electronic
eigenvalue problem for fixed atomic nuclei positions. Restricting to one active electron of spin $1/2$
per atom, one immediately gets two ground potential curves, the singlet one
corresponding to the total spin $S=0$,
and the triplet one corresponding to the total spin $S=1$.
An external magnetic field $B$ is applied to activate the Feshbach resonance.
This magnetic field couples mainly to the total electronic spin and thus induces
{\sl different} Zeeman shifts for the singlet and triplet curves.
In reality, the problem is further complicated by the existence of the nuclear spin and the hyperfine
coupling, that couples the singlet channel to the triplet channel for nearby atoms,
and that induces a hyperfine splitting within the ground electronic state for well separated atoms.
A detailed discussion is given e.g. in \cite{revue_feshbach,Timm}. Here we take
the simplified view depicted in Fig.\ref{fig:feshbach}: The atoms
interact {\sl via} two potential curves, $V_{\rm open}(r)$ and
$V_{\rm closed}(r)$.
At large distances, $V_{\rm open}(r)$ conventionally tends to zero,
whereas $V_{\rm closed}(r)$ tends to a positive value $V_\infty$, one of
the hyperfine energy level spacings for a single atom in the applied magnetic field.
In the two-body scattering problem, the atoms come from $r=+\infty$
in the internal state corresponding to $V_{\rm open}(r)$, the so-called open channel,
with a kinetic energy $E\ll V_\infty$.
Due to a coupling between the two channels,
the two interacting atoms can have access to the internal state corresponding
to the curve $V_{\rm closed}(r)$, but only at short distances; at long distances,
the external atomic wavefunction in this so-called closed channel
is an evanescent wave that decays exponentially with $r$ since $E<V_{\infty}$.
Now assume that, in the absence of coupling between the channels, the closed channel
supports a bound state of energy $E_b$, called in what follows {\sl the molecular state}, or {\sl the closed-channel molecule}.
Assume also that, by applying a judicious magnetic field, one tunes the energy of
this molecular state close to zero, that is to the dissociation limit of the
open channel. In this case one may expect that the scattering amplitude of
two atoms is strongly affected, by a resonance effect, given the non-zero coupling
between the two channels. This is in essence how the Feshbach resonance takes place.
\begin{figure}[b]
\sidecaption
\includegraphics[width=7cm,clip=]{feshbach.eps}
\caption{Simple view of a Feshbach resonance.
The atomic interaction is described by two curves (solid line:
open channel, dashed line: closed channel).
When one neglects the interchannel coupling $\Lambda$, the closed channel
has a molecular state of energy $E_b$ close to the dissociation
limit of the open channel. The energy spacing $V_\infty$
was greatly exaggerated, for clarity.}
\label{fig:feshbach}
\end{figure}
The central postulate of the theory of quantum gases is that the short range details
of the interaction are unimportant, only the low-momentum scattering amplitude $f_k$ between
two atoms is relevant. As a consequence, any simplified model for the interaction, leading
to a different scattering amplitude $f_k^{\rm model}$, is acceptable provided that
\begin{equation}
\label{eq:postu}
f_k^{\rm model} \simeq f_k
\end{equation}
for the relevant values of the relative momentum $k$ populated in the gas.
We insist here that we impose similar scattering amplitudes over some momentum range, not just
equal scattering lengths $a$. For spin 1/2 fermions, typical values of $k$ can be
\begin{equation}
k_{\rm typ}\in \{a^{-1}, k_F, \lambda_{\rm dB}^{-1}\}
\end{equation}
where the Fermi momentum is defined in Eq.(\ref{eq:kf}) and the thermal de Broglie
wavelength in Eq.(\ref{eq:ldb}).
The appropriate value of $k_{\rm typ}$ depends on the physical
situation. The first choice $k_{\rm typ}\sim a^{-1}$ is well suited to the case
of a condensate of dimers ($a>0$) since it is the relative momentum of two atoms forming
the dimer. The second choice $k_{\rm typ}\sim k_F$ is well suited to a degenerate
Fermi gas of atoms (not dimers). The third choice $k_{\rm typ} \sim \lambda^{-1}$
is relevant for a non-degenerate Fermi gas.
The strategy is thus to perform an accurate calculation of the ``true" $f_k$,
to identify the validity conditions of the simple models and of the unitary regime
assumption Eq.(\ref{eq:fku}).
One needs a realistic, though analytically tractable, model of the Feshbach resonance.
This is provided by the so-called {\sl two-channel} models \cite{Timm,Holland0,Holland}.
We use here the version presented in \cite{Tarruell}, which is a particular case of the one used
in~\cite{revue_feshbach,KoehlerBurnett} and Refs. therein:
The open channel part consists of the original gas of spin $1/2$ fermions that interact
{\sl via} a separable potential, that is in first quantized form for two opposite spin fermions,
in position space:
\begin{equation}
\langle \mathbf{r}_1, \mathbf{r}_2 | V_{\rm sep} | \mathbf{r}_1',\mathbf{r}_2'\rangle =
\delta\left(\frac{\mathbf{r}_1+\mathbf{r}_2}{2}-\frac{\mathbf{r}_1'+\mathbf{r}_2'}{2}\right)
g_0 \chi(\mathbf{r}_2-\mathbf{r}_1)\chi(\mathbf{r}_2'-\mathbf{r}_1').
\end{equation}
This potential does not affect the atomic center of mass, so it conserves total momentum
and respects Galilean invariance. Its matrix element involves the product of a function of the relative position
in the ket and of the same function of the relative position in the bra, hence the name {\sl separable}.
The separable potential is thus in general non local. As we shall take a function $\chi$ of width $\approx b$
this is clearly not an issue. The coupling constant $g_0$ of the separable potential
is well-defined by the normalization condition for $\chi$, $\int d^3r \, \chi(\mathbf{r})=1$.
In the presence of this open channel interaction only, the scattering length between fermions,
the so-called background scattering length $a_{\rm bg}$, is usually small, of the order of the potential
range $b$, hence the necessity of the Feshbach resonance to reach the unitary limit.
In the closed channel part, a single two-particle state is kept, the one corresponding to the
molecular state, of energy $E_b$ and of spatial range $\lesssim b$. The atoms thus exist in that channel
not in the form of spin $1/2$ fermions, but in the form of bosonic spinless molecules, of mass twice
the atomic mass.
The coupling between the two channels simply corresponds to the possibility for each boson
to decay in a pair of opposite spin fermions, or the inverse process that two opposite spin fermions
merge into a boson, in a way conserving the total momentum. This coherent Bose-Fermi conversion
may take place only if the positions $\mathbf{r}_1$ and $\mathbf{r}_2$ of the two fermions are within a distance $b$, and is thus described
by a relative position dependent amplitude $\Lambda \chi(\mathbf{r}_1-\mathbf{r}_2)$, where for simplicity one takes
the same cut-off function $\chi$ as in the separable potential.
It is important to realize that the Bose-Fermi conversion effectively induces an interaction
between the fermions, which becomes resonant for the right tuning of $E_b$ and leads to
the diverging total scattering length $a$.
The model is best summarized in second quantized form \cite{Tarruell}, introducing the fermionic field operators
$\psi_\sigma(\mathbf{r})$, $\sigma=\uparrow,\downarrow$,
obeying the usual fermionic anticommutation relations, and the bosonic field operator
$\psi_b(\mathbf{r})$ obeying the usual bosonic commutation relations:
\begin{eqnarray}
\label{eq:definition_H2}
&& H = \int d^3r \left[\sum_{\sigma=\uparrow,\downarrow}
\psi_\sigma^\dagger\left(-\frac{\hbar^2}{2m}\Delta_\mathbf{r} + U\right) \psi_\sigma +
\psi_b^\dagger\left(E_b-\frac{\hbar^2}{4m}\Delta_\mathbf{r} + U_b\right) \psi_b
\right] \\
&& +
\Lambda \int d^3r_1 d^3r_2\, \chi(\mathbf{r}_1-\mathbf{r}_2)
\left\{
\psi_b^\dagger[(\mathbf{r}_1+\mathbf{r}_2)/2]
\psi_\downarrow(\mathbf{r}_1) \psi_\uparrow(\mathbf{r}_2)
+ {\rm h. c.}
\right\}
\nonumber \\
&& + g_0 \int d^3\!R\, d^3\!r\, d^3\!r'
\chi(\mathbf{r}) \chi (\mathbf{r}')
\psi_\uparrow^\dagger(\mathbf{R}-\mathbf{r}/2) \psi_\downarrow^\dagger(\mathbf{R}+\mathbf{r}/2)
\psi_\downarrow(\mathbf{R}+\mathbf{r}'/2) \psi_\uparrow(\mathbf{R}-\mathbf{r}'/2),
\nonumber
\end{eqnarray}
where $U(\mathbf{r})$ and $U_b(\mathbf{r})$ are the trapping potentials for the fermions and the bosons, respectively.
\subsubsection{Scattering amplitude and universal regime}
\label{subsubsec:fk_2canaux}
In free space, the scattering problem of two fermions
is exactly solvable
for a Gaussian cut-off function $\chi(\mathbf{r})\propto \exp[-r^2/(2 b^2)]$~ \cite{revue_feshbach,Tarruell}. A variety of parameterizations are possible.
To make contact with typical notations, we assume that the energy $E_b$ of the molecule in the closed
channel is an affine function of the magnetic field $B$, a reasonable assumption close to the Feshbach resonance:
\begin{equation}
\label{eq:reasas}
E_b(B) = E_b^0 + \mu_b (B-B_0)
\end{equation}
where $B_0$ is the magnetic field value right on resonance and $\mu_b$ is the effective magnetic moment
of the molecule. Then the scattering length for the model Eq.(\ref{eq:definition_H2}) can be exactly
written as the celebrated formula
\begin{equation}
\label{eq:celeb}
a = a_{\rm bg} \left(1-\frac{\Delta B}{B-B_0}\right),
\end{equation}
where $\Delta B$, such that $E_b^0 + \mu_b \Delta B = \Lambda^2/g_0$,
is the so-called width of the Feshbach resonance. As expected, for $|B-B_0|\gg |\Delta B|$, one finds
that $a$ tends to the background scattering length $a_{\rm bg}$ solely due to the open channel interaction.
With $\Delta B$ one forms a length $R_*$ \cite{PetrovBosons} which is always non-negative:
\begin{equation}
\label{eq:Ret}
R_* \equiv \frac{\hbar^2}{m a_{\rm bg} \mu_b \Delta B} =
\left(\frac{\Lambda}{2\pi b E_b^0}\right)^2,
\end{equation}
where the factor $2\pi$ is specific to our choice of $\chi$.
Physically, the length $R_*$ is also directly related to the effective range on resonance:
\begin{equation}
r_e^{\rm res} = - 2 R_* + \frac{4b}{\sqrt{\pi}},
\end{equation}
where the numerical coefficient in the last term depends on the choice of $\chi$.
The final result for the scattering amplitude for the model Eq.(\ref{eq:definition_H2}) is
\begin{equation}
\label{eq:lerpf}
-\frac{1}{f_{k}} = ik + \frac{e^{k^2b^2}}{a}
\left[1-\left(1-\frac{a}{a_{\rm bg}}\right)\frac{k^2}{k^2-Q^2}\right]
-i k\, \mathrm{erf}\,(-i k b)
\end{equation}
where erf is the error function, that vanishes linearly in zero, and the wavevector $Q$, such that
\begin{equation}
Q^2 \equiv \frac{m}{\hbar^2 g_0} \left(g_0 E_b-\Lambda^2\right)
= \frac{-1}{a_{\rm bg} R_*(1-a_{\rm bg}/a)},
\end{equation}
may be real or purely imaginary.
The unitary limit assumption Eq.(\ref{eq:fku}) implies that all the terms in the right hand side
of Eq.(\ref{eq:lerpf}) are negligible, except for the first one.
We now discuss this assumption, restricting for simplicity to an infinite scattering length $a^{-1}=0$ (i.e. a magnetic field sufficiently close to resonance)
and a typical relative momentum $k_{\rm typ}=k_F$ (i.e. a degenerate gas).
To satisfy Eq.(\ref{eq:postu}), with $f_k^{\rm model}=-1/(ik)$, one should then have,
in addition to the gas phase requirement $k_F b\ll 1$, that
\begin{equation}
\label{eq:cond_valid}
\frac{k R_*}{|1+k^2 a_{\rm bg} R_*|} \ll 1 \ \ \forall k\in [0,k_F].
\end{equation}
Table~\ref{tab:valid} summarizes the corresponding conditions to reach the unitary limit.\footnote{We discarded for simplicity the rather peculiar case where $k_F \sqrt{|a_{\rm bg}| R_*}$ is $\le1$ but not $\ll1$.} \footnote{
An additional condition actually has to be imposed
to
have a universal gas, as
we will see after Eq.(\ref{eq:Nb_unitaire}).}
Remarkably, the condition $k_F |r_e^{\rm res}|\ll 1$ obtained in Eq.(\ref{eq:zrl})
from the expansion of $1/f_k$ to order $k^2$ is not
the end of the story.
In particular,
if $a_{\rm bg} < 0$, $Q_{\rm res}^2\equiv-1/(a_{\rm bg} R_*)$ is positive and $1/f_k$ diverges
for $k=Q_{\rm res}$; if the location of this divergence is within the Fermi sea, the unitary
limit is not reachable. This funny case however requires huge values of $R_* a_{\rm bg}$,
that is extremely small values of the resonance width $\Delta B$:
\begin{equation}
\label{eq:narrow}
| \mu_b \Delta B | \lesssim \frac{\hbar^2 k_F ^2}{2m}.
\end{equation}
This corresponds to
very narrow Feshbach resonances \cite{Gurarie},
whose experimental use requires a good control of the magnetic
field homogeneity and is more delicate.
Current experiments rather use broad Feshbach resonances such as on lithium 6,
where $r_e^{\rm res}=4.7$nm \cite{Strinati},
$a_{\rm bg}=-74$ nm, $R_*=0.027$nm \cite{Grimm}, leading to $1/(|a_{\rm bg}| R_*)^{1/2}
=700 (\mu\mathrm{m})^{-1}$ much larger than $k_F \approx$ a few $(\mu\mathrm{m})^{-1}$, so that the unitary
limit is indeed well reached.
\begin{table}
\caption{In the two-channel model, conditions deduced from Eq.(\ref{eq:cond_valid}) (supplementary to the gas phase
condition $k_F b \ll 1$) to reach the unitary limit
for a degenerate gas of spin 1/2 fermions of Fermi momentum $k_F$. It is assumed
that the magnetic field is tuned right on resonance, so that the scattering length is infinite.
The last column corresponds to narrow Feshbach resonances satisfying Eq.(\ref{eq:narrow}).}
\label{tab:valid}
\begin{tabular}{p{3.1cm}p{4.1cm}p{4.1cm}}
\hline\noalign{\smallskip}
&
$k_F \sqrt{|a_{\rm bg}| R_*} \ll 1$ &
$k_F \sqrt{|a_{\rm bg}| R_*}> 1$ \\
\noalign{\smallskip}\svhline\noalign{\smallskip}
$a_{\rm bg} > 0$ & $k_F R_* \ll 1$ &
$(R_*/a_{\rm bg})^{1/2} \ll 1$
\\
$a_{\rm bg} < 0$ & $k_F R_* \ll 1$ & unreachable \\
\noalign{\smallskip}\hline\noalign{\smallskip}
\end{tabular}
\end{table}
\subsubsection{Relation between number of closed channel molecules and ``contact''}
\label{subsubsec:nccm}
The fact that the two-channel model includes the underlying atomic physics of the Feshbach resonance
allows to consider an observable that is simply absent from
single channel models,
namely
the number of molecules in the closed channel, represented by the operator:
\begin{equation}
N_b \equiv \int d^3\mathbf{r}\ \psi_b^\dagger(\mathbf{r}) \psi_b(\mathbf{r})
\end{equation}
where $\psi_b$ is the molecular field operator. The mean number $\langle N_b\rangle$ of closed
channel molecules was recently measured by laser molecular excitation techniques \cite{Hulet_mol}.
This mean number can be calculated from a two-channel model by a direct application of
the Hellmann-Feynman theorem~\cite{BraatenLong,Tarruell}
(see also~\cite{ZhangLeggett}). The key point is that the only quantity depending on the magnetic
field in the Hamiltonian Eq.(\ref{eq:definition_H2}) is the internal energy $E_b(B)$ of a closed
channel molecule.
At thermal equilibrium in the canonical ensemble,
we thus have
\begin{equation}
\label{eq:hell}
\left(\frac{dE}{dB}\right)_{\!S}= \langle N_b\rangle \frac{dE_b}{dB}.
\end{equation}
Close to the Feshbach resonance, we may assume that $E_b$ is an affine function of $B$, see Eq.(\ref{eq:reasas}),
so that the scattering length $a$ depends on the magnetic field as in Eq.(\ref{eq:celeb}).
Parameterizing $E$ in terms of the inverse scattering length rather than $B$, we can replace
$dE/dB$ by $dE/d(1/a)$ times $d(1/a)/dB$.
The latter can be calculated explicitly from (\ref{eq:celeb}). Thus
\begin{equation}
\langle N_b\rangle = \frac{C}{4\pi} R_* \left(1-\frac{a_{\rm bg}}{a}\right)^2,
\label{eq:CvsNb}
\end{equation}
where
$C$
is the contact defined in~Eq.(\ref{eq:defC}),
and
we introduced the length $R_*$ defined in Eq.(\ref{eq:Ret}).
If the interacting gas is in the universal zero range regime, its energy $E$ depends on the interactions only via the scattering length,
independently of the microscopic details of the atomic interactions, and its dependence with $1/a$ may
be calculated by any convenient model. Then, at zero temperature, for the unpolarized case $N_\uparrow
=N_\downarrow$,
the equation of state of the homogeneous gas can be expressed as
\begin{equation}
e_0 = e_0^{\rm ideal} f\left(\frac{1}{k_F a}\right),
\label{eq:EOS_T=0}
\end{equation}
where $e_0$ and $e_0^{\rm ideal}$ are the ground state energy per particle for the interacting gas
and for the ideal gas with the same density,
and the Fermi wavevector $k_F$ was defined in Eq.(\ref{eq:kf}).
In particular, $f(0)=\xi$, where the number
$\xi$ was introduced in Eq.(\ref{eq:e0}).
Setting $\zeta\equiv -f'(0)$,
we have for the homogeneous unitary gas
\begin{equation}
\frac{C^{\rm hom}}{V} = \zeta\frac{2}{5\pi}k_F^4,
\label{eq:C_hom_vs_zeta}
\end{equation}
so that
\begin{equation}
\frac{\langle N_b\rangle^{\rm hom}}{N} = \frac{3}{10} k_F R_* \zeta.
\label{eq:Nb_unitaire}
\end{equation}
This expression is valid for a universal gas consisting mainly of fermionic atoms, which requires that $\langle N_b\rangle^{\rm hom}/N\ll1$, i. e. $k_F R* \ll 1$. This condition was already obtained in \S\ref{subsubsec:fk_2canaux}
for the broad resonances of the left column of Table~\ref{tab:valid}.
In the more exotic case of the narrow resonances of the second column of Table~\ref{tab:valid}, this condition has to be imposed in addition to the ones of Table~\ref{tab:valid}.
\subsubsection{Application of general relations: Various measurements of the contact}
The relation (\ref{eq:CvsNb}) allowed us to extract in~\cite{Tarruell} the contact $C$
of the trapped gas
[related to the derivative of the total energy of the trapped gas {\it via}~(\ref{eq:defC})]
from the values of $N_b$ measured in~\cite{Hulet_mol}.
The result is shown in Fig.\ref{fig:hulet}, together with a theoretical zero-temperature curve resulting from
the local density approximation
in the harmonically trapped case
where $U(\mathbf{r})=\frac{1}{2} m\sum_\alpha \omega_\alpha^2 x_\alpha^2$,
the function $f$ of~(\ref{eq:EOS_T=0}) being obtained by interpolating between
the fixed-node Monte-Carlo data of~\cite{Giorgini, LoboGiorgini}
and the known asymptotic expressions in the BCS and BEC limits\footnote{See~\cite{Tarruell} for details. The cusp at unitarity is of course an artefact of this
interpolation procedure.}.
While
this is the
first direct measurements of the contact in the BEC-BCS crossover,
it has also been measured more recently:
\begin{itemize}
\item
using Bragg scattering, {\it via} the
large-momentum tail of the structure factor, directly related
by Fourier transformation to the
short-distance singularity Eq.(\ref{eq:C_g2}) of the pair correlation function~\cite{Australiens}, see the cross at unitarity in Fig.\ref{fig:hulet}
\item
{\it via} the tail of the momentum distribution Eq.(\ref{eq:C_nk}) measured by abruptly turning off both trapping potential and interactions~\cite{Jin_tail}, see the squares in Fig.\ref{fig:hulet}
\item
{\it via}
(momentum resolved)
radio-frequency spectroscopy~\cite{Jin_tail,BraatenChap}.
\end{itemize}
\begin{figure}[b]
\centerline{\includegraphics[width=9cm,clip=]{Ctrap_for_chap.eps}}
\caption{The contact $C=\frac{dE}{d(-1/a)}\frac{4\pi m}{\hbar^2}$ of a trapped unpolarized Fermi gas.
The circles are obtained from the measurements of
$\langle N_b\rangle$ in \cite{Hulet_mol}, combined with the two channel model theory linking
$\langle N_b\rangle$ to $C$ [Eq.(\ref{eq:CvsNb})].
The cross was obtained in~\cite{Australiens} by measuring the structure factor.
The squares were obtained in~\cite{Jin_tail} by measuring the momentum distribution.
Solid line: zero-temperature theoretical prediction extracted from \cite{Giorgini}
as detailed in \cite{Tarruell}.
Here the Fermi wavevector $k_F^{\rm trap}$ of the trapped gas is defined by
$\hbar^2 (k_F^{\rm trap})^2/(2m)=(3N)^{1/3}\hbar \bar{\omega}$,
with $\bar{\omega}$ the geometric mean of the three oscillation frequencies
$\omega_\alpha$ and $N$ the total atom number.
}
\label{fig:hulet}
\end{figure}
For the homogeneous unitary gas,
the contact is conveniently expressed in terms of the dimensionless parameter $\zeta$ [see (\ref{eq:C_hom_vs_zeta})].
The experimental value
$\zeta=0.91(5)$ was obtained
by measuring the equation of state of the homogeneous gas
(with the technique proposed in \cite{Ho_pascal})
and taking the derivative of the energy with respect to the inverse scattering length [Eq.(\ref{eq:defC})]~(see
\cite{NavonEOS} and the contribution
of F. Chevy and C. Salomon to this volum
).
From the fixed-node Monte-Carlo calculations, one gets
$\zeta\simeq 1$ by taking a derivative of the data of~\cite{Giorgini} for the function $f$,
while the data of~\cite{LoboGiorgini} for the pair correlation function together with the relation~(\ref{eq:C_g2}) give $\zeta\simeq0.95$.\footnote{This value is also compatible with the data of~\cite{LoboGiorgini} for the one-body density matrix, whose short-range singular part is related by Fourier transformation to the large-$k$ tail of the momentum distribution~\cite{Tangen}.}
At unitarity,
the contact $C$ of the trapped gas is directly related to the contact of the homogeneous gas, i.e. to $\zeta$:
Indeed the average over the trap can be done analytically within
the local density approximation,
yielding \cite{Tarruell}
\begin{equation}
\frac{C}{N k_F^{\rm trap}}=\frac{512}{175}\,\frac{\zeta}{\xi^{1/4}}.
\end{equation}
In conclusion, the smallness of the interaction range leads to mathematical singularities; at first sight this may seem to complicate the problem as compared to other strongly interacting systems;
however these singularities are well understood and have a useful consequence:
the existence of exact relations resulting from the Hellmann-Feynman theorem
\cite{Lieb} and from properties of the Fourier transform \cite{Olshanii}.
In particular this provides a ``useful check on mutual consistency of various experiments'', as foreseen in~\cite{Leggett_IHP}.
\section{Dynamical symmetry of the unitary gas
}
\label{sec:ds}
In this section, we present some remarkable properties
of the unitary gas,
derived
from the zero-range model.
The starting point is that the time evolution of the gas
in a
time dependent
isotropic harmonic trap
may be expressed exactly
in terms of a gauge and scaling transform, see subsection \ref{subsec:tds}.
This implies the existence of a SO(2,1) dynamical (or hidden)
symmetry of the system, a formal property that we shall link to concrete
consequences, such as the existence of an exactly decoupled
bosonic degree of freedom (the breathing mode of the gas),
see subsection \ref{subsec:so21},
or the separability of the $N$-body wavefunction in hyperspherical
coordinates, see subsection \ref{subsec:separability},
which holds both in an isotropic harmonic trap and in free space and
has several important consequences
such as the analytical solution of the trapped three-body
problem, see subsection \ref{subsec:pcots}.
In subsec.~\ref{subsec:viscosity} we
use the existence of the undamped breathing mode
to rederive a remarkable property of the homogeneous unitary gas: its bulk viscosity vanishes.
Subsection~\ref{subsec:scaling_laws} concerns short-range scaling laws, which are related to the separability in hyperspherical coordinates, but hold for any scattering length and external potential.
\subsection{Scaling solution in a time-dependent trap}
\label{subsec:tds}
In this section, we shall assume that the trapping potential
$U(\mathbf{r})$ introduced in Eq.(\ref{eq:hamil}) is an isotropic harmonic
potential. Whereas the hypothesis of harmonicity
may be a good approximation in present experiments for small enough
atomic clouds, the isotropy is not granted and requires some
experimental tuning that, to our knowledge, remains to be done.
On the other hand, we allow a general time dependence of the trap
curvature, so that Schr\"odinger's equation for the $N$-body
wavefunction defined in Eq.(\ref{eq:def_psi}) is
\begin{equation}
\label{eq:tdse}
i\hbar \partial_t \psi(\mathbf{X},t) =
\left[-\frac{\hbar^2}{2m} \Delta_\mathbf{X} + \frac{1}{2} m \omega^2(t) X^2
\right] \psi(\mathbf{X},t),
\end{equation}
where we recall that $\mathbf{X}$ is the set of all particle coordinates,
and $\omega(t)$ is the instantaneous angular oscillation frequency.
The interaction between particles is described by the contact
conditions Eq.(\ref{eq:bpN}), written here for the unitary
gas, that is for $a^{-1}=0$:
\begin{equation}
\label{eq:bpu}
\psi(\mathbf{X}) = \frac{A_{ij} (\mathbf{R}_{ij}; (\mathbf{r}_k)_{k\neq i,j})}{r_{ij}}+O(r_{ij}).
\end{equation}
Let us consider the particular case, quite relevant experimentally,
where the gas is initially at equilibrium in a static trap
$\omega(t=0)=\omega$. The gas is then in a statistical mixture of
stationary states, so we can assume that the initial $N$-body
wavefunction is an eigenstate of the Hamiltonian with energy $E$.
At $t>0$, the trap curvature is varied, which leads to an arbitrary time
dependent $\omega(t)$. In typical experiments, one either sets abruptly
$\omega(t)$ to zero, to perform a time of flight measurement,
or one modulates $\omega(t)$ at some frequency to study
the gas collective modes. Can we predict the evolution of the system~?
As shown in \cite{CRASCastin}, the answer is yes, as we now explain.
In the absence of interactions, it is well known \cite{scaling_ideal}
that $\psi(\mathbf{X},t)$ is deduced from the $t=0$ wavefunction by
a simple gauge plus scaling ansatz:
\begin{equation}
\label{eq:ansatz}
\psi(\mathbf{X},t) = \frac{e^{i\theta(t)}}{\lambda^{3N/2}(t)}
\exp\left[\frac{im\dot\lambda(t)}{2\hbar\lambda(t)} X^2\right]
\psi(\mathbf{X}/\lambda(t),0),
\end{equation}
where $\dot\lambda(t)=d\lambda(t)/dt$.
At time $t=0$, one clearly has $\theta(0)=0$,
\begin{equation}
\label{eq:inco}
\lambda(0) = 1 \ \ \ \ \mbox{and} \ \ \ \dot\lambda(0) = 0.
\end{equation}
Inserting this ansatz into Schr\"odinger's equation (\ref{eq:tdse}),
we obtain a Newton like equation of motion for $\lambda$:
\begin{equation}
\label{eq:russe}
\ddot\lambda(t) = \frac{\omega^2}{\lambda^3(t)} -\omega^2(t) \lambda(t)
\end{equation}
to be solved with the initial conditions (\ref{eq:inco}).
We recall that $\omega$ stands for the {\sl initial} angular
oscillation frequency.
The equation (\ref{eq:russe}) is well studied in the literature,
under the name of the Ermakov equation
\cite{eq_russe}, and is in particular amenable to a linear form:
One recognizes an equation for the distance to the origin
for a two-dimensional harmonic oscillator
of angular frequency $\omega(t)$, as
obtained from Newton's equation and from the law of equal areas.
In particular, if $\omega(t)=\omega_{\rm ct}$ is a constant over some time
interval, $\lambda(t)$ oscillates
with a period $\pi/\omega_{\rm ct}$ over that time interval.
The global phase $\theta(t)$ is given by
\begin{equation}
\label{eq:tmod}
\theta(t) = -\frac{E}{\hbar} \int_0^{t} \frac{dt'}{\lambda^2(t')}.
\end{equation}
This suggests that $\theta$ still evolves at the stationary pace
$-E/\hbar$ provided that one introduces a modified time,
as done in \cite{Shlyapnikov} in a bosonic mean field context:
\begin{equation}
\label{eq:modt}
\tau(t) = \int_0^{t} \frac{dt'}{\lambda^2(t')}.
\end{equation}
We shall come back to this point below.
In presence of interactions, one has to check that the ansatz
(\ref{eq:ansatz}) obeys the contact conditions (\ref{eq:bpu}).
First, the ansatz includes a scaling transform. As discussed
in subsection \ref{subsec:bp}, this preserves the contact
conditions and the domain of the Hamiltonian
for the unitary gas. Second, the ansatz includes a quadratic gauge
transform. Turning back to the definition of the contact conditions,
we select an arbitrary pair of particles
$i$ and $j$ and we take the limit $r_{ij}\to 0$ for a fixed centroid
position $\mathbf{R}_{ij}=(\mathbf{r}_i+\mathbf{r}_j)/2$. In the gauge factor,
the quantity $X^2 = \sum_{k=1}^{N} r_k^2$ appears.
The positions $\mathbf{r}_k$ of the particles other than $i$ and $j$
are fixed. What matters is thus $r_i^2 + r_j^2$ that we rewrite as
\begin{equation}
r_i^2 + r_j^2 = 2 R_{ij}^2 + \frac{1}{2} r_{ij}^2.
\end{equation}
$R_{ij}$ is fixed. $r_{ij}$ varies but it appears squared in the gauge
transform, so that
\begin{equation}
\label{eq:chrcbp}
\exp\left[\frac{im\dot\lambda(t)}{2\hbar\lambda(t)} r_{ij}^2/2\right]
\left[\frac{1}{r_{ij}} + O(r_{ij})\right] = \frac{1}{r_{ij}} + O(r_{ij})
\end{equation}
and the contact conditions are preserved by the gauge
transform, even if the scattering length $a$ was finite.
We thus conclude that the ansatz (\ref{eq:ansatz}) gives
the solution also for the unitary gas. This has interesting practical
consequences. For measurements in position space,
one has simple scaling relations,
not only for the mean density $\rho_\sigma(\mathbf{r},t)$
in each spin component $\sigma$:
\begin{equation}
\rho_\sigma(\mathbf{r},t) = \frac{1}{\lambda^{3}(t)}\, \rho_\sigma(\mathbf{r}/\lambda(t),0)
\end{equation}
but also for higher order density correlation functions:
For example, the second order density correlation function
defined in terms of the fermionic field operators as
\begin{equation}
g^{(2)}_{\sigma\sigma'}(\mathbf{r},\mathbf{r}') \equiv
\langle \psi_\sigma^\dagger(\mathbf{r}) \psi_{\sigma'}^\dagger(\mathbf{r}')
\psi_{\sigma'}(\mathbf{r}') \psi_\sigma(\mathbf{r})\rangle,
\end{equation}
evolves in time according to the scaling
\begin{equation}
g^{(2)}_{\sigma\sigma'}(\mathbf{r},\mathbf{r}',t) = \frac{1}{\lambda^6(t)}
g^{(2)}_{\sigma\sigma'}(\mathbf{r}/\lambda(t),\mathbf{r}'/\lambda(t),0).
\end{equation}
As a consequence, if one abruptly switches off the trapping potential
at $t=0^+$, the gas experiences a ballistic expansion with a
scaling factor
\begin{equation}
\lambda(t) = [1+\omega^2 t^2]^{1/2},
\end{equation}
which acts as a perfect magnifying lens on the density distribution.
For non-diagonal observables in position space, some information
is also obtained, with the gauge transform now contributing.
E.g.\ the first order coherence function
\begin{equation}
g^{(1)}_{\sigma\sigma} (\mathbf{r},\mathbf{r}') \equiv
\langle \psi_{\sigma}^\dagger(\mathbf{r}') \psi_\sigma(\mathbf{r}) \rangle,
\end{equation}
which is simply the matrix element of the one-body density operator
between $\langle \mathbf{r},\sigma|$ and $|\mathbf{r}',\sigma\rangle$,
evolves according to
\begin{equation}
g^{(1)}_{\sigma\sigma} (\mathbf{r},\mathbf{r}',t)=
\frac{1}{\lambda^3(t)}
\exp\left[\frac{im\dot\lambda(t)}{2\hbar\lambda(t)} (r^2-r'^{2})\right]
g^{(1)}_{\sigma\sigma}(\mathbf{r}/\lambda(t),\mathbf{r}'/\lambda(t),0).
\end{equation}
The momentum distribution $n_\sigma(\mathbf{k})$ in the spin component $\sigma$
is the Fourier transform over $\mathbf{r}-\mathbf{r}'$ and the integral over
$(\mathbf{r}+\mathbf{r}')/2$ of the first order coherence function.
For a ballistic expansion, directly transposing to three dimensions
the result obtained in \cite{Gangardt} from a time dependent scaling
solution for the one-dimensional gas of impenetrable bosons,
one has that the momentum distribution
of the ballistically expanding unitary gas is asymptotically
homothetic to the gas initial spatial density profile:
\begin{equation}
\lim_{t\to +\infty} n_\sigma(\mathbf{k},t) =
\left(\frac{2\pi\hbar}{m\omega}\right)^{3}
\rho_\sigma\left(\mathbf{r}=\frac{\hbar\mathbf{k}}{m\omega},0\right).
\end{equation}
We emphasize that the above results hold for an arbitrary gas polarization,
that is for arbitrary numbers of particles in each of the two
spin states $\sigma=\uparrow,\downarrow$. If the initial state is
thermal, they hold whatever the value of the temperature, larger
or smaller than the critical temperature $T_c$.
These results however require the unitary limit (in particular
$|a|=+\infty$) and a perfect isotropy of the harmonic trap.
If the experimental goal is simply to have the ballistic expansion
as a perfect magnifying lens, these two requirements
remarkably may be removed, as shown in \cite{Lobo}, if one is ready
to impose an appropriate time dependence to the scattering length
$a(t)$ and to the trap aspect ratio, in which case
the ansatz (\ref{eq:ansatz}) holds at all times. In the particular case of
an isotropic trap, the procedure of \cite{Lobo} is straightforward
to explain: If $\psi(t=0)$ obeys the contact conditions with a
finite scattering length $a$, the ansatz
(\ref{eq:ansatz}) obeys the contact conditions for a scattering
length $\lambda(t)a$ so one simply has to adjust the actual scattering
length in a time dependent way:
\begin{equation}
a(t) = \lambda(t) a
\end{equation}
where $\lambda$ evolves according to Eq.(\ref{eq:russe}).
As shown in the next subsection,
the time dependent solution in the unitary case,
apart from providing convenient scaling relations on the density,
is connected to several interesting intrinsic properties of the system,
whereas the procedure of \cite{Lobo} does not imply such properties.
To be complete, we finally address the general case where
the initial wavefunction of the unitary gas is not necessarily
a stationary state but is arbitrary \cite{WernerPRA}.
Then the observables of the gas have in general a non-trivial time dependence,
even for a fixed trap curvature.
If the trap curvature is time dependent, we modify
the gauge plus scaling ansatz as follows:
\begin{equation}
\psi(\mathbf{X},t) = \frac{1}{\lambda^{3N/2}(t)}
\exp\left[\frac{im\dot\lambda(t)}{2\hbar\lambda(t)} X^2\right]
\tilde{\psi}(\mathbf{X}/\lambda(t),\tau(t)),
\end{equation}
where $\tau(t)$ is the modified time introduced in Eq.(\ref{eq:tmod}),
$\lambda(t)$ evolves according to Eq.(\ref{eq:russe}) with the initial
conditions (\ref{eq:inco}),
and the time-dependent wavefunction $\tilde{\psi}$ coincides
with $\psi$ at time $t=\tau=0$ and obeys the unitary gas contact
conditions. Then this ansatz obeys the contact conditions.
When inserted in the time dependent Schr\"odinger equation
(\ref{eq:tdse}), it leads to a Schr\"odinger equation
for $\tilde{\psi}$ in the time independent external potential
fixed to the $t=0$ trap:
\begin{equation}
i\hbar \partial_\tau \tilde{\psi}(\mathbf{X},\tau)=
\left[-\frac{\hbar^2}{2m} \Delta_\mathbf{X} + \frac{1}{2} m \omega^2 X^2
\right] \tilde{\psi}(\mathbf{X},\tau).
\end{equation}
The gauge plus scaling transform, and the redefinition of time,
have then totally cancelled the time dependence of the trap.
If the initial wavefunction is an eigenstate of energy $E$,
as was previously the case, one simple has
$\tilde{\psi}(\tau) = \exp(-i E\tau/\hbar) \psi(t=0)$
and one recovers the global phase factor in Eq.(\ref{eq:ansatz}).
\subsection{SO(2,1) dynamical symmetry and the decoupled breathing mode}
\label{subsec:so21}
As shown in \cite{Rosch} for a two-dimensional Bose gas with $1/r^2$ interactions,
the existence of a scaling solution such as Eq.(\ref{eq:ansatz})
reflects a hidden symmetry of the Hamiltonian, the SO(2,1)
dynamical symmetry. Following \cite{WernerPRA},
we construct explicitly this dynamical symmetry
for the unitary gas and we show that it has interesting consequences
for the energy spectrum in a static isotropic harmonic trap.
Let us consider a gedankenexperiment: Starting from the unitary gas
is an energy eigenstate $\psi$, we modify in an infinitesimal
way the trap curvature during the time interval $[0,t_f]$, and for $t>t_f$
we restore the initial trap curvature, $\omega(t)=\omega(0)=\omega$.
Linearizing Eq.(\ref{eq:russe}) around $\lambda=1$ for $t>t_f$,
we see that the resulting change in the scaling parameter $\lambda$ is
\begin{equation}
\lambda(t)-1 = \epsilon \, e^{-2i\omega t} + \epsilon^* \ e^{2i\omega t}
+O(\epsilon^2)
\label{eq:mode_castin}
\end{equation}
where $\epsilon$ is proportional to the infinitesimal curvature change.
Since $\lambda$ oscillates indefinitely at frequency $2\omega$,
this shows the existence of an undamped mode of frequency $2\omega$.
This conclusion actually extends
to excitations during $[0,t_f]$ of arbitrarily large amplitudes, as noted
below Eq.(\ref{eq:russe}) \cite{Rosch}.
We calculate the resulting change in the $N$-body wavefunction,
expanding Eq.(\ref{eq:ansatz}) to first order in $\epsilon$,
putting in evidence the components that oscillate with Bohr
frequencies $\pm 2\omega$:
\begin{equation}
\label{eq:cwf}
\psi(\mathbf{X},t) = e^{i\alpha}\left[e^{-iEt/\hbar}-\epsilon e^{-i(E+2\hbar\omega)t/\hbar}
L_+ + \epsilon^* e^{-i(E-2\hbar\omega)t/\hbar} L_-\right]
\psi(\mathbf{X},0) + O(\epsilon^2).
\end{equation}
The time independent phase $\alpha$ depends on the details of the excitation procedure.
We have introduced the operators
\begin{equation}
L_{\pm} = \pm i D + \frac{H}{\hbar\omega} - \frac{m\omega}{\hbar} X^2
\end{equation}
where $D$ is the generator of the scaling transforms, as defined
in Eq.(\ref{eq:defD}), and $L_+=L_-^\dagger$.
We then read on Eq.(\ref{eq:cwf}) the remarkable property that the
action of $L_+$ on an energy eigenstate $\psi$ of energy $E$ produces
an energy eigenstate of energy $E+2 \hbar \omega$ \footnote
{As shown in \cite{WernerPRA}, $L_+ \psi$ cannot be zero.}.
Similarly, the action of $L_-$ on $\psi$ produces an energy eigenstate
of energy $E-2\hbar \omega$, or eventually gives zero since
the spectrum is bounded from below by $E\ge 0$ according to the
virial theorem (\ref{eq:virth}) applied to $U(\mathbf{r})=\frac{1}{2} m\omega^2 r^2$.
We see that the spectrum has thus a very simple structure, it is a
collection of semi-infinite ladders, each ladder being made
of equidistant energy levels separated by $2\hbar \omega$,
see Fig.~\ref{fig:ladder}, and $L_\pm$ acting respectively as
a raising/lowering operator in that structure.
Within each ladder, we call $\psi_g$ the wavefunction corresponding
to the ground step of that ladder, such that
\begin{equation}
\label{eq:innocent}
L_- \psi_g = 0.
\end{equation}
\begin{figure}[b]
\sidecaption
\includegraphics[width=4cm,clip=]{ladder.eps}
\caption{
The energy spectrum of the unitary gas in an isotropic harmonic trap
is a collection of semi-infinite ladders such as the one depicted
in the figure, with various ground step energies $E_g$.
This structure is related to the existence of
a decoupled bosonic mode, and holds whatever the numbers of fermions in each
of the two spin components.
\label{fig:ladder}
}
\end{figure}
As shown in \cite{Rosch}, this structure implies a dynamical SO(2,1) symmetry,
meaning that the Hamiltonian $H$ is part of the SO(2,1) Lie algebra.
One starts with the commutation relations:
\begin{equation}
[H,L_{\pm}] = \pm 2 \hbar \omega L_{\pm}
\end{equation}
\begin{equation}
[ {L}_{+},{L}_{-} ] = -4 H /(\hbar\omega).
\label{eq:comm_ut}
\end{equation}
The first relation was expected from the raising/lowering nature
of $L_{\pm}$. Both relations can be checked from the commutation relations
Eqs.(\ref{eq:comm1},\ref{eq:comm2}) and from
\begin{equation}
[\frac{1}{2}X^2, -\frac{1}{2}\Delta_\mathbf{X}] = i D.
\end{equation}
We emphasize again the crucial point that the operators $L_\pm$
preserve the domain of the Hamiltonian in the present unitary case,
since $D$ and $X^2$ do. Obtaining the canonical commutation
relations among the generators $T_1,T_2$ and $T_3$ of the SO(2,1)
Lie algebra,
\begin{equation}
\label{eq:la}
[T_1,T_2] \equiv -i T_3, \ \ \ [T_2,T_3]\equiv i T_1, \ \ \
[T_3,T_1] \equiv i T_2,
\end{equation}
is then only a matter of rewriting:
\begin{equation}
T_1 \pm i T_2 \equiv \frac{1}{2} L_\pm \ \ \ \
\mbox{and}\ \ \ \ T_3 = \frac{H}{2\hbar\omega}.
\end{equation}
Note the sign difference in the first commutator of Eq.(\ref{eq:la})
with respect to the other two commutators, and with respect to the
more usual SO(3) or SU(2) Lie algebra.
Have we gained something in introducing the SO(2,1) Lie algebra,
or is it simply a formal rewriting of the ladder structure
already apparent in the simple minded approach Eq.(\ref{eq:cwf}),
may ask a reader unfamiliar with dynamical symmetries.
Well, an advantage is that we can immediately exhibit the so-called Casimir
operator $C$,
\begin{equation}
C \equiv -4 [T_1^2 + T_2^2 - T_3^2] = H^2-\frac{1}{2} (\hbar\omega)^2
(L_+ L_- + L_- L_+),
\end{equation}
guaranteed to commute with all the elements $T_1$, $T_2$ and
$T_3$ of the algebra, so that $C$ is necessarily a scalar within each
ladder. Taking as a particularly simple case the expectation value
of $C$ within the ground step $\psi_g$ of the ladder of energy
$E_g$, and using
Eq.(\ref{eq:comm_ut}) to evaluate $\langle \psi_g| L_- L_+ |\psi_g\rangle$,
we obtain $C |\psi_g\rangle = E_g (E_g - 2 \hbar \omega) |\psi_g\rangle$.
Inverting this relation thanks to the property
$E_g \ge 3\hbar\omega/2$
\footnote{To obtain this inequality, one uses a virial theorem after
separation of the center of mass motion \cite{WernerPRA}.},
we can define the ground energy step operator $H_g$:
\begin{equation}
H_g = \hbar\omega + [C+(\hbar\omega)^2]^{1/2},
\end{equation}
which is scalar and equal to $E_g$ within each ladder.
A useful application of $H_g$ is to rescale the raising
and lowering operator $L_\pm$ to obtain simpler commutation
relations: It appears that
\begin{equation}
b = [2(H+H_g)/(\hbar\omega)]^{-1/2} L_-
\end{equation}
is a bosonic annihilation operator, which obeys the usual
bosonic commutation relations, in particular with its hermitian conjugate
\begin{equation}
[b,b^\dagger] = 1.
\end{equation}
$b^\dagger $ and $b$ have the same raising/lowering properties
as $L_\pm$, and commute with $H_g$. They have the usual simple
matrix elements, e.g.\ $b^\dagger|n\rangle = (n+1)^{1/2} |n+1\rangle$
where $|n\rangle$ is the step number $n$ of a ladder, $n$ starting from
0. They allow an illuminating rewriting of the Hamiltonian:
\begin{equation}
H = H_g + 2\hbar\omega b^\dagger b
\end{equation}
revealing that the unitary gas in a harmonic isotropic trap
has a fully decoupled bosonic degree of freedom.
This bosonic degree of freedom, physically, is simply
the undamped breathing mode of the gas of frequency $2\omega$,
identified for a different system in \cite{Rosch}.
We now give two simple applications of the above formalism \cite{WernerPRA}.
First, one can calculate the various moments of the trapping Hamiltonian
$H_{\rm trap}= \frac{1}{2} m \omega^2 X^2$, from the identity
\begin{equation}
\label{eq:htrap}
H_{\rm trap} = \frac{1}{2} H -\frac{\hbar\omega}{4} (L_+ + L_-)
= \frac{\hbar\omega}{2} A^\dagger A
\end{equation}
where $A=[b^\dagger b+H_g/(\hbar\omega)]^{1/2}-b$.
Taking the expectation value of Eq.(\ref{eq:htrap}) within a given
eigenstate of energy $E$, or within a statistical mixture of eigenstates,
immediately gives
\begin{equation}
\langle H_{\rm trap}\rangle = \frac{1}{2} \langle H\rangle,
\end{equation}
a particular case of the virial theorem Eq.(\ref{eq:virth}).
Taking the expectation value of $H_{\rm trap}^2$ for the thermal
equilibrium density operator gives
\begin{equation}
4 \langle H_{\rm trap}^2 \rangle = \langle H^2 \rangle
+ \langle H\rangle \hbar\omega [2\langle b^\dagger b\rangle +1]
\end{equation}
where we used $\langle H_g b^\dagger b\rangle= \langle H_g\rangle \langle
b^\dagger b\rangle$ for the thermal equilibrium.
From the Bose formula, one has also
$\langle b^\dagger b\rangle=[\exp(2\beta\hbar\omega)-1]^{-1}$,
with $\beta=1/(k_B T)$ and $T$ is the temperature.
The second, more impressive,
application is to uncover a very interesting structure
of the $N$-body wavefunction $\psi_g(\mathbf{X})$ of the ground energy step
of an arbitrary ladder. We introduce hyperspherical coordinates
$(X,\mathbf{n}=\mathbf{X}/X)$, where $\mathbf{n}$ is a unit vector is the space of
$3N$ real coordinates. The innocent equation (\ref{eq:innocent}) becomes
\begin{equation}
\label{eq:innocent2}
\left[-\frac{3N}{2} - X\partial_X +\frac{E_g}{\hbar\omega}
-\frac{m\omega}{\hbar} X^2 \right] \psi_g(\mathbf{X})=0.
\end{equation}
This is readily integrated for a fixed hyperdirection $\mathbf{n}$:
\begin{equation}
\label{eq:fascinating}
\psi_g(\mathbf{X}) = e^{-m\omega X^2/(2\hbar)}
X^{\frac{E_g}{\hbar\omega} -\frac{3N}{2}}
f(\mathbf{n})
\end{equation}
where $f(\mathbf{n})$ is an unknown function of the hyperdirection.
Eq.(\ref{eq:fascinating}) has fascinating consequences.
First, it shows that $\psi_g$, being the product of a function
of the modulus $X$ and of a function of the hyperdirection,
is {\sl separable} in hyperspherical coordinates. The physical consequences
of this separability, in particular for the few-body problem,
are investigated in subsection \ref{subsec:pcots}.
Note that this separability holds for all the other steps of the ladder,
since $L_+$ only acts on the hyperradius.
Second, we take the limit $\omega\to 0$ in Eq.(\ref{eq:fascinating}):
According to Eq.(\ref{eq:simple}), $E_g/(\hbar\omega)$ is a constant,
and $E_g\to 0$, whereas the Gaussian factor tends to unity.
$\mathbf{n}$ is dimensionless, and we can take $f(\mathbf{n})$ to be $\omega$ independent
if we do not normalize $\psi_g$ to unity.
We thus obtain in this limit a zero energy eigenstate
of the free space problem,
\begin{equation}
\label{eq:psifsi}
\psi^{\rm free}(\mathbf{X}) = X^{\frac{E_g}{\hbar\omega} -\frac{3N}{2}} f(\mathbf{n})
\end{equation}
which is independent of $\omega$. This zero energy eigenstate
is scaling invariant, in the sense that
\begin{equation}
\label{eq:sca_inv}
\psi_\lambda^{\rm free} (\mathbf{X}) = \frac{1}{\lambda^\nu} \psi^{\rm free}(\mathbf{X})
\ \ \ \forall \lambda>0,
\end{equation}
where $\psi_\lambda$ is defined in Eq.(\ref{eq:sctr}) and
\begin{equation}
\nu = \frac{E_g}{\hbar\omega}.
\end{equation}
In summary, starting from the wavefunction $\psi_g$ of any ladder ground state of
the trapped gas spectrum,
one gets a scaling-invariant zero-energy free-space eigenstate $\psi_\lambda^{\rm
free}$,
simply by removing the gaussian factor $e^{-m\omega X^2/(2\hbar)}$
in the expression (\ref{eq:fascinating}) of $\psi_g$.
Remarkably, the reverse property is true. Let us imagine that we know
a zero energy eigenstate $\psi^{\rm free}$ of the free space problem
$H_{\rm free}=-\frac{\hbar^2}{2m} \Delta_\mathbf{X}$,
\begin{equation}
\label{eq:eselen}
\Delta_\mathbf{X} \psi^{\rm free}(\mathbf{X}) = 0,
\end{equation}
that of course also obeys the Wigner-Bethe-Peierls contact conditions
for the unitary gas. Since $H_{\rm free}$
commutes with the generator $D$ of the scaling transforms,
we generally expect $\psi^{\rm free}$ to obey Eq.(\ref{eq:sca_inv})
with some exponent $\nu$, so that
\begin{equation}
i D \psi^{\rm free} = \nu \psi^{\rm free}.
\label{eq:dpsi}
\end{equation}
Since $\psi^{\rm free}$ is not square integrable, the hermiticity of
$D$ does not imply that $\nu\in i\mathbb{R}$; on the contrary,
we will see that $\nu\in \mathbb{R}$.
Let us multiply $\psi^{\rm free}$ with a Gaussian factor:
\begin{equation}
\label{eq:mapping}
\psi(\mathbf{X}) \equiv e^{-m\omega X^2/(2\hbar)} \psi^{\rm free}(\mathbf{X}).
\end{equation}
As we did for the gauge transform, see Eq.(\ref{eq:chrcbp}),
we can show that $\psi$ so defined obeys the Wigner-Bethe-Peierls
contact conditions.
Calculating the action on $\psi$ of the Hamiltonian $H$ of the trapped gas,
and using Eq.(\ref{eq:dpsi}), we directly obtain
\begin{equation}
H \psi = \nu \hbar \omega \psi,
\end{equation}
i.e.\ $\psi$ is indeed an eigenstate of the trapped gas with the
eigenenergy $\nu \hbar \omega$.
This $\psi$ corresponds to the ground energy step of a ladder.
Repeated action of $L_+$ will generate the other states of the ladder.
We have thus constructed a mapping between the trapped case
and the zero energy free space case, for the unitary gas in an
isotropic harmonic trap. A similar mapping
(restricting to the ground state) was constructed by Tan in
an unpublished work \cite{TanScaling}.
\subsection{Separability in internal hyperspherical coordinates}
\label{subsec:separability}
As shown in subsection \ref{subsec:so21}, the SO(2,1) dynamical symmetry
of the unitary gas in an isotropic harmonic trap implies that the eigenstate
wavefunctions $\psi(\mathbf{X})$ may be written as the product of a function
of the modulus $X$ and of a function of the direction $\mathbf{X}/X$.
Here, following \cite{WernerPRA}, we directly use this property at the level of the $N$-body Schr\"odinger
equation, for $N>2$, and we derive an effective Schr\"odinger equation for a
hyperradial wavefunction, with interesting consequences discussed
in subsection \ref{subsec:pcots}. The derivation is restricted here to the case of particles of
identical masses, as in the previous sections, but the separability in internal spherical coordinates
may also hold for particles of different masses, as detailed in Appendix~3.
First, we introduce a refinement to the separability of subsection \ref{subsec:so21}: In a harmonic trap,
the center of mass of the system is totally decoupled from the internal
variables, that is from the relative coordinates $\mathbf{r}_i-\mathbf{r}_j$ of the particles.
This is quite straightforward in Heisenberg picture, for an interaction modeled by a potential
$V(|\mathbf{r}_i-\mathbf{r}_j|)$. The Heisenberg equations of motion for the center of mass position
\begin{equation}
\mathbf{C} \equiv \frac{1}{N} \sum_{i=1}^{N} \mathbf{r}_i
\end{equation}
and the center of mass momentum $\mathbf{P}=\sum_{i=1}^{N} \mathbf{p}_i$
are indeed coupled only among themselves, due in particular to the fact that the interaction
potential cannot change the total momentum $\mathbf{P}$ of the system:
\begin{eqnarray}
\frac{d}{dt} \hat{\mathbf{P}}(t) &=& -N m \omega^2 \hat{\mathbf{C}}(t) \\
\frac{d}{dt} \hat{\mathbf{C}}(t) &=& \frac{\hat{\mathbf{P}}(t)}{Nm}.
\end{eqnarray}
The center of mass of the system thus behaves as a fictitious particle of mass $Nm$
trapped in the harmonic potential $N m\omega^2 \mathbf{C}^2/2$, with a Hamiltonian
\begin{equation}
H_{\rm CM} = -\frac{\hbar^2}{2Nm} \Delta_{\mathbf{C}} + \frac{1}{2} N m \omega^2 C^2.
\end{equation}
The center of mass has of course the same
angular oscillation frequency as the individual particles.
This center of mass decoupling property clearly holds in the general harmonic {\sl anisotropic} case.
It persists in the zero range limit so it holds also for the zero-range model.
We can thus split the Hamiltonian Eq.(\ref{eq:hamil})
as the sum of the center of mass Hamiltonian $H_{\rm CM}$ and the internal Hamiltonian
$H_{\rm internal} \equiv H - H_{\rm CM}$.
As a consequence, we introduce as new spatial coordinates the center of mass position
$\mathbf{C}$ and the set of internal coordinates
\begin{equation}
\mathbf{R} \equiv (\mathbf{r}_1-\mathbf{C},\ldots,\mathbf{r}_N-\mathbf{C}),
\end{equation}
and we can seek eigenstates in the factorized form $\psi(\mathbf{X}) = \psi_{\rm CM}(\mathbf{C}) \psi_{\rm internal}(\mathbf{R})$.
The crucial step is then to define {\sl internal} hyperspherical coordinates, consisting in the
hyperradius
\begin{equation}
R = \left[\sum_{i=1}^{N} (\mathbf{r}_i-\mathbf{C})^2\right]^{1/2}
\end{equation}
and a convenient parameterization of the set of dimensionless internal coordinates $\mathbf{R}/R$.
There is a technical subtlety due to the
fact that the coordinates of $\mathbf{R}$ are not independent variables: Since the sum of the components
of $\mathbf{R}$ along each spatial direction $x$, $y$ and $z$ is exactly zero, and since $\mathbf{R}/R$ is a unit vector,
the vector $\mathbf{R}/R$ contains actually only $3N-4$ independent dimensionless real variables.
We then use the following result, that may be obtained with the appropriate Jacobi coordinates
\footnote{For particles of equal masses one introduces the Jacobi coordinates
$\mathbf{u}_i=(\frac{N-i}{N+1-i})^{1/2} \left[\mathbf{r}_i-(N-i)^{-1} \sum_{j=i+1}^{N}\mathbf{r}_j\right]$
for $1\leq i\leq N-1$. Then $\Delta_\mathbf{X} = N^{-1} \Delta_\mathbf{C} + \sum_{i=1}^{N-1} \Delta_{\mathbf{u}_i}$
and $R^2=X^2-N C^2 =\sum_{i=1}^{N-1} \mathbf{u}_i^2$. The general case of arbitrary masses is detailed in
the Appendix~3.}
\cite{Jacobi}:
There exists a parameterization of $\mathbf{R}/R$
by a set of $3N-4$ internal hyperangles that we call $\mbox{\boldmath$\Omega$}$,
such that the internal Hamiltonian takes the form
\begin{equation}
\label{eq:hamil_interne}
H_{\rm internal} = -\frac{\hbar^2}{2m} \left[ \partial_R^2 + \frac{3N-4}{R} \partial_R
+\frac{1}{R^2} \Delta_{\mbox{\boldmath$\Omega$}}\right] + \frac{1}{2} m \omega^2 R^2,
\end{equation}
where $\Delta_{\mbox{\boldmath$\Omega$}}$ is the Laplacian on the unit sphere of dimension $3N-4$.
The expression between square brackets is the standard form for
the usual Laplacian in dimension $d=3N-3$, written in hyperspherical coordinates,
which justifies the name of ``internal hyperspherical coordinates".
The separability in internal spherical coordinates means that
the internal eigenstates in a trap can be written as products of a function of $R$ and a function of $\mbox{\boldmath$\Omega$}$.
This
basically results from the reasoning
below Eq.(\ref{eq:eselen}), with the little twist that one can further assume that
the zero-energy free space eigenstate $\psi^{\rm free}(\mathbf{X})$ has a zero total momentum,
i.\ e.\ it is independent of the center of mass position
\footnote{The reasoning below Eq.(\ref{eq:innocent2}) can also be adapted
by putting the center of mass in its ground state $\psi_{\rm CM}(\mathbf{C})\propto
\exp[-Nm\omega C^2/(2\hbar)]$ and by constructing purely internal raising and lowering operators
of an internal SO(2,1) dynamical symmetry, that do not excite the center of mass
motion contrarily to $L_+$ and $L_-$ \cite{WernerPRA}.}.
The scale invariance Eq.(\ref{eq:sca_inv}) or equivalently Eq.(\ref{eq:dpsi})
then implies
\begin{equation}
\label{eq:psifreeR}
\psi^{\rm free}(\mathbf{X}) = R^{s-(3N-5)/2} \phi(\mbox{\boldmath$\Omega$})
\end{equation}
with some exponent $s$ shifted for convenience by $(3N-5)/2$.
The challenge is of course to determine the unknown function $\phi(\mbox{\boldmath$\Omega$})$ and the corresponding value
of $s$.
From Schr\"odinger's equation $\Delta_\mathbf{X} \psi^{\rm free}=0$ and the expression
of the internal Laplacian in hyperspherical coordinates, see Eq.(\ref{eq:hamil_interne}),
one finds that $s^2$ solves the eigenvalue problem
\begin{equation}
\label{eq:hypera}
\left[-\Delta_{\mbox{\boldmath$\Omega$}} + \left(\frac{3N-5}{2}\right)^2 \right] \phi(\mbox{\boldmath$\Omega$}) = s^2 \phi(\mbox{\boldmath$\Omega$}),
\end{equation}
where $\phi(\mbox{\boldmath$\Omega$})$ has to obey the Wigner-Bethe-Peierls contact conditions Eq.(\ref{eq:bpu})
reformulated in hyperangular coordinates.\footnote{These reformulated contact conditions are given explicitly in~\cite{theseFelix}, Eq.(1.38).} The merit of the shift $(3N-5)/2$ is thus
to reveal a symmetry $s\leftrightarrow -s$.
The generalization of the zero energy free space solution Eq.(\ref{eq:psifreeR})
to the finite energy trapped problem
is simply provided by the ansatz:
\begin{equation}
\psi(\mathbf{X}) = \psi_{\rm CM}(\mathbf{C}) \phi(\mbox{\boldmath$\Omega$}) R^{-(3N-5)/2} F(R).
\label{eq:psisepargen}
\end{equation}
Here $\psi_{\rm CM}(\mathbf{C})$ is any center of mass eigenstate wavefunction of energy $E_{\rm CM}$,
$\phi(\mbox{\boldmath$\Omega$})$ is any solution
of the eigenvalue problem Eq.(\ref{eq:hypera}).
Injecting the ansatz into Schr\"o\-din\-ger's equation of eigenenergy $E$
and using Eq.(\ref{eq:hamil_interne}), one finds that
\begin{equation}
E = E_{\rm CM} + E_{\rm internal},
\end{equation}
where the hyperradial wavefunction $F(R)$ and the internal eigenenergy $E_{\rm internal}$ solve the eigenvalue problem:
\begin{equation}
\label{eq:schr_F}
-\frac{\hbar^2}{2m} \left[F''(R) + \frac{1}{R} F'(R) \right] +
\left(\frac{\hbar^2 s^2}{2m R^2} + \frac{1}{2} m \omega^2 R^2\right) F(R) =
E_{\rm internal} F(R).
\end{equation}
We note that, as detailed in the Appendix~3,
this separability remarkably
also holds in the case where the $N$ particles have different masses \cite{WernerPRA}, provided that they all have
the same angular oscillation frequency $\omega$ in the trap, and that the Wigner-Bethe-Peierls model still defines a self-adjoint Hamiltonian
for the considered mass ratios. The separability even holds when the Wigner-Bethe-Peierls model {\it supplemented with an additional boundary condition for $R\to0$ and fixed $\mbox{\boldmath$\Omega$}$} is self-adjoint, as is the case e.g. for $N=3$ bosons, see below; indeed such a boundary condition only affects the hyperradial problem.
In practice the explicit calculation of $s$ is possible
for the few-body problem.
The most natural approach in general is to try to calculate the functions $A_{ij}$ in Eq.(\ref{eq:bpu}) in momentum space.
From Eq.(\ref{eq:bpu}) it appears that $A_{ij}$ is scaling invariant with an exponent
$s+1-(3N-5)/2$. Its Fourier transform\footnote{Since the Fourier transform
$\tilde{A}_{ij}(\mathbf{K})=\int d^{3(N-2)}Y\, e^{-i\mathbf{K}\cdot\mathbf{Y}} A_{ij}(\mathbf{Y})$ may lead to non-absolutely
converging integrals at infinity, the calculation has to be performed using the language of distributions,
with a regularizing factor $e^{-\eta Y}$, $\eta\to 0^+$.}
is then also scaling invariant, with an exponent given
by a simple power-counting argument: Since $A_{ij}$ is a function of $3(N-2)$ variables,
if one takes into account the fact that it does not depend on the center of mass position $\mathbf{C}$, and
since one has $[s+1-(3N-5)/2]+3(N-2)=s+(3N-5)/2$, its Fourier transform $\tilde{A}_{ij}$ scales as
\begin{equation}
\label{eq:Afs}
\tilde{A}_{ij}(\mathbf{K}) = K^{-[s+(3N-5)/2]} f_{ij}(\mathbf{K}/K)
\end{equation}
where $\mathbf{K}$ collects all the $3(N-2)$ variables of $\tilde{A}_{ij}$
and $f_{ij}$ denotes some functions to be determined.
Remarkably it is the same quantity $(3N-5)/2$ which appears in both Eqs.(\ref{eq:psifreeR},\ref{eq:Afs}).
This momentum space approach leads to integral equations.
For $N=3$, this integral equation was obtained in~\cite{TerMartirosian}; it was solved analytically in~\cite{Danilov},
the allowed values of $s$ being the solutions of a transcendental equation.
This transcendental equation was rederived from a direct analytical solution of (\ref{eq:hypera}) in position space in \cite{Efimov70},
and generalised to arbitrary angular momenta, masses and statistics in~\cite{Efimov70,Efimov73};
for equal masses it is conveniently expressed in the form~(\cite{Birse} and refs. therein):
\begin{equation}
\frac{\Gamma\left( l+\frac{3}{2} \right)}
{\Gamma\left(\frac{l+1+s}{2}\right) \Gamma\left(\frac{l+1-s}{2}\right)}
=
\frac{\eta}{\sqrt{3\pi}(-2)^l}\ _2F_1\left(\frac{l+1+s}{2},\frac{l+1-s}{2};l+\frac{3}{2};\frac{1}{4}\right)
\label{eq:trans_hypergeo}
\end{equation}
or alternatively~\cite{WernerPRL}
\begin{eqnarray*}
\Bigg[i^l\sum_{k=0}^l\frac{(-l)_k(l+1)_k}{k!}\frac{(1\!-\!s)_l}{(1\!-\!s)_k}
\Big(
2^{-k}i(k-s)e^{is\frac{\pi}{2}}
+\eta (-1)^l\frac{4}{\sqrt{3}}e^{i\frac{\pi}{6}(2k+s)}
\Big) \Bigg]
\nonumber
\end{eqnarray*}
\begin{eqnarray*}
-\Bigg[(-i)^l\sum_{k=0}^l\frac{(-l)_k(l+1)_k}{k!}\frac{(1\!-\!s)_l}{(1\!-\!s)_k}
\Big(
2^{-k}(-i)(k-s)e^{-is\frac{\pi}{2}}
+\eta (-1)^l\frac{4}{\sqrt{3}}e^{-i\frac{\pi}{6}(2k+s)}
\Big) \Bigg] \nonumber
\end{eqnarray*}
\begin{equation}
\ \ \ \ =0
\label{eq:trans}
\end{equation}
where $l$ is the total internal angular momentum quantum number,
$\eta$ is $-1$ for fermions ($N_\uparrow=2$, $N_\downarrow=1$) or $+2$ for spinless bosons,
$\,_2F_1$ is a hypergeometric function,
and
$(x)_n\equiv x (x+1)\ldots (x+n-1)$ with $(x)_0\equiv 1$.
The equation (\ref{eq:trans_hypergeo}) has some spurious integer solutions
($l=0,s=2$ for fermions; $l=0,s=4$ and $l=1,s=3$ for bosons)
which must be eliminated.
For $N=4$ there is no known analytical solution of the integral equation.
Using the scale invariance of $\tilde{A}_{ij}(\mathbf{K})$ as in Eq.(\ref{eq:Afs}) and rotational
symmetry however brings it to a numerically tractable integral equation involving the exponent
$s$, that allowed to predict a four-body Efimov effect for three same-spin state
fermions interacting with a lighter particle \cite{CMP}.
\subsection{Physical consequences of the separability}
\label{subsec:pcots}
As seen in the previous section \ref{subsec:separability}, the solution of the $N$-body problem ($N>2$)
for the unitary gas in a harmonic isotropic trap boils down to (i) the calculation
of exponents $s$ from zero-energy free space solutions, and (ii) the solution of the hyperradial eigenvalue
problem Eq.(\ref{eq:schr_F}). Whereas (i) is the most challenging part on a practical point of
view, the step (ii) contains a rich physics that we now discuss.
Formally, the hyperradial problem Eq.(\ref{eq:schr_F}) is Schr\"odinger's equation for one (fictitious) particle
moving in two dimensions with zero angular momentum in the (effective) potential
\begin{equation}
U_{\rm eff}(R)=\frac{\hbar^2}{2m}\,\frac{s^2}{R^2}
+ \frac{1}{2}m\omega^2 R^2
.
\label{Ueff}
\end{equation}
We will see that the nature of this problem is very different depending on the sign of $s^2$.
The case $s^2\ge0$, i.e. $s$ real, happens
for $N=3$ fermions ($N_\uparrow=2$,\,$N_\downarrow=1$),
not only for equal masses [as can be tested numerically from (\ref{eq:trans}) and even demonstrated analytically from the corresponding hyperangular eigenvalue problem~\cite{WernerPRL}] but also for unequal masses
provided $m_\uparrow/m_\downarrow$ is below the critical value $13.60\ldots$ where one of the $s$
(in the angular momentum $l=1$ channel) becomes imaginary~\cite{Efimov73}.
For $N=4$ fermions with ($N_\uparrow=3$, $N_\downarrow=1$),
the critical mass ratio above which one of the $s$ (in the angular momentum $l=1$ channel) becomes imaginary is slightly smaller, $m_\uparrow/m_\downarrow\simeq13.384$~\cite{CMP}.
In the physics literature,
$s$ is believed to be real for fermions for any $(N_\uparrow, N_\downarrow)$ for equal masses,
this belief being supported by numerical and experimental evidence.
For $3$ identical bosons, it is well-known that one of the values of $s$ (in the $l=0$ channel) is imaginary~\cite{Efimov70}, all other values being real.
\subsubsection{Universal case}
In this subsection we assume that $s$ is real and we can take the sign convention $s\ge0$.
We impose that the hyperradial wavefunction $F(R)$ is bounded for $R\to0$; indeed, allowing $F(R)$ to diverge would physically correspond to a $N$-body resonance
(see Appendix~6).
The spectrum and
the corresponding hyperradial wavefunctions
then are~\cite{WernerPRA}
\begin{equation}
E_{\rm internal} = (s+1+2 q)\hbar\,\omega,\ \ \ q\in\mathbb{N}
\label{eq:Eint}
\end{equation}
\begin{equation}
F(R)=
\sqrt{\frac{2\,q !}{\Gamma(s+1+q)}}\
\frac{R^s}{(a_{ho})^{s+1}}
\, e^{-\left(\frac{R}{2 a_{ho}}\right)^2}
L_q^{(s)}\!\left( \left(\frac{R}{a_{ho}}\right)^2\right)
\label{eq:F(R)}
\end{equation}
where $L_q^{(s)}$ is a generalised Laguerre polynomial of order $q$, $a_{ho}\equiv\sqrt{\frac{\hbar}{m\omega}}$ is the harmonic oscillator length, and the normalisation is such that
$\int_0^\infty dR\, R\, F(R)^2=1$.
Eq.~(\ref{eq:Eint}) generalises to excited states the result obtained
for the ground state
in~\cite{TanScaling}.
We thus recover the $2\hbar\omega$ spacing of the spectrum discussed in section~\ref{subsec:so21}. We can also reinterpret the scaling solution of section~\ref{subsec:tds} as a time-evolution of the hyperradial wavefunction with a time-independent hyperangular wavefunction; in particular, the undamped breathing mode corresponds to an oscillation of the fictitious particle in the effective potential (\ref{Ueff}).\footnote{Strictly speaking, such a time evolution of the wavefunction in {\it internal} hyperspherical coordinates corresponds to an internal scaling solution where the center of mass wavefunction is constant, whereas the scaling solution of~\ref{subsec:tds} corresponds to a hyperradial motion in the hyperspherical coordinates $(X,\mathbf{n})$.}
The expression (\ref{eq:F(R)}) of $F(R)$ immediately yields the probability distribution $P(R)$ of the hyperradius
via $P(R)=F(R)^2\,R$. This analytical prediction is in good agreement with the numerical results obtained in~\cite{Blume} for up to $17$ fermions.
In the large $N$ limit (more precisely if $N_\uparrow$ and $N_\downarrow$ tend to infinity and their ratio goes to a constant),
the ground state energy of the trapped unitary gas is expected to be given in an asymptotically exact way by hydrostatics (also called local density approximation).
Amusingly, this allows to predict the large-$N$ asymptotics of the smallest possible value of $s$.
For $N_\uparrow=N_\downarrow=N/2\to\infty$ this gives~\cite{TanScaling,Blume}
\begin{equation}
s\sim\sqrt{\xi}\frac{(3N)^{4/3}}{4}
\end{equation}
where $\xi$ appears in the expression
Eq.(\ref{eq:e0}) for the ground state energy of the homogeneous unitary gas.
For spin-1/2 fermions, Eqs. (\ref{eq:Eint},\ref{eq:F(R)}), combined with
the transcendental equation (\ref{eq:trans_hypergeo})
and the expression of the hyperangular wavefunctions~\cite{Efimov73}, provide the complete solution of the unitary $3$-body problem in an isotropic harmonic trap~\cite{WernerPRL} (for completeness, one also has to include the eigenstates which are common to the unitary and the non-interacting problem~\cite{WernerPRL}, mentioned at the end of subsection~\ref{subsubsec:ZRM}).
This was first realised for the ground state in~\cite{TanScaling}.
Remarkably, this $3$-body spectrum in a trap allows to compute the third virial coefficient of the homogeneous unitary gas~\cite{Hu}, whose value was confirmed experimentally (see~\cite{ChevyNature} and the contribution of F. Chevy and C. Salomon to this volume).
For spinless bosons, the unitary $3$-body problem in an isotropic harmonic trap has two families of eigenstates (apart from the aforementioned common eigenstates with the non-interacting problem)~\cite{Jonsell,WernerPRL}: the states corresponding to real solutions $s$ of the transcendental equation~(\ref{eq:trans_hypergeo}), which we call universal states; and the states corresponding to the imaginary solution for $s$, which we call efimovian. Eqs.~(\ref{eq:Eint},\ref{eq:F(R)}) apply to universal states. The efimovian states are discussed in the next subsection.
\subsubsection{Efimovian case}
In this subsection we consider the case $s^2<0$,
i.e. $s$ is purely imaginary.
In this case, {\rm all} solutions of the Schr\"{o}dinger-like equation (\ref{eq:schr_F}) are bounded and oscillate more and more rapidly when $R\to 0$.
In order to obtain a hermitian problem with a discrete spectrum, one has to impose the boundary condition~\cite{Danilov,theseFelix}:
\begin{equation}
\exists A/\ F(R)
\sim
A\ \mbox{Im}\, \left[\left(\frac{R}{R_t}\right)^{s}\right]
\ \ \ {\rm for}\ R\to 0
,
\label{eq:CL_Rt}
\end{equation}
where $R_t$ is an additional 3-body parameter.~An equivalent form is:
\begin{equation}
\exists A'/\ F(R)
\sim
A'\ \sin\left[|s|\ln\left(\frac{R}{R_t}\right)\right]
\ \ \ {\rm for}\ R\to 0
.
\label{eq:CL_sin}
\end{equation}
The corresponding hyperradial wavefunctions are
\begin{equation}
F(R) = R^{-1}\, W_{E/2,s/2}(R^2/a_{ho}^2)
\label{eq:Fefimovien}
\end{equation}
where $W$ is a Whittaker function, and the spectrum is given by the implicit equation
\begin{equation}
\mbox{arg}\, \Gamma\left[\frac{1+s-E/(\hbar\omega)}{2}\right] =
-|s| \ln (R_t/a_{ho})
+\mbox{arg}\, \Gamma(1+s) \ \mbox{mod}\ \pi
\label{eq:Eefimovien}
\end{equation}
obtained in~\cite{Jonsell}, whose solutions form a discrete series, which is unbounded from below, and can be labeled by a quantum number $q\in\mathbb{Z}$.
In free space ($\omega=0$), there is a geometric series of bound states
\begin{equation}
E_q=-\frac{2\hbar^2}{mR_t^{\phantom{.}2}} \exp\left({-q \frac{2\pi}{|s|}+\frac{2}{|s|}{\rm arg\,}\Gamma(1+s)}\right),\ q\in\mathbb{Z}
\label{eq:Eefimov}
\end{equation}
\begin{equation}
F(R)=K_s\left(R\sqrt{2m|E|/\hbar^2}\right)
\label{eq:Fefimov}
\end{equation}
where $K$ is a Bessel function.
For $3$ particles this corresponds to the well-known series of Efimov $3$-body bound states \cite{Efimov70,Efimov73}.
This also applies to the $4$-body bound states in the aforementioned case of $(3+1)$ fermions with $m_\uparrow/m_\downarrow$ between $\simeq 13.384$ and $13.607\ldots$~\cite{CMP}.
As expected,
in the limit $E\to -\infty$, the spectrum of the efimovian states in the trap (\ref{eq:Eefimovien}) approaches the free space spectrum (\ref{eq:Eefimov}).
The unboundedness of the spectrum in the zero-range limit is a natural consequence of the Thomas effect and of the limit cycle behavior~\cite{theseFelix}.
\subsection{Vanishing bulk viscosity}
\label{subsec:viscosity}
In this subsection, we give a simple rederivation of the fact that the bulk viscosity of the unitary gas in the normal phase is zero.
This result was obtained in~\cite{SonBulkVisco}
(see also \cite{NishidaSonChap}).
It helps analysing e.~g. the ongoing experimental studies of the shear viscosity, whose value is of fundamental importance~(\cite{SchaeferLong2010} and refs. therein).
Although the superfluid regime was also treated in~\cite{SonBulkVisco},
we omit it here for simplicity.
In our rederivation we shall use the scaling solution
and the existence of the undamped breathing mode.~\footnote{
In article~\cite{SonBulkVisco}, the vanishing of the bulk viscosity was deduced from the so-called general coordinate and conformal invariance,
the
scaling solution being
unknown to its author of at the time of writing
(although it had been obtained in~\cite{CRASCastin}).
The scaling solution was recently rederived using this
general coordinate and conformal invariance~\cite{NishidaSonChap}.
Several other results presented in subsections~\ref{subsec:so21}, \ref{subsec:separability} and \ref{subsec:pcots} were also rederived
using this field theoretical formalism (\cite{NishidaSonChap} and refs. therein).}
In the hydrodynamic theory for a normal compressible viscous fluid~\cite{LandauHydro,SonBulkVisco},
the (coarse-grained) evolution of the gas in a trapping potential $U(\mathbf{r},t)$ is described by the atom number density $\rho(\mathbf{r},t)$,
the velocity vector field $\mathbf{v}(\mathbf{r},t)$, and the entropy per particle (in units of $k_B$) $s(\mathbf{r},t)$.
These $5$ scalar functions solve $5$ equations
which are given for completeness in Appendix~4
although we will not directly use them here.
We will only need the equation for the increase of the total entropy $S=\int \rho s\, d^3r$ of the gas
\begin{equation}
\frac{dS}{dt}=\int \frac{\kappa \|\mathbf{\nabla} T\|^2}{T^2} d^3r
+
\int \frac{\eta}{2T}\sum_{ik}\left(
\frac{\partial v_i}{\partial x_k}+\frac{\partial v_k}{\partial x_i}-\frac{2}{3}\delta_{ik}\mathbf{\nabla}\cdot\mathbf{v} \right)^2 d^3r
+
\int \frac{\zeta}{T}\|\mathbf{\nabla}\cdot\mathbf{v}\|^2 d^3r
\label{eq:dSdt}
\end{equation}
which follows from the hydrodynamic equations~(\ref{eq:entr_prod},\ref{eq:cte});
note that the thermal conductivity $\kappa$, the shear viscosity $\eta$ and the bulk viscosity $\zeta$ have to be $\geq 0$ so that $dS/dt\geq0$~\cite{LandauHydro}.
The hydrodynamic theory is expected to become exact in the limit where the length (resp. time) scales on which the above functions vary are much larger than microscopic length (resp. time) scales such as $1/k_F$ (resp. $\hbar/E_F$).
We consider
the following gedankenexperiment:
Starting with the gas at thermal equilibrium in a trap of frequency $\omega$,
we suddenly switch the trapping frequency
at $t=0$ to a different value $\omega_+$.
As we have seen in subsection~\ref{subsec:tds}
and at the beginning of subsection~\ref{subsec:so21},
this excites an {\it undamped} breathing mode:
For $t>0$, the size of the gas oscillates {\it indefinitely}.
This rigorously periodic evolution of the system implies that
the total entropy $S(t)$ is periodic, and since it cannot decrease, it has to be constant.
Thus each of the terms in the right-hand-side of~(\ref{eq:dSdt}), and in particular the last term, has to vanish.
Thus
$\zeta(\mathbf{r},t) \|\mathbf{\nabla}\cdot\mathbf{v}(\mathbf{r},t)\|^2\equiv0$.
This implies that $\zeta$ is identically zero, as we now check.
From the scaling evolution~(\ref{eq:ansatz}) of each many-body eigenstate, one can deduce (using the quantum-mechanical expression for the particle flux) that
\begin{equation}
\mathbf{v}(\mathbf{r},t)=\frac{\dot{\lambda}}{\lambda}\mathbf{r},
\label{eq:v_echelle}
\end{equation}
so that $\mathbf{\nabla}\cdot\mathbf{v}=3\dot{\lambda}/\lambda$.
For $t$ approaching $0$ from above,
we have $\dot{\lambda}(t)\neq0$,
as is intuitively clear and can be checked from Eqs.~(\ref{eq:inco},\ref{eq:russe});
thus $\zeta(\mathbf{r},t)=0$ and by continuity $\zeta(\mathbf{r},t=0)=0$. Since the central density and temperature in the initial equilibrium state of the gas are arbitrary,
we conclude that $\zeta(\rho,T)=0$ for all $\rho$ and $T$.
An alternative derivation of this result is presented in Appendix~5.
\subsection{Short-range scaling laws}
\label{subsec:scaling_laws}
As opposed to the previous subsections, we now consider an arbitrary scattering length and an arbitrary external potential, possibly with periodic boundary conditions.
Of all the particles $1,\ldots,N$, let us consider a subset $J\subset\{1,\ldots,N\}$ containing $n_\uparrow$ particles of spin $\uparrow$ and $n_\downarrow$ particles of spin $\downarrow$.
From the particle positions $(\mathbf{r}_i)_{i\in J}$, we can define
a hyperradius $R_J$ and hyperangles $\mbox{\boldmath$\Omega$}_J$, and
a center of mass position $\mathbf{C}_J$.
The positions of all particles that do not belong to $J$ are denoted by $\mathbf{\mathcal{R}}_J=(\mathbf{r}_i)_{i\notin J}$.
In the absence of a $(n_\uparrow+n_\downarrow)$-body resonance
(see Appendix~6),
one expects that, for any eigenstate,
in the limit $R_J\to0$ where all particles belonging to the subset $J$ approach each other while $(\mbox{\boldmath$\Omega$}_J,\mathbf{C}_J,\mathbf{\mathcal{R}}_J)$ remain fixed, there exists a function $A_J$ such that
\begin{equation}
\psi(\mathbf{r}_1,\ldots,\mathbf{r}_N) = R_J^\nu\,\phi(\mbox{\boldmath$\Omega$}_J)
\,A_J(\mathbf{C}_j,\mathbf{\mathcal{R}}_J)
+o(R_J^\nu).
\label{eq:RJ->0}
\end{equation}
Here, $\nu=s_{\rm min}(n_\uparrow,n_\downarrow)-\frac{3(n_\uparrow+n_\downarrow)-5}{2}$ with $s_{\rm min}(n_\uparrow,n_\downarrow)$ the smallest possible value of $s$ for the problem of $n_\uparrow$ particles of spin $\uparrow$ and $n_\downarrow$ particles of spin $\downarrow$ ($s$ being defined in Sec.~\ref{subsec:separability})
and
$\phi(\mbox{\boldmath$\Omega$}_J)$ is the corresponding hyperangular wavefunction (also defined in Sec.~\ref{subsec:separability}).
This statement is essentially contained in~\cite{PetrovShlyapSalomon,TanScaling}. It comes from the intuition that, in the limit where the $n_\uparrow+n_\downarrow$ particles approach each other, the $N$-body wavefunction should be proprtional to the $(n_\uparrow+n_\downarrow)$-body zero-energy free space wavefunction Eq.(\ref{eq:psifreeR}).
This was used in~\cite{PetrovShlyapSalomon} to predict that
the formation rate of deeply bound molecules by three-body recombination, $\Gamma\equiv-\dot{N}/N$, behaves as $\hbar \Gamma/E_F\sim K\cdot(k_F b)^{2 s_{\rm min}(2,1)}$ in the low-density limit, with $b$ on the order of the van der Waals range
and $K$ a numerical prefactor which depends on short-range physics.
The analytical solution of the hyperangular three-body problem [Eq.(\ref{eq:trans})] yields $s_{\rm min}(2,1)=1.772724\ldots$ (this value is reached in the angular momentum $l=1$ channel).
Experimentally, this scaling has not been checked,
but the smallness of $\hbar\Gamma/E_F$ is one of the crucial ingredients which allow to realise the unitary gas.
\acknowledgement
We thank O. Goulko,
O. Juillet,
T.~Sch\"afer,
D.~T. Son,
B.~Svistunov
and S.~Tan
for helpful discussions while writing this manuscript,
and the authors of~\cite{Jin_p} for their data.
F.~W. is supported by NSF grant PHY-0653183, Y.~C. is member of IFRAF and acknowledges support from the ERC project FERLODIM N.228177.
\section*{Appendix 1: Effective range in a lattice model}
\addcontentsline{toc}{section}{Appendix 1}
\label{app:re}
To calculate the effective range $r_e$
[defined by Eq.(\ref{eq:lke})]
for the lattice model
of subsection \ref{subsec:lm},
it is convenient to perform in the expression~(\ref{eq:fklm}) of the scattering amplitude an analytic continuation
to purely imaginary incoming
wavevectors $k_0$, setting
$k_0=i q_0$ with $q_0$ real and positive.
Eliminating $1/g_0$ thanks to Eq.(\ref{eq:g0_gen}) we obtain the useful expression:
\begin{equation}
\label{eq:flu}
-\frac{1}{f_{k_0}} = \frac{1}{a} +4\pi
\int_{\mathcal{D}} \frac{d^3k}{(2\pi)^3}
\left[\frac{1}{q_0^2+2m\epsilon_{\mathbf{k}}/\hbar^2}-\frac{1}{2m\epsilon_{\mathbf{k}}/\hbar^2}\right].
\end{equation}
We first treat the case of the parabolic dispersion relation Eq.(\ref{eq:pdr}).
A direct expansion of Eq.(\ref{eq:flu}) in powers of $q_0$ leads to an infrared
divergence. The trick is to use the fact that the integral over $\mathcal{D}$
in Eq.(\ref{eq:flu}) can be written as the integral of the same integrand over the whole
space minus the integral over the supplementary space $\mathbb{R}^3\setminus \mathcal{D}$.
The integral over the whole space may be performed exactly using
\begin{equation}
\int_{\mathbb{R}^3}
\frac{d^3k}{(2\pi)^3} \left[\frac{1}{q_0^2+k^2}-\frac{1}{k^2}\right] = -\frac{q_0}{4\pi}.
\end{equation}
This leads to the transparent expression, where the term corresponding to $ik$ in Eq.(\ref{eq:u(k)}),
and which is non-analytic in the energy $E$, is now singled out:
\begin{equation}
-\frac{1}{f_{k_0}^{\rm parab}} = \frac{1}{a} - q_0 -4\pi \int_{\mathbb{R}^3\setminus\mathcal{D}} \frac{d^3k}{(2\pi)^3}
\left[\frac{1}{q_0^2+k^2}-\frac{1}{k^2}\right].
\end{equation}
This is now expandable in powers of $q_0^2$, leading to the effective range for the parabolic dispersion relation:
\begin{equation}
\label{eq:repa}
r_e^{\rm parab} = \frac{1}{\pi^2} \int_{\mathbb{R}^3\setminus\mathcal{D}} \frac{d^3k}{k^4}.
\end{equation}
We now turn back to the general case. The trick is to consider the difference between the inverse scattering amplitudes
of the general case and the parabolic case with a common value of the scattering length:
\begin{equation}
\frac{1}{f_{k_0}^{\rm parab}} - \frac{1}{f_{k_0}} =
4\pi \int_{\mathcal{D}} \frac{d^3k}{(2\pi)^3}
\left[\frac{1}{q_0^2+2m\epsilon_{\mathbf{k}}/\hbar^2}-\frac{1}{q_0^2+k^2}
-\frac{1}{2m\epsilon_{\mathbf{k}}/\hbar^2} + \frac{1}{k^2}\right].
\end{equation}
This is directly expandable to second order in $q_0$, leading to:
\begin{equation}
r_e - r_e^{\rm parab} = 8\pi \int_{\mathcal{D}} \frac{d^3k}{(2\pi)^3} \left[
\frac{1}{k^4} - \left(\frac{\hbar^2}{2m\epsilon_{\mathbf{k}}}\right)^2
\right].
\end{equation}
The numerical evaluation of this integral for the Hubbard dispersion relation
Eq.(\ref{eq:disphub}) leads to the Hubbard model effective range Eq.(\ref{eq:reh}).
Finally, we specialize the general formula to the parabolic plus quartic form
Eq.(\ref{eq:mixte}). Setting $\mathbf{k}=(\pi/b)\mathbf{q}$ and using Eq.(\ref{eq:repa}), we obtain
\begin{equation}
\frac{\pi^3 r_e^{\rm mix}}{b} = \int_{\mathbb{R}^3\setminus[-1,1]^3} \frac{d^3q}{q^4}
+ \int_{[-1,1]^3} \frac{d^3q}{q^4} \left[1-\frac{1}{(1-Cq^2)^2}\right].
\end{equation}
The trick is to split the cube $[-1,1]^3$ as the union of $B(0,1)$, the sphere of center $0$ and unit radius,
and of the set $X=[-1,1]^3\setminus B(0,1)$. One has also $\left(\mathbb{R}^3\setminus[-1,1]^3\right) \cup X=\mathbb{R}^3\setminus
B(0,1)$ so that
\begin{equation}
\label{eq:interm}
\frac{\pi^3 r_e^{\rm mix}}{b} = \int_{\mathbb{R}^3\setminus B(0,1)} \frac{d^3q}{q^4}
+\int_{B(0,1)} \frac{d^3q}{q^4} \left[1-\frac{1}{(1-Cq^2)^2}\right]
- \int_X \frac{d^3q}{q^4} \frac{1}{(1-Cq^2)^2}.
\end{equation}
One then moves to spherical coordinates of axis $z$.
The first two terms in the right hand side may be calculated exactly. In particular, one introduces
a primitive of $q^{-2}(1-C q^2)^{-2}$, given by $C^{1/2} \Phi(C^{1/2}q)$ with
\begin{equation}
\Phi(x) = \frac{x}{2(1-x^2)} + \frac{3}{2}\mbox{arctanh}\, x -\frac{1}{x}.
\end{equation}
In the last term of Eq.(\ref{eq:interm}) one integrates over the modulus $q$ of $\mathbf{q}$ for a fixed direction
$(\theta,\phi)$ where $\theta$ is the polar angle and $\phi$ the azimuthal angle.
One then finds that $q$ ranges from $1$ to some maximal value $Q(\theta,\phi)$, and the integral
over $q$ provides the difference $\Phi(C^{1/2}Q)-\Phi(C^{1/2})$. Remarkably, the term
$-\Phi(C^{1/2})$ cancels the contribution of the first two integrals in the right hand side
of Eq.(\ref{eq:interm}), so that
\begin{equation}
\frac{r_e^{\rm mix}}{b} = -\frac{C^{1/2}}{\pi^3} \int_0^{2\pi} d\phi \int_{-1}^{1} du\, \Phi[C^{1/2}Q(\theta,\phi)]
\end{equation}
where as usual we have set $u=\cos\theta$.
Using the symmetry under parity along each Cartesian axis, which adds a factor $8$, and restricting to the face
$q_x=1$ of the cube, which adds a factor $3$, the expression of $Q(\theta,\phi)$ is readily obtained,
leading to
\begin{equation}
\label{eq:remi}
\frac{r_e^{\rm mix}}{b} = -\frac{24 C^{1/2}}{\pi^3} \int_0^{\pi/4} d\phi \int_0^{\frac{\cos\phi}{\sqrt{1+\cos^2\phi}}}
du \, \Phi\left(\frac{C^{1/2}}{\cos\phi \sqrt{1-u^2}}\right).
\end{equation}
In the limit $C\to 0$, $r_e^{\rm mix}\to r_e^{\rm parab}$, and Eq.(\ref{eq:remi}) may be calculated analytically
with $\Phi(x)\sim -1/x$ and with an exchange of the order of integration: This leads to Eq.(\ref{eq:rep}).
For a general value of $C\in [0,1/3[$ we have calculated Eq.(\ref{eq:remi}) numerically,
and we have identified the magic value of $C$ leading to a zero effective range, see
Eq.(\ref{eq:magic}). With the same technique, we can calculate the value of $K$ appearing in Eq.(\ref{eq:g0_exp})
from the expression
\begin{equation}
K = \frac{12}{\pi C^{1/2}} \int_0^{\pi/4} d\phi \int_0^{\frac{\cos\phi}{\sqrt{1+\cos^2\phi}}}
du \, \mbox{arctanh}\, \frac{C^{1/2}}{\cos\phi \sqrt{1-u^2}}.
\end{equation}
\section*{Appendix 2: What is the domain of a Hamiltonian?}
\addcontentsline{toc}{section}{Appendix 2}
\label{app:domain}
Let us consider a Hamiltonian $H$ represented by a differential
operator also called $H$.
A naive and practical definition
of the domain $D(H)$ of $H$ is that it is the set
of wavefunctions over which the action of the Hamiltonian is
indeed represented by the considered differential operator.
In other words, if a wavefunction $\psi_{\rm bad}$
does not belong to $D(H)$, one should not calculate the action of $H$
on $\psi_{\rm bad}$ directly using the differential operator $H$.
If $H$ if self-adjoint, one should rather expand $\psi_{\rm bad}$
on the Hilbert basis of eigenstates of $H$ and calculate the
action of $H$ in this basis.
For example, for a single particle
in one dimension in a box with infinite walls in $x=0$
and $x=1$, so that $0\leq x \leq 1$, one has the Hamiltonian
\begin{equation}
\label{eq:dfa}
H = -\frac{1}{2} \frac{d^2}{dx^2},
\end{equation}
with the boundary conditions on the wavefunction
\begin{equation}
\label{eq:bca}
\psi(0)=\psi(1)=0
\end{equation}
representing the effect of the box.
To be in the domain, a wavefunction $\psi(x)$ should be twice
differentiable for $0<x<1$
and should obey the boundary conditions (\ref{eq:bca}).
An example of a wavefunction which is not in the domain is the constant
wavefunction $\psi(x)=1.$
An example of wavefunction in the domain is
\begin{equation}
\label{eq:ind}
\psi(x) = 30^{1/2} x (1-x).
\end{equation}
If one is not careful, one may obtain wrong results. Let us calculate
the mean energy and the second moment of the energy for $\psi$
given by (\ref{eq:ind}). By repeated action of $H$ onto $\psi$,
and calculation of elementary integrals, one obtains
\begin{eqnarray}
\label{eq:m1}
\langle H\rangle_\psi &=& 5 \\
\langle H^2\rangle_\psi &=& 0 ?!
\label{eq:m2}
\end{eqnarray}
Eq.(\ref{eq:m1}) is correct, but Eq.(\ref{eq:m2}) is wrong (it would lead
to a negative variance of the energy) because
$H\psi$ is not in $D(H)$ and the subsequent illicit
action of $H$ as the differential
operator (\ref{eq:dfa}) gives zero.
How to calculate the right value of $\langle H^2\rangle_\psi$~?
One introduces the orthonormal Hilbert basis of eigenstates of $H$,
\begin{equation}
\psi_n(x) = 2^{1/2} \sin [\pi(n+1)x], \ \ \
n\in \mathbb{N},
\end{equation}
with the eigenenergy $\epsilon_n=\frac{\pi^2}{2} (n+1)^2$.
Then $\psi$ of Eq.(\ref{eq:ind}) may be expanded as
$\sum_n c_n \psi_n(x)$, and the $k^{\rm th}$ moment of the energy may be
defined as
\begin{equation}
\langle H^k\rangle_\psi = \sum_{n\in \mathbb{N}} (\epsilon_n)^k
|c_n|^2.
\end{equation}
Since $c_n=4\sqrt{15}[1+(-1)^n]/[\pi(n+1)]^3$, one recovers
$\langle H\rangle_\psi =5$ and one obtains the correct value
$\langle H^2\rangle_\psi=30$, that leads to a positive energy
variance as it should be. Also $\langle H^k\rangle_\psi=+\infty$
for $k\geq 3$.
The trick of expanding $\psi$ in the eigenbasis of $H$ is thus quite powerful,
it allows to define the action of $H$ on any wavefunction $\psi$
in the Hilbert space (not belonging to the domain). It may be applied of
course only if $H$ is self-adjoint, as it is the case in our simple
example.
\section*{Appendix 3: Separability and Jacobi Coordinates for arbitrary masses}
\addcontentsline{toc}{section}{Appendix 3}
We here consider $N\geq 2$ harmonically trapped particles interacting in the unitary limit,
with possibly different masses $m_i$ but with the same isotropic angular
oscillation frequency
$\omega$. The Hamiltonian reads
\begin{equation}
H = \sum_{i=1}^{N}
\left[ -\frac{\hbar^2}{2m_i} \Delta_{\mathbf{r}_i} +
\frac{1}{2} m_i \omega^2 r_i^2
\right]
\label{eq:hamiljacob}
\end{equation}
and the unitary interaction is described by the Wigner-Bethe-Peierls contact conditions on the $N$-body wavefunction:
For all pairs of particles $(i,j)$, in the limit $r_{ij}=|\mathbf{r}_i-\mathbf{r}_j|\to 0$ with a {\sl fixed} value of the centroid of the particles
$i$ and $j$, $\mathbf{R}_{ij}\equiv (m_i\mathbf{r}_i+m_j \mathbf{r}_j)/(m_i+m_j)$,
that differs from the positions $\mathbf{r}_k$ of the other particles, $k\neq i,j$, there exists a function $A_{ij}$ such that
\begin{equation}
\psi(\mathbf{r}_1,\ldots,\mathbf{r}_N) = \frac{A_{ij}(\mathbf{R}_{ij};(\mathbf{r}_k)_{k\neq i,j})}{r_{ij}} + O(r_{ij}).
\end{equation}
As is well known and as we will explain below,
the internal Hamiltonian
$H_{\rm internal}= H - H_{\rm CM}$, where $H_{\rm CM}=
-\frac{\hbar^2}{2M}\Delta_\mathbf{C} + \frac{1}{2} M \omega^2 C^2$,
takes the form
\begin{equation}
H_{\rm internal} = \sum_{i=1}^{N-1}
\left[
-\frac{\hbar^2}{2\bar{m}} \Delta_{\mathbf{u}_i}
+ \frac{1}{2} \bar{m} \omega^2 u_i^2
\right]
\label{eq:Hjacodreduit}
\end{equation}
in suitably defined Jacobi coordinates [see Eqs.(\ref{eq:def_jac_y},\ref{eq:ui_vs_yi})].
Here $\mathbf{C}=\sum_{i=1}^{N} m_i\mathbf{r}_i/M$ is the center of mass position, $M=
\sum_{i=1}^{N} m_i$ is the total mass, and $\bar{m}$ is some arbitray
mass reference, for example the mean mass $M/N$.
Then it is straightforward to express Eq.(\ref{eq:Hjacodreduit})
in hyperspherical coordinates, the vector $(\mathbf{u}_1,\ldots,\mathbf{u}_{N-1})$ with $3N-3$ coordinates
being expressed in terms of its modulus $R$ and a set
of $3N-4$ hyperangles $\mbox{\boldmath$\Omega$}$, so that
\begin{equation}
\label{eq:hamilseparjacob}
H_{\rm internal} = -\frac{\hbar^2}{2\bar{m}}
\left[\partial_R^2 + \frac{3N-4}{R} \partial_R
+\frac{1}{R^2} \Delta_{\mbox{\boldmath$\Omega$}}
\right]
+\frac{1}{2} \bar{m} \omega^2 R^2
\end{equation}
where $\Delta_{\mbox{\boldmath$\Omega$}}$ is the Laplacian over the unit sphere of dimension $3N-4$.
As we shall see, the expression for the hyperradius is simply
\begin{equation}
R^2 \equiv \sum_{i=1}^{N-1} u_i^2 = \frac{1}{\bar{m}} \sum_{i=1}^{N} m_i (\mathbf{r}_i - \mathbf{C})^2.
\label{eq:Rjacob}
\end{equation}
This form of the Hamiltonian is then useful to show the separability of Schr\"odinger's equation
for the unitary gas in hyperspherical coordinates \cite{theseFelix,WernerPRA} for $N\geq 3$ and arbitrary masses.
The separability Eq.(\ref{eq:psisepargen}) that was described for simplicity in the case of equal mass particles in
subsection \ref{subsec:separability} indeed still holds in the case of different masses, if the Wigner-Bethe-Peierls model defines a self-adjoint Hamiltonian.
\footnote{Strictly speaking, it is sufficient that the Laplacian on the unit sphere
together with the Wigner-Bethe-Peierls boundary conditions reexpressed in terms of hyperangles
is self-adjoint, as extensively
used in \cite{CMP}. This is less restrictive than having the full Hamiltonian self-adjoint, since it allows for example
to have a $N$-body Efimov effect while the $N-1$ zero-range model is perfectly well-defined and does not experience any Efimov effect.}
We recall here the various arguments. First, for zero energy free space eigenstates, the form Eq.(\ref{eq:psifreeR}) is expected from scale invariance,
if the Hamiltonian is self-adjoint \cite{WernerPRA}. Second, the form Eq.(\ref{eq:psisepargen}) for the general case, including non-zero energy
and an isotropic harmonic trap, is expected because (i)
the Hamiltonian (\ref{eq:hamiljacob}), after separation of the center of mass,
has the separable form (\ref{eq:hamilseparjacob}) in hyperspherical coordinates, and (ii) Eq.(\ref{eq:psisepargen})
obeys the Wigner-Bethe-Peierls contact conditions if Eq.(\ref{eq:psifreeR}) does.
This point (ii) results from the fact that the Wigner-Bethe-Peierls conditions are imposed, for each pair of particles $(i,j)$,
for $r_{ij}\to 0$ with a {\sl fixed} value of $\mathbf{R}_{ij}$
that differs from the positions $\mathbf{r}_k$ of the other particles, $k\neq i,j$. Using $\mathbf{r}_i=\mathbf{R}_{ij}+[m_j/(m_i+m_j)] \mathbf{r}_{ij}$
and $\mathbf{r}_j=\mathbf{R}_{ij}-[m_i/(m_i+m_j)] \mathbf{r}_{ij}$, with $\mathbf{r}_{ij}\equiv \mathbf{r}_i-\mathbf{r}_j$, we indeed find that
\begin{equation}
\bar{m} R^2 = \frac{m_i m_j}{m_i+m_j} r_{ij}^2 + (m_i+m_j) (\mathbf{R}_{ij}-\mathbf{C})^2 + \sum_{k\neq i,j} m_k (\mathbf{r}_k-\mathbf{C})^2.
\end{equation}
For $N\geq 3$, we see that $\lim_{r_{ij}\to 0} R^2 > 0$, so that $R$ varies only to second order in $r_{ij}$
in that limit. Provided that the function $F(R)$ in Eq.(\ref{eq:psisepargen}) has no singularity at non-zero $R$,
the Wigner-Bethe-Peierls contact conditions are preserved [similarly to the argument Eq.(\ref{eq:chrcbp})].
Third, bosonic or fermionic exchange symmetries imposed on the $N$-body wavefunction cannot break the separability
in hyperspherical coordinates: Exchanging the positions of particles of same mass does not change the value of the hyperradius
$R$, it only affects the hyperangles and thus the eigenvalues $[(3N-5)/2]^2-s^2$ of the Laplacian on the unit sphere.
To derive the form Eq.(\ref{eq:Hjacodreduit}) of the internal Hamiltonian,
we introduce the usual Jacobi coordinates given for example in \cite{libro_italiano}:
\begin{equation}
\mathbf{y}_i \equiv \mathbf{r}_i -\frac{\sum_{j=i+1}^{N} m_j \mathbf{r}_j}{\sum_{j=i+1}^{N} m_j}
\ \ \ \mbox{for} \ 1\leq i \leq N-1.
\label{eq:def_jac_y}
\end{equation}
We note that $\mathbf{y}_i$ simply gives the relative coordinates
of particle $i$ with respect to the center of mass of the particles from
$i+1$ to $N$. To simplify notations, we also set $\mathbf{y}_N \equiv \mathbf{C}$.~\footnote{Alternatively, Eq.(\ref{eq:Hjacodreduit}) can be derived easily by recursion, see p.~63 of~\cite{theseFelix}.}
In compact form, the Jacobi change of variables corresponds to
setting
$\mathbf{y}_i = \sum_{j=1}^{N} M_{ij} \mathbf{r}_j$ for $1\leq i \leq N$,
where the non-symmetric matrix $M$ is such that:
\begin{itemize}
\item In the case $1\leq i<N$, one has: $M_{ij}=0$ for $1\leq j < i$,
$M_{ij}=1$ for $j=i$, and $M_{ij}=-m_j/(\sum_{k=i+1}^{N} m_k)$
for $i< j \leq N$.
\item $M_{Nj} = m_j/(\sum_{k=1}^{N} m_k)$ for $1 \leq j \leq N$.
\end{itemize}
From the formula giving the derivative of a composite function,
the kinetic energy operator writes
\begin{equation}
H_{\rm kin} \equiv \sum_{i=1}^{N} -\frac{\hbar^2}{2m_i} \Delta_{\mathbf{r}_i}
= -\frac{\hbar^2}{2} \sum_{j=1}^{N} \sum_{k=1}^{N} S_{jk} \mathrm{grad}_{\mathbf{y}_j}
\cdot \mathrm{grad}_{\mathbf{y}_{k}},
\end{equation}
where the {\sl symmetric} matrix $S$ is defined as
$S_{jk} = \sum_{i=1}^{N} M_{ji} M_{ki}/m_i.$
The explicit calculation of the matrix elements $S_{jk}$
is quite simple. Taking advantage of the fact that $S$ is symmetric,
one has to distinguish three cases,
(i) $1\leq j,k\leq N-1$, with $j=k$ and $j<k$ as subcases,
(ii) $j=k=N$, and (iii) $j<N, k=N$.
One then finds that $S$ is purely diagonal, with $S_{ii}=1/\mu_i$
for $1\leq i\leq N-1$ and $S_{NN}=1/M$. Here $\mu_i$ is the reduced
mass for the particle $i$ and for a fictitious particle of mass
equal to the sum of the masses of the particles from $i+1$ to $N$:
\begin{equation}
\frac{1}{\mu_i} = \frac{1}{m_i} + \frac{1}{\sum_{j=i+1}^{N} m_j} \ \ \
\mbox{for} \ \ 1\leq i\leq N-1.
\end{equation}
This results in the following form
\begin{equation}
H_{\rm kin} = -\frac{\hbar^2}{2M} \Delta_{\mathbf{C}} -
\sum_{i=1}^{N-1} \frac{\hbar^2}{2\mu_i} \Delta_{\mathbf{y}_i}.
\label{eq:kinjac}
\end{equation}
The next step is to consider the trapping potential energy term.
Inspired by Eq.(\ref{eq:kinjac}) one may consider the guess
\begin{equation}
H_{\rm trap} \equiv \sum_{i=1}^{N} \frac{1}{2} m_i \omega^2 r_i^2
\stackrel{?}{=} \frac{1}{2} M\omega^2 C^2 + \sum_{i=1}^{N-1}
\frac{1}{2} \mu_i \omega^2 y_i^2.
\label{eq:guessjacob}
\end{equation}
Replacing each $\mathbf{y}_i$ by their expression in the guess gives
\begin{equation}
M C^2 + \sum_{i=1}^{N-1} \mu_i y_i^2 = \sum_{j=1}^{N} \sum_{k=1}^{N}
Q_{jk} \mathbf{r}_j \cdot \mathbf{r}_k
\end{equation}
where $Q$ is uniquely defined once it is imposed to be a
{\sl symmetric} matrix. Setting $M_i=\sum_{j=i+1}^{N} m_j$
for $0\leq i \leq N-1$, and $M_N=0$, we find for the off-diagonal
matrix elements
\begin{equation}
Q_{jk} =
-\frac{\mu_{\min(j,k)}m_{\max(j,k)}}{M_{\min(j,k)}} + \frac{m_j m_k}{M}
+m_j m_k \sum_{i=1}^{\min(j,k)-1} \frac{\mu_i}{M_i^2}
\end{equation}
where $1\leq j,k\leq N$,
$\min(j,k)$ and $\max(j,k)$ respectively stand for the smallest
and for the largest of the two indices $j$ and $k$.
The key relation is then that
\begin{equation}
\frac{\mu_i}{M_i^2} = \frac{1}{M_i} - \frac{1}{m_i+M_i}
=\frac{1}{M_i} - \frac{1}{M_{i-1}}
\label{eq:trick1}
\end{equation}
since $M_i + m_i = M_{i-1}$ for $1\leq i\leq N$. This allows to calculate
the sum over $i$ of $\mu_i/M_i^2$, as all except the border terms
compensate by pairs. E.g. for $j < k$:
\begin{equation}
\sum_{i=1}^{j-1} \frac{\mu_i}{M_i^2}
= \frac{1}{M_{j-1}}- \frac{1}{M}
\label{eq:trick2}
\end{equation}
since $M_0=M$. One then finds that the off-diagonal elements
of the matrix $Q$ vanish. The diagonal elements of $Q$ may
be calculated using the same tricks (\ref{eq:trick1},\ref{eq:trick2}),
one finds $Q_{ii}= m_i$ for $1\leq i\leq N$.
As a consequence, the guess was correct and the question mark
can be removed from Eq.(\ref{eq:guessjacob}).
The last step to obtain Eq.(\ref{eq:Hjacodreduit}) is to appropriately
rescale the usual Jacobi coordinates, setting
\begin{equation}
\mathbf{u}_i \equiv (\mu_i/\bar{m})^{1/2} \mathbf{y}_i
\label{eq:ui_vs_yi}
\end{equation}
where $\bar{m}$ is an arbitrarily chosen mass.
A useful identity is
the expression
for the square of the hyperradius, Eq.(\ref{eq:Rjacob}). Starting from the definition
[first identity in Eq.(\ref{eq:Rjacob})] we see that $R^2=\sum_{i=1}^{N-1} \frac{\mu_i}{\bar{m}} y_i ^2$.
Then the second identity in Eq.(\ref{eq:Rjacob}) results from the fact that the guess
in Eq.(\ref{eq:guessjacob}) is correct.
\section*{Appendix 4: Hydrodynamic equations}\label{app:hydro}
\addcontentsline{toc}{section}{Appendix 4}
The hydrodynamic equations for a normal compressible viscous fluid are~(see~\cite{SonBulkVisco}\footnote{There is a typo in Eq.(10) of~\cite{SonBulkVisco}: $\mathbf{\nabla}_i(\rho v^i \partial_i s)$ should be replaced by $\mathbf{\nabla}_i(\rho v^i s)$.}, or \S15 and~\S49 in~\cite{LandauHydro}):
\begin{itemize}
\item
the continuity equation
\begin{equation}
\frac{\partial \rho}{\partial t}+\mathbf{\nabla}\cdot(\rho\mathbf{v}) =0,
\label{eq:cte}
\end{equation}
\item
the equation of motion
\begin{eqnarray}
m \rho \left(\frac{\partial v_i}{\partial t}+\mathbf{v}\cdot\mathbf{\nabla} v_i\right)&=&-\frac{\partial p}{\partial x_i}-\rho\frac{\partial U}{\partial x_i}+\sum_k\frac{\partial}{\partial x_k}\left[\eta\left(\frac{\partial v_i}{\partial x_k}+\frac{\partial v_k}{\partial x_i}-\frac{2}{3}\delta_{ik}\mathbf{\nabla}\cdot\mathbf{v}\right)\right]
\nonumber
\\ & &+\frac{\partial}{\partial x_i}\left(\zeta\, \mathbf{\nabla}\cdot\mathbf{v}\right)
\label{eq:motion}
\end{eqnarray}
where $m$ is the atomic mass, $\eta$ is the shear viscosity, $\zeta$ is the bulk viscosity, and the pressure $p(\mathbf{r},t)$
[as well as the temperature $T(\mathbf{r},t)$ appearing in the next equation]
is as always expressible in terms of $\rho(\mathbf{r},t)$ and $s(\mathbf{r},t)$ {\it via} the equation of state,\footnote{If we would neglect the position-dependence of $\eta$ and $\zeta$, (\ref{eq:motion}) would reduce to the Navier-Stokes equation.}
\item
the entropy-production equation
\begin{equation}
\rho T \left(\frac{\partial s}{\partial t}+\mathbf{v}\cdot\mathbf{\nabla}s\right)
= \mathbf{\nabla}\cdot(\kappa\mathbf{\nabla}T)+\frac{\eta}{2}\sum_{i,k}\left(
\frac{\partial v_i}{\partial x_k}+\frac{\partial v_k}{\partial x_i}-\frac{2}{3}\delta_{ik}\mathbf{\nabla}\cdot\mathbf{v} \right)^2
+\zeta\|\mathbf{\nabla}\cdot\mathbf{v}\|^2
\label{eq:entr_prod}
\end{equation}
where $\kappa$ is the thermal conductivity.
\end{itemize}
\section*{Appendix 5: Alternative derivation of the vanishing bulk viscosity}\label{app:hydro_alternatif}
\addcontentsline{toc}{section}{Appendix 5}
Consider the particular case of a unitary gas initially prepared at thermal
equilibrium in an isotropic harmonic trap at a temperature $T$ above
the critical temperature. When the harmonic trap becomes time dependent,
$U(\mathbf{r},t)=\frac{1}{2} m \omega^2(t) r^2$,
each many-body eigenstate of the statistical mixture evolves under
the combination Eq.(\ref{eq:ansatz}) of a time dependent gauge transform and a
time dependent scaling transform of scaling factor $\lambda(t)$.
The effect of the gauge transform is to shift the momentum
operator $\mathbf{p}_i$ of each particle $i$ by the spatially slowly varying
operator $m\mathbf{r}_i \dot{\lambda}/\lambda$. In the hydrodynamic framework,
this is fully included by the velocity field Eq.(\ref{eq:v_echelle}).
\footnote{To formalize this statement, we consider
a small but still macroscopic element of the equilibrium
gas of volume $dV$ around
point $\bar{\mathbf{r}}$, with $k_F^{-1} \ll dV^{1/3} \ll R$ where
$k_F$ is the Fermi momentum and $R$ the Thomas-Fermi radius of the gas.
We can define the density operator $\hat{\rho}_{\rm elem}$ of this element
by taking the trace of the full $N$-body density operator over
the spatial modes outside the element. Since the gauge transform
in Eq.(\ref{eq:ansatz}) is local in position space, $\hat{\rho}_{\rm elem}$
experiences the same unitary gauge transform. It would be tempting
to conclude from the general formula $dS = -k_B \mbox{Tr}[\hat{\rho}_{\rm elem}
\ln \hat{\rho}_{\rm elem}]$ that the entropy $dS$ of the element
is not changed by
the gauge transform. This is a valid conclusion however only if
the gauge transform does not bring $\hat{\rho}_{\rm elem}$ too far
from local thermal equilibrium.
To check this, we split the gauge transform for a single particle
of position $\mathbf{r}$ as $m r^2 \dot{\lambda}/(2 \hbar \lambda) =
m \dot{\lambda}/(2 \hbar \lambda) [\bar{r}^2 + 2 \bar{\mathbf{r}}\cdot (\mathbf{r}-
\bar{\mathbf{r}}) + (\mathbf{r}-\bar{\mathbf{r}})^2].$ The first term is an innocuous
uniform phase shift. The second term performs a uniform shift
in momentum space by the announced value $m\mathbf{v}(\bar{\mathbf{r}},t)$.
Due to Galilean invariance, this has no effect on the thermodynamic
quantities of the small element, such as its temperature, its pressure,
its density, its entropy.
With the estimate $\dot{\lambda}/\lambda \sim \omega$, $\bar{r}\sim
R$, $m\omega R \sim \hbar k_F$,
this second term is of order $k_F dV^{1/3}\gg 1$, not negligible.
The third term is of order $m\omega dV^{2/3}/\hbar
\sim N^{-1/3} k_F^2 dV^{2/3}$, negligible in the thermodynamic limit.}
Using the macroscopic consequences of a spatial scaling
Eqs.(\ref{eq:resca_macro_dens},\ref{eq:resca_macro_temp},\ref{eq:resca_macro_entropie_par_particule},\ref{eq:resca_macro_pression}),
one {\sl a priori} obtains a time dependent solution of the
hydrodynamic equations:
\begin{eqnarray}
T(\mathbf{r},t) &=& T(t=0)/\lambda^2(t) \\
\rho(\mathbf{r},t) &=& \rho(\mathbf{r}/\lambda,0)/\lambda^3(t) \\
s(\mathbf{r},t) &=& s(\mathbf{r}/\lambda,0) \\
p(\mathbf{r},t) &=& p(\mathbf{r}/\lambda,0)/\lambda^5(t) \\
v_i(\mathbf{r},t) &=& x_i \dot{\lambda}(t)/\lambda(t).
\end{eqnarray}
One then may {\sl a posteriori} check that Eq.(\ref{eq:cte}) is inconditionally satisfied, and that Eq.(\ref{eq:entr_prod}) is satisfied if $\zeta\equiv 0$.
Setting $\zeta\equiv 0$ in Eq.(\ref{eq:motion}), and using the hydrostatic condition $\mathbf{\nabla}p=-\rho\mathbf{\nabla} U$
at time $t=0$, one finds that Eq.(\ref{eq:motion}) holds provided that $\lambda(t)$ solves
Eq.(\ref{eq:russe}) as it should be.
\section*{Appendix 6: $n$-body resonances}
\addcontentsline{toc}{section}{Appendix 6}
\label{app:res_Ncorps}
Usually in quantum mechanics one takes the boundary condition that the wavefunction is bounded when two particles approach each other; in contrast, the Wigner-Bethe-Peierls boundary condition (\ref{eq:bpN}) expresses the existence of a $2$-body resonance.
If the interaction potential is fine-tuned not only to be close to a two-body resonance (i.e. to have $|a|\gg b$) but also to be close to a $n$-body resonance (meaning that a real or virtual $n$-body bound state consisting of $n_\uparrow$ particles of spin $\uparrow$ and $n_\downarrow$ particles of spin $\downarrow$ is close to threshold),
then one similarly expects that, in the zero-range limit, the interaction potential can be replaced by the
Wigner-Bethe-Peierls boundary condition,
{\it together with an additional boundary condition in the limit where any subset of $n_\uparrow$ particles of spin $\uparrow$ and $n_\downarrow$ particles of spin $\downarrow$ particles approach each other}.
Using the notations of Section~\ref{subsec:scaling_laws},
this additional boundary condition reads~\cite{PetrovBosons, WernerPRA,NishidaSonTan,theseFelix}:
\begin{equation}
\psi(\mathbf{r}_1,\ldots,\mathbf{r}_N) = \left(R_J^{-s}-\frac{\epsilon}{l^{2s}} R_J^s\right)\,R_J^{-\frac{3 n-5}{2}}\,\phi(\mbox{\boldmath$\Omega$}_J)
\,A_J(\mathbf{C}_j,\mathbf{\mathcal{R}}_J)
+o(R_J^\nu)
\label{eq:NbodyRes}
\end{equation}
where $s=s_{\rm min}(n_\uparrow,n_\downarrow)$,
while $l>0$ and $\epsilon=\pm1$
are parameters of the model playing a role analogous to the absolute value and the sign of the two-body scattering length.
This approach is only possible if the wavefunction remains square integrable,
i.~e. if $0\le s<1$,
which we assume in what follows.
This condition is satisfied e.g. for $n_\uparrow=2, n_\downarrow=1$ for a mass ratio $m_\uparrow/m_\downarrow \in ]8.62\ldots;13.6\ldots]$~\cite{Efimov73}.
Moreover we are assuming for simplicity that $s\neq0$.
Let us now consider the particular case where the two-body scattering length is infinite, and the external potential is either harmonic isotropic, or absent.
Then the
separability in internal hyperspherical coordinates of Section~\ref{subsec:separability} still holds for $n=N$. Indeed, Eq.(\ref{eq:NbodyRes}) then translates into the boundary condition on the hyperradial wavefunction
\begin{equation}
\exists A\in\mathbb{R}/\ F(R)\underset{R\to 0}{=} A \cdot \left( R^{-s} - \frac{\epsilon}{l^{2s}} R^s \right) + O\left(R^{s+2}\right)
\end{equation}
and does not affect the hyperangular problem.
Consequently~\cite{theseFelix},
\begin{itemize}
\item
For the $n$-body bound state, which exists if $\epsilon=+1$:
\begin{equation}
E=-\frac{2\hbar^2}{m\,l^2} \left[\frac{\Gamma(1+s)}{\Gamma(1-s)}\right]^{\frac{1}{s}},\end{equation}
\begin{equation}
F(R)=K_s\left(R \sqrt{-2 E \frac{m}{\hbar^2}}\right).
\end{equation}
\item
For the eigenstates in a trap:
\begin{equation} E {\rm\ solves:\ \ }
-\epsilon\cdot\left(\frac{\hbar}{m\omega\,l^2}\right)^s =
\frac{\Gamma\left(
\frac{1+s-E/(\hbar\omega)}{2}
\right) \Gamma(-s)}
{\Gamma\left(
\frac{1-s-E/(\hbar\omega)}{2}
\right) \Gamma(s)},
\end{equation}
\begin{equation}
F(R)=\frac{1}{R}\,W_{\frac{E}{2\hbar\omega},\frac{s}{2}}\left(R^2\frac{m\omega}{\hbar}\right).
\end{equation}
\end{itemize}
In particular, for $l=\infty$,
we are exactly at the $n$-body resonance, since
the energy of the $n$-body bound state vanishes. The spectrum in a trap then is $E=(-s+1+2 q)\hbar\omega$ with $q\in\mathbb{N}$.
Note that, most often, $s\geq1$, in which case one would have to use an approach similar to the one developped by Pricoupenko for the case of $2$-body resonances in non-zero angular momentum channels, and to introduce a modified scalar product~\cite{LudoPRL_ondeP,Ludo_onde_L}.
\input references.tex
\end{document}
|
\section{Introduction}
Einstein's general theory of relativity predicts the existence of
gravitational waves (GWs), oscillations in the space-time metric that
propagate at the speed of light. The Laser Interferometer
Gravitational-Wave Observatory, LIGO, is designed to detect and study
astrophysical GWs, with the promise of studying qualitatively new
physics and astrophysics.~\citep{Abramovici:1992p2546} In particular,
the direct detection of GWs will provide information about systems in
which strong-field gravitation dominates, a virtually untested regime
in which space-time curvature self-interacts. Such GW sources include
compact binary coalescences in which a neutron star or black hole
binary system inspirals together, coalescing to form a black hole; the
stellar core collapse thought to power Type II supernovae; rapidly
rotating asymmetric neutron stars; and possibly cosmic-scale processes that
produce a stochastic background of GWs.~\citep{Cutler:2002p7832}
In the past few years, the first generation of long-baseline
gravitational wave detectors has successfully operated at or near
design sensitivity. In collaboration with the Virgo 3~km and the GEO
600~m interferometers,~\citep{Acernese:2008p9535, Willke:2002p3770}
LIGO anchors a worldwide network of instruments in search of the first
direct detection of gravitational waves. The LIGO detectors operated
from November 2005 to October 2007, with joint data taking with Virgo
starting in May 2007. The data is currently being analyzed for GW
signals from inspiraling binary systems, burst sources, a stochastic
GW background, and rapidly rotating neutron stars. The status of
these searches and their astrophysical importance is discussed by
other authors in these proceedings.
This article focuses on the next generation of LIGO interferometers,
in particular Advanced LIGO, currently being designed and assembled at
two sites in the United States. The Livingston Parish, Louisiana
observatory will operate a single interferometer, L1, with 4~km long
arms while the Hanford, Washington observatory will operate two 4~km
interferometers within a common vacuum envelope, H1 and H2. The
second generation Advanced LIGO detectors will improve the sensitivity
of ground-based gravitational wave detectors by an order of magnitude
over current detectors. A preliminary Advanced LIGO design was
described in Ref.~\citep{Fritschel:2003p9356}, here we provide an
overview of the final design as construction begins. We first
describe the Advanced LIGO optical configuration, then follow with a
description of the dominant noise terms and anticipated sensitivities.
Finally, we conclude with comments on the initial tests of Advanced
LIGO and progress towards the first science runs.
\section{Advanced LIGO Optical Design}
\label{sec:advanced-ligo-design}
From a detector perspective, gravitational waves can be thought of as
a quadrupole strain of space, $h = \delta L / L$, which can be probed
by monitoring the relative positions of inertial test masses with
light. Equivalently, in a fixed Lorentzian frame, gravitational waves
create a tidal force, a force proportional to the distance from a
chosen origin. As with electromagnetic waves, GWs are transverse
waves that travel at the speed of light. Unlike electromagnetism, GWs
are constrained by mass and momentum conservation to be quadrupolar:
the strain (or tidal forces) contracts along one transverse dimension
while expanding the orthogonal dimension. Also unlike
electromagnetism, GWs are very weak and interact very weakly as they
propagate through space; detectable GWs are generated only by the
coherent acceleration of stellar masses at relativistic velocities. The
strongest nearby sources will produce strains on Earth no larger than
$h \approx 10^{-21}$. Finally, it's worth noting that GW detectors
measure the amplitude of a GW (as opposed to the power) so that the
observed volume of space scales cubicly with the detector sensitivity.
The Advanced LIGO optical design consists of a Michelson
interferometer with Fabry-Perot arm cavities, a power recycling cavity
and a signal recycling cavity as shown in Fig.~\ref{fig:layout}. The
Michelson topology is well matched to the quadrupole strain: a
properly oriented, linearly polarized GW propagating normal to the
interferometer plane generates a positive strain along one arm, a
negative strain along the other and vice versa, oscillating in time.
In the Advanced LIGO configuration, the arm lengths are controlled so
that Michelson interferometer reflects the input laser beam back
towards the laser while the anti-symmetric port is dark. The
differential motions of the two arms -- the GW signal --
constructively interfere at the beam splitter's Anti-Symmetric port,
labeled ``AS'' in Fig.~\ref{fig:layout}. The common mode signals
generated by common mode motion of the end mirrors, by laser frequency
noise, and by laser intensity noise constructively interfere at the
beam splitter's Symmetric port; to first order, this configuration
eliminates laser technical noise couplings to the GW signals.
Fabry-Perot cavities in the interferometer arms defined by the
partially reflecting Input Test Mass (ITM) and high reflectance End
Test Mass (ETM) resonate the input laser light to increase the power
in the arms. Similarly, the partially reflecting Power Recycling
Mirror (PRM) resonates the light that returns toward the laser from
the beam splitter's symmetric port. Together the arm and power
recycling cavities build up the laser power in the arms by a factor of
$\simeq 6000$. The power recycled Fabry-Perot Michelson topology is
identical to Initial LIGO and has been described in detail in
Ref.~\citep{Abbott:2009p9436}.
\begin{figure}[tb]
\centering
\includegraphics[height=4in]{IFO_diagram_modified.pdf}
\caption{The Advanced LIGO optical layout. The triangular suspended
input mode cleaner filters frequency and amplitude noise from the
laser (not shown) and provides a stable input beam. The Faraday
Isolator (FI) isolates the laser from the interferometer reflected
beam (REFL) used to control the laser frequency. The 4~km long
Fabry-Perot arm cavities are formed between the ITM and ETM test
masses. The Power Recycling Mirror (PRM) and Signal Recycling
Mirrors (SRM) form folded cavities discussed in the text. The GW
signals are carried by the light transmitted through the Output
Mode Cleaner at the Anti-Symmetric (AS) port.}
\label{fig:layout}
\end{figure}
Advanced LIGO has several significant changes in the optics relative
to Initial LIGO: the Signal Recycling Mirror (SRM) is added to the AS
port of the interferometer to form a signal recycling cavity, the
power and signal recycling cavities have a stable geometry, the GW
signal is detected using DC readout, and the laser power is increased.
The SRM forms a resonant cavity for the differential mode signal,
altering the interferometer
dynamics.~\citep{Meers:1988p8021,Mizuno:1993p10166} The impact on the
interferometer quantum noise is discussed further in the
\S\ref{sec:quantum} below.
In Initial LIGO, the $\approx 10$~m long power recycling cavity was
formed by mirrors having $\ge 10$~km radii of curvature, effectively a
flat-flat resonator geometry. That configuration was extremely
sensitive to changes in the curvatures caused by unavoidable thermal
lensing. To reduce the thermal sensitivity, the Advanced LIGO signal
and power recycling cavities are each formed by a folded chain of
three curved mirrors. In effect, the Advanced LIGO recycling cavities
incorporate beam expanding telescopes that reduce the sensitivity to
thermal lenses in the ITMs. In a similar change, the Fabry-Perot arm
cavities have a near concentric configuration motivated by the reduced
coating thermal noise from large spot sizes, discussed in
\S\ref{sec:therm}, and from the improved response to optical torques
for near concentric cavities, discussed in
Ref.~\citep{Sidles:2006p4148}.
The Advanced LIGO differential arm length will be detected in a
homodyne scheme known as DC readout. In DC readout, the GW signals are
measured directly as amplitude modulations on a static field at the AS
port. The static field is created by a small offset in the
differential arm length; GWs make oscillations around the offset,
modulating the output. However, many fields are present at the AS
port that don't carry GW information such as auxiliary control fields
and scattered non-resonant light. Advanced LIGO incorporates an
Output Mode Cleaner (OMC) at the AS port to select only those fields
containing GW signal. The OMC is a $\sim1~$m long optical cavity
which filters the interferometer output before detection, transmitting
only light from the arm cavity.
Finally, the Initial LIGO laser will be upgraded from a 10~W Master
Oscillator/Power Amplifier (MOPA) to a 180~W MOPA for Advanced
LIGO.~\citep{Willke:2008p10173} The input optics are upgraded to match
the laser: high power versions of electro-optic phase modulators,
photodetectors, and Faraday isolators replace conventional
components. With these changes, the maximum Advanced LIGO arm power
approaches 800~kW, improving the shot noise limited sensitivity by a
factor of $\approx 6$ with respect to Initial LIGO.
\section{Advanced LIGO Noise Contributions}
\label{sec:noiselimits}
The Advanced LIGO sensitivity limits are estimated from calculations of
technical and fundamental noises; many of these have been studied with
Initial LIGO and other dedicated experiments.\footnote{See
Ref.~\citep{Abbott:2009p9436} and references therein.} Below 10~Hz,
the sensitivity is limited by the seismic motion of the earth, at
intermediate frequencies thermal noise dominates, and at the highest
frequencies photon shot noise limits the sensitivity. The
interferometer noise contributions, modeled with the GWINC-v2 software
package and plotted in Fig.~\ref{fig:noise_budget}, are described in
the following sections.
\begin{figure}[tb]
\centering
\includegraphics[height=4in]{NoiseTermsPlot-crop.pdf}
\caption{The modeled noise budget for an Advanced LIGO
interferometer with $\phi_{SRM}= 0$ and 125~W input
power. The total noise (grey) is the incoherent sum of each of
the listed noise terms, described in detail below.}
\label{fig:noise_budget}
\end{figure}
\subsection{Acoustic and Seismic Isolation}
\label{sec:isolation}
The Advanced LIGO detector requires a residual RMS differential arm
motion of $\delta x \leq 10^{-15}$~m to maintain the arm power buildup
and to minimize the coupling of laser noise into the GW signal. In
the GW detection band, the required test mass displacement is $\delta
x \leq 10^{-19}~m/\sqrt{Hz}$ at 10 Hz and $\delta x \leq 2 \times
10^{-20}~m/\sqrt{Hz}$ at 100 Hz. Here, $\delta x$ refers only to the
differential arm motion; the common arm motion and the motion of the
other length degrees of freedom may be somewhat larger. To meet these
requirements, the interferometer is be isolated from environmental
influences such as acoustic noise, gas produced phase noise, and
seismic noise. Thus, the interferometer optics are enclosed in an
ultra-high vacuum system. The facility specifications have been
determined by the standard quantum sensitivity limit for a future
interferometer with 1 ton test masses. The $\simeq 10^{-9}$~torr
vacuum reduces the gas produced phase noise to an equivalent strain of
$h \approx 10^{-25}$, well below the anticipated Advanced LIGO
sensitivity. In addition, the interferometer detection beam paths,
including the photodetectors, are enclosed within the vacuum on the
seismic isolation platforms to eliminate acoustic coupling and reduce
the motion of light scatterers.
Isolating the interferometer optics from ground motion is a task
divided into several stages, with each stage providing isolation for
the following stage. The effect of each stage of isolation can be
seen in Fig.~\ref{fig:seismic}. At the lowest frequencies, an active,
6 degree of freedom, hydraulic, external pre-isolator (HEPI) reduces
the motion between 0.1 and 5~Hz by a factor of $\simeq
10$.~\citep{Fritschel:2004p9420} A two stage, active, internal seismic
isolation (ISI) system enclosed within the vacuum reduces ground
motion by a further factor of $\simeq 300$ at 1~Hz and $\simeq 3000$
at 10~Hz.~\citep{Abbott:2002p4431} Both active isolation systems consist
of a spring mounted, actuated platform outfitted with a suite of
motion sensors. The sensors measure the platform motion and feedback
to the actuators, thereby stabilizing the platform to a level limited by the
sensor noise floors and mechanical cross coupling.
\begin{figure}[tb]
\centering
\includegraphics[height=3in]{SeismicSpectra-crop.pdf}
\includegraphics[height=3in]{SuspensionSketch.pdf}
\caption{The Advanced LIGO seismic isolation consists of several
stacked stages of active isolation shown schematically on the
right, with colors matching the curve and the test mass shaded
gray. The reference ground motion (blue) is the average ground
displacement at the Livingston observatory. The HEPI curve (green)
is the motion atop the hydraulic pre-isolator located outside the
vacuum system. The anticipated ISI payload motion (red) is
calculated from models of the two stage active isolator
performance with the HEPI motion as the input spectrum. The
motion of the test mass at the end of the four-stage suspension is
calculated from a model and shown in the Quad curve (black).}
\label{fig:seismic}
\end{figure}
The interferometer optics are suspended from the ISI platforms using a
coupled pendulum system based on the GEO-600 three-stage
suspensions.~\citep{Plissi:2000p9494} The beam splitter, power
recycling cavity and signal recycling cavity optics are hung from
three-stage suspensions consisting of two metal masses linked with
steel wire, followed by the optic itself. Vertical isolation is
provided by cantilevered blade springs mounted to the metal masses.
Above the pendulum resonances, the triple suspension provides
isolation proportional to $1/f^6$. The ITMs and ETMs are suspended
from four-stages consisting of two metal masses with vertical springs,
a fused silica intermediate mass, and the test mass. Welded fused
silica fibers join the test mass to the fused silica mass
above.~\citep{Robertson:2002p9490} The additional isolation stage is
necessary to meet the displacement noise goals at 10~Hz, and the fused
silica fibers reduce the thermal noise as discussed below.
Altogether, the Advanced LIGO isolation systems reduce the
seismic-induced test mass motion by 10 orders of magnitude to $\delta
x \simeq 10^{-20}~m/\sqrt{Hz}$ at 10 Hz, opening the frequency band
from 10 to 40~Hz for gravitational wave searches.
\subsection{Thermal Noise Sources}
\label{sec:therm}
Between 10 and 200 Hz, thermal noise sources limit the interferometer
sensitivity. An unavoidable consequence of energy dissipation, thermal
noise is modeled by applying the fluctuation-dissipation theorem to
all aspects of the system which influence the motion or measurement of
the arm cavity test masses. The fluctuation dissipation theorem is
closely related to Brownian motion, hence thermal noise associated
with mechanical motion is often called Brownian noise. Three sources
of dissipation dominate the Advanced LIGO thermal noise: mechanical
loss in each test mass leads to fluctuations in the mirrored surface
of the test mass; dissipation within the suspension fibers generates a
fluctuating force on the test masses; and losses within the mirror's
dielectric coating generate a fluctuating phase shift of the reflected
light.
The test mass mechanical loss determines the level of substrate
Brownian noise that causes fluctuations of the surface with respect to
the center of mass. Early Advanced LIGO designs considered sapphire
test masses for their superior mechanical and thermal
properties. ~\citep{Rowan:2000p10031} Since then, fused silica test
masses have demonstrated sufficiently low loss and have been adopted
as the substrate material. The Advanced LIGO test masses are 40~kg,
high purity, low-inclusion fused silica cylinders 34~cm in diameter.
To maintain the substrates' excellent mechanical properties, no lossy
materials contact the test mass (eg. magnets). The masses are
suspended via fused silica mounting blocks hydroxy-catalysis bonded to
each side.~\citep{Rowan:1998p10037} Fused silica fibers are welded to
the blocks and connect to the upper fused silica mass in a similar
fashion. The test masses are actuated using a non-contact, low-force
and low-noise electrostatic drive. As a result of these measures, the
test masses have very low loss (mechanical quality factors
exceeding $10^7$), and correspondingly low fluctuations, contributing
to the strain noise at $h \approx 3 \times 10^{-24} f^{-1/2}\;
Hz^{-1/2}$.
Loss in the mechanical structure supporting the test masses generates
fluctuating forces at the masses' suspension
points.~\citep{Gonzalez:2000p3745} The extremely low mechanical loss
needed to limit the fluctuations motivates the fused silica fiber
stage of the four-stage suspension. Since the loss is dominated by
the bending regions near the fiber ends, the cylindrical fibers are
laser polished and drawn from fused silica rod with a carefully
controlled, variable diameter. At the ends where the fiber is welded
to the test mass, the fibers have a large diameter to reduce flexing
of the (potentially higher loss) welded joints. The fibers then taper
with an optimized profile so that the bending occurs predominantly in
a low loss region of the fiber. Suspension thermal noise contributes
to the detector noise below 30~Hz, limiting the low-frequency
sensitivity to $h \approx 2\times 10^{-21}f^{-2}\; Hz^{-1/2}$.
Finally, thermal noise from the dielectric mirror coatings limits the
detector noise between 40 and 140~Hz, the most sensitive region. The
coatings are alternating layers of SiO$_2$ and titanium-doped
TaO$_2$.\citep{Harry:2007p9489} Both thermo-optic and mechanical loss
contribute to the thermal noise. The thermo-optic noises include
thermo-refractive fluctuations in the layers' index of refraction as
well as the thermo-elastic fluctuations that modify the layer
thickness and hence the magnitude and phase of the reflected field.
The coating mechanical loss dominates the thermal noise by an order of
magnitude, primarily in the thick ETM high reflector. Because the
thin film coatings have a high mechanical loss relative to the
substrate, the coating Brownian noise exceeds that of the substrate by
nearly an order of magnitude. The coating noise is inversely
proportional to the beam diameter, motivating large spot sizes on the
optics as described in \S\ref{sec:advanced-ligo-design}. Coating
thermal noise limits the detector sensitivity to $h\approx 2.5 \times
10^{-23} f^{-1/2}\; Hz^{-1/2}$.
\subsection{Quantum Noise}
\label{sec:quantum}
Quantum mechanics limits the precision at which the test mass
positions can be determined. At high frequencies, photon shot noise
limits the sensitivity to $h \propto \sqrt{ f / P}$, while at low
frequencies radiation pressure limits the sensitivity to $h \propto
\sqrt{P}/f^2$. The Advanced LIGO interferometer is a realization of a
Heisenberg microscope: the high laser power required to determine the
position of the test masses exerts a fluctuating radiation pressure
which perturbs the test mass positions. In the absence of
position-momentum correlations, the Advanced LIGO strain sensitivity
is limited by the Standard Quantum Limit (SQL) $h_{SQL} = 1.8 \times
10^{-22}/f\; Hz^{-1/2}$. Because the signal recycling cavity couples
the test mass position and momentum, sub-SQL sensitivity is possible
over a frequency bandwidth $\Delta f \sim f$ at the expense of the
sensitivity at other wavelengths.~\citep{Buonanno:2001p2} The primary
difference between the three curves shown in
Fig.~\ref{fig:noise_curves} is the degree of correlation between the
test mass position and momentum as determined by the SRM reflectivity
and the signal recycling cavity length tuning.\footnote{These noise
curves can be found at
https://dcc.ligo.org/cgi-bin/private/DocDB/ShowDocument?docid=2974
and the documents therein.} The noise curves include the thermal
and seismic noises calculated by GWINC-v2.
\begin{figure}[tb]
\centering
\includegraphics[height=3.5in]{AnticipatedSensitivity-crop.pdf}
\caption{Amplitude spectral densities of the anticipated
sensitivities of the Advanced LIGO interferometers as a function
of the tuning of the signal recycling phase, $\phi_{SRM}$. \textbf{No
SRM} is a potential initial interferometer configuration with no
signal recycling mirror with modest sensitivity; the
\textbf{Broadband} configuration has good sensitivity at all
frequencies; and \textbf{NS/NS} is optimized for the detection of
two coalescing 1.4~M$_\odot$ neutron stars. }
\label{fig:noise_curves}
\end{figure}
\section{Advanced LIGO Progress}
\label{sec:advanc-ligo-progr}
Some Advanced LIGO hardware has already been installed on the Initial LIGO
interferometers. These components include: a 35~W laser
Master-Oscillator/Power-Amplifier; a high-power, in vacuum Faraday
Isolator; a single stage, in-vacuum seismic isolation system; and DC
readout using an in-vacuum Output Mode Cleaner. With these systems,
LIGO has begun another science run, the sixth, with significantly
improved high frequency performance.
The Advanced LIGO project began in 2008, with plans for the first
in-vacuum hardware installation in early 2011. To evaluate the
greatly increased chances for direct GW detection, we consider compact
binary coalescences for which the source rate can be estimated from
observed binary pulsar systems. Once the instruments reach the
anticipated sensitivities, we can expect to detect between 1 and 1,000
compact binary coalescences per year. As installation and
commissioning progress, Advanced LIGO will transform the field from
searching for the first direct GW detection to exploring the rich
phenomena of GW astrophysics.
\section*{Acknowledgments}
We gratefully acknowledge the support of the United States National
Science Foundation for the construction and operation of the LIGO
Laboratory and the Science and Technology Facilities Council of the
United Kingdom, the Max-Planck-Society, and the State of
Niedersachsen/ Germany for support of the construction and operation
of the GEO600 detector. We also gratefully acknowledge the support of
the research by these agencies and by the Australian Research Council,
the Council of Scientific and Industrial Research of India, the
Istituto Nazionale di Fisica Nucleare of Italy, the Spanish Ministerio
de Educacion y Ciencia, the Conselleria d'Economia Hisenda i Innovacio
of the Govern de les Illes Balears, the Scottish Funding Council, the
Scottish Universities Physics Alliance, The National Aeronautics and
Space Administration, the Carnegie Trust, the Leverhulme Trust, the
David and Lucile Packard Foundation, the Research Corporation and the
Alfred P Sloan Foundation.
\bibliographystyle{unsrt}
|
\section{Introduction}
\label{sec:Introduction}
Radio surveys of the sky in the time domain have often been used to
identify new astrophysical phenomena. Highly variable radio sources
can serve as signposts to compact, high energy objects which are
accompanied by high magnetic fields and/or relativistic particle
acceleration. Radio variability from quasars and $\gamma$-ray bursts
(Dent 1965; Frail et al. 1997) was used to infer bulk relativistic
motions in these objects (Rees 1967; Goodman et al. 1987). Notable new
phenomena identified from radio time-domain surveys include the
discovery of the first pulsars (Hewish et al. 1968), the Galactic
high-energy binary LSI+61$^\circ$303 (Gregory and Taylor 1978), the
anomalous variability of 4C\,21.53 that lead to the discovery of
millisecond pulsars (Backer et al. 1982),
and the still-mysterious
extreme scattering events (Fiedler et al. 1987).
More recent surveys have found several new types of radio transients
whose identity has remained unknown or not well understood (e.g., Hyman et al. 2005;
McLaughlin et al. 2006; Bower et al. 2007; Lorimer et al. 2007;
Niinuma et al. 2007; Kida et al. 2008; Matsumura et al. 2009).
Specifically, Bower et al.
(2007) re-analyzed 944 epochs of Very Large Array\footnote{The Very
Large Array is operated by the National Radio Astronomy Observatory (NRAO),
a facility of the National Science Foundation operated under
cooperative agreement by Associated Universities, Inc.} (VLA)
observations, taken about once per week for twenty two years, of a
single calibration field. These authors discovered a total of ten
transients, eight in the 5-GHz band and two in the 8-GHz band. Eight
of these transients were detected in a single epoch.
Therefore, their duration is shorter than the
time between successive epochs (one week) and longer than the exposure
time (20\,min).
Moreover, the
majority of these sources do not have any optical counterpart coinciding
with their position.
The lack of optical counterparts down to limiting
magnitudes of 27.6 in $g$-band and 26.5 in $R$-band is especially puzzling
and significantly limits the classes of objects that can be associated
with these events (Ofek et al. 2010).
In a possibly related work, Kuniyoshi et al. (2006), Niinuma et al.
(2007), and Kida et al. (2008) reported a search for radio
transients using an East-West interferometer of the Nasu Pulsar
Observatory (located in Tochigi Prefecture, Japan) of Waseda
University. To date, this program reported 11 bright radio transients
with flux densities above 1\,Jy in the 1.4-GHz band.
Recently, in Ofek et al. (2010) we suggested that the properties of
the single epoch ``Bower et al. transients'' and the Nasu transients
are consistent with emerging from a single
class of objects, namely isolated old Neutron Stars (NS).
Specifically, the NS hypothesis is consistent with the rate,
energetics, sky surface density, source number count function and the
lack of optical counterparts.
In this paper we present a new VLA survey for radio transients and
variables at low Galactic latitudes. Our main motivation for this
survey was to detect more examples of this new class of short-lived
radio transients, with the goal of identifying them in real-time in
order to find their counterparts at other wavelengths for further
study. A second, and equally important motivation for this survey was
to characterize the transient and variable
radio sky with a sensitivity and cadence
which had not been carried out previously.
The organization of this paper is as follows.
In \S\ref{Prev} we provide a summary of previous radio transient and
variability surveys.
In \S\ref{SurveyObs} we present the observations,
while the data reduction is outlined in \S\ref{Red}.
The results from our real time transients search
are provided in \S\ref{RealTimeTran}.
\S\ref{PostSurveyCat} present the source catalogs generated in the post
survey phase.
The final post survey transient search is described
in \S\ref{PostSurveyTran} while the sources variability study is
presented in \S\ref{Variability}.
The implications of this study are discussed in \S\ref{Disc}
and we summarize in \S\ref{Sum}.
In addition three appendices discussing:
flux calibration;
the statistics of max/min of a time series;
and transient areal density calculation in the case of
a beam with non-uniform sensitivity
are provided.
\section{Previous GHz Surveys for Transients and Variables}
\label{Prev}
Existing 0.8-8\,GHz surveys have already explored,
to some extent, the dynamic radio
sky with a wide range of sensitivities, angular resolution and
cadences.
However, compared with synoptic surveys at higher frequencies
(infra-red to $\gamma$-rays)
the radio sky remains poorly explored.
In Table~\ref{Tab:PreviousSurveys} we summarize
past synoptic radio surveys.
For each survey we list also the number of transients, as well as variables
which vary by more than 50\%.
We note however, that comparison of these numbers is complicated
due to several factors.
A radio image
may be accomplished either through a single pointing,
or adding several scans taken at different times.
If the time span, $\delta{t}$, containing
all the observation composing a single ``epoch''
is larger than the transient duration
(or variability time scale)
then the survey sensitivity to transients (variables) is degraded.
Additionally, the probability of
detecting significant variability
depends on $\delta{t}$ and the typical time scale
between epochs ($\Delta{t}$), through the variability structure function.
Depending on the statistical method used to define
the variability amplitude, it may also affected by the
number of epochs ($N_{{\rm ep}}$) in the survey.
\begin{deluxetable*}{llllllllllll}
\tablecolumns{12}
\tabletypesize{\scriptsize}
\tablecaption{Previous GHz Transient and variability Surveys}
\tablehead{
\colhead{$\nu$} &
\colhead{Area} &
\colhead{Direction} &
\colhead{$\Delta{\theta}$} &
\colhead{N$_{{\rm ep}}$} &
\colhead{$\delta{t}$} &
\colhead{$\Delta{t}$} &
\colhead{rms} &
\colhead{Sources} &
\colhead{Tran.} &
\colhead{Var.} &
\colhead{Ref.} \\
\colhead{GHz} &
\colhead{deg$^{2}$} &
\colhead{deg} &
\colhead{$''$} &
\colhead{} &
\colhead{} &
\colhead{} &
\colhead{mJy} &
\colhead{} &
\colhead{} &
\colhead{} &
\colhead{}
}
\startdata
0.84 &2776 & $\delta<-30$ & $\sim45$ &2\tablenotemark{a} & 12\,hr & 1\,day--20\,yr & 2.8 & 29730 & 15 & $\sim10$ & [14] \\
1.4 & 0.22 & $l=150$, $b=+53$& 4.5 & 3 & 6\,hrs & 19\,d, 17\,m & 0.015 &\nodata & 0 & $2\%$ & [1] \\
1.4 & 2.6 & $l=151$, $b=+24$& 60 & 16 & 12\,hrs & 1-12\,d, 1-3\,m & 0.7 & 245 & 0 & $\sim1\%$ & [2] \\
1.4 & 120 & S. Galactic Cap & 5 & 2 & days & 7\,yr & 0.15 & 9086 & 0 & 1.4\% & [3] \\
1.4 & 2500 & $b\gtorder30$ & 45 & 2 & days & $\sim$years & 0.45 & 7181 & 1 &\nodata & [5,6,7] \\
1.4 & 2870\tablenotemark{b}& $+32>\delta>+42$&$24'\times2.4'$&$\sim1000$&4\,min &1\,d & 300 &\nodata &$11$ &\nodata & [8,9,10,11]\\
1.4 & 0.2 & $l=57$, $b=+81$ & 20 & 1852 & minutes & 1\,day--23\,yr & 2 & 10 & 0 &\nodata & [19] \\
1.4 & 690 & $l=70$, $b=+64$ & 150 & 2 & months & 15\,yr & 3.94 & 4408 & 0 &$\sim0.1\%$& [4] \\
1.4 & 690 & $l=70$, $b=+64$ & 150 & 12 & $>1$\,day & days--months & 38 & 4408 & 0 &$\ltorder0.5\%$& [20] \\
1.4 & 0.2 & phase calib. & \nodata & 151 & $5$\,min & days-years & $\sim1$&\nodata & 0 & \nodata & [21] \\
3.1 & 10 & $l=57$, $b=+67$ & 100 & 2 &months &15\,yr & 0.25 & 425 & 1\tablenotemark{c}&\nodata& [12] \\
4.9 & 0.1 & phase calib. & \nodata&$\sim390$\tablenotemark{d}&$5$\,min & days-years & $\sim1$&\nodata & 0 & \nodata & [21] \\
4.9 & 0.69 & Extragalactic & 0.5-15 & 2 &60\,min & 1-100d & 0.05 & \nodata & 0 & \nodata & [15] \\
4.9 & 23.2 &$\vert b\vert<0.4$& 5 & 3 &90\,s & 2\,m--15\,yr & 0.2 &2700 & 0 & 15 & [16] \\
4.9 & 500 & $\vert b\vert<2$& 180 & 16 &2\,min & 1\,day--5\,yr & 4.6 & 1274 & 1 & $\sim0.5$ & [18] \\
4.9 & 19924 & $75>\delta>0$ & 210 & 2 & $\sim$week & 1\,yr & 5 & 75162 & 0 & $>40$ & [17] \\
4.9 & 0.07 & $l=115$, $b=+36$& 5 & 626 &20\,mim & 1\,week--22\,yr & 0.05 & 8 & 7\tablenotemark{e}& 0 & [13] \\
8.5 & 0.02 & $l=115$, $b=+36$& 3 & 599 &20\,min & 1\,week--22y & 0.05 & 4 & 1\tablenotemark{f}& 0 & [13] \\
8.5 & 0.04 & phase calib. & \nodata&$\sim308$\tablenotemark{g}&$5$\,min & days-years & $\sim1$&\nodata & 0 & \nodata & [21] \\
\hline
4.9 & 2.6 & $\vert b\vert\approx7$ & 15 & 16 &50\,s & 1\,d--2\,yr & 0.15 &$\sim200$& 1 &$0.3-30\%$& This paper
\enddata
\tablecomments{{\it Columns description:} $\Delta{\theta}$ is the beam full width at half power; $N_{{\rm ep}}$ is the number of epochs;
$\delta{t}$ is the time span over which each epoch was obtained (see text);
$\Delta{t}$ is the range of time separations between epochs;
Sources is the number of persistent sources detected;
Tran. is the number of transients found by the survey;
Var. is the number or percentage of variables showing variability larger than 50\%.
We note that strong variables are defined differently by each survey.
Therefore, these numbers provide only a qualitative comparison
between the surveys.
References: [1] Carilli et al.~(2003), [2] Frail et al. (1994), [3] de Vries et al. (2004), [4] Croft et al. (2010), [5] Levinson et al. (2002), [6] Gal-Yam et al. (2006), [7] Ofek et al. (2010), [8] Matsumura et al. (2009), [9] Kida et al. (2008), [10] Kuniyoshi et al. (2007), [11] Matsumura et al. (2007), [12] Bower et al. (2010), [13] Bower et al. (2007), [14] Bannister et al. (2010), [15] Frail et al. (2003), [16] Becker et al. (2010), [17] Scott (1996), [18] Gregory \& Taylor (1986), [19] Bower \& Saul (2011), [20] Croft et al. (2011), [21] Bell et al. (2011).}
\tablenotetext{a}{Smaller fraction of the sky was observed more than twice.}
\tablenotetext{b}{The total surveyed area is about 2870\,deg$^{2}$, but about 460\,deg$^{2}$ was surveyed every day. These parameters are deduced from the Kida et al. (2008) and Matsumura et al. (2009) papers (see \S\ref{Prev}).}
\tablenotetext{c}{Marginal detection ($4.3\sigma$). Ignored in Figure~\ref{Fig:ArealDen_Flux_SurveySummary}.}
\tablenotetext{d}{Mean number of epochs per field - Seven fields were observed on 2732 epochs.}
\tablenotetext{e}{In addition, one transient was found by combining two month worth of data and no transients were found by combining 1\,yr worth of data.}
\tablenotetext{f}{In addition, one transient was found by combining two month worth of data and no transients were found by combining 1\,yr worth of data.}
\tablenotetext{g}{Mean number of epochs per field - Seven fields were observed on 2154 epochs.}
\label{Tab:PreviousSurveys}
\end{deluxetable*}
Here we provide a summary of some of the sky surveys listed in Table~\ref{Tab:PreviousSurveys}.
\subsection{Surveys at frequencies below 2\,GHz}
Carilli et al.~(2003) used a deep, single VLA pointing at
1.4 GHz toward the Lockman hole. They found
that only a small fraction, $\leq 2$\%, of radio sources above a
flux density limit of 0.1 mJy are highly ($>$50\%) variable on 19 day
and 17 month timescales. No transients were identified. Frail et al. (1994) imaged a
much larger field at 1.4 GHz toward a $\gamma$-ray burst with the
Dominion Radio Astrophysical Observatory
synthesis telescope, making daily measurements for two weeks and then
on several single epochs for up to three months. No transients were
identified on these timescales, and no sources above a flux density
limit of 3.5\,mJy were seen to vary by more than $4\sigma$.
There have also been a number of wide field surveys of the sky
at 1.4 GHz.
de Vries et al. (2004) used the VLA to image a region, toward
the South Galactic cap, twice on a seven year timescale. No transients were found
above a limit of 2\,mJy.
Croft et al. (2010; 2011) presented results from
the Allen
Telescope Array Twenty-centimeter Survey (ATATS).
They surveyed 690\,deg$^{2}$
of an extragalactic field on 12 epochs.
They compared the individual images with
their combined image and the combined image with the
NRAO VLA Sky Survey (NVSS; Condon et al. 1998).
No transients were found above a
flux density limit of 40 mJy in the combined image, with respect to the NVSS survey (Croft et al. 2010).
In addition, no transients were found in the
individual epochs above flux density of about 100\,mJy (Croft et al. 2011).
A systematic search for transients was
made between the two largest radio sky surveys,
Faint Images of the Radio Sky at Twenty-Centimeters
(FIRST; Becker et al. 1995) and
NVSS (Condon et al. 1998) by Levinson et al. (2002).
Nine transient candidates were identified. Follow-up observations of
these established that only one was a genuine transient -- a likely
radio supernova in NGC\,4216
(Gal-Yam et al. 2006; Ofek et al. 2010).
We note that each FIRST and NVSS image is composed of $\approx4$ overlapping beams
of adjacent regions taken typically with $\delta{t}\sim$days
(Becker et al. 1995; Condon et al. 1998; Ofek \& Frail 2011).
Therefore, if the duration ($t_{{\rm dur}}$) of the Bower et al.
transients is shorter than this typical time between images
composing ``one epoch'' then the sensitivity of these surveys for transients is degraded by
$\approx\sqrt{4}$.
However, the Levinson et al. survey used a flux density limit of 6\,mJy
which is ($\gtorder2$) higher than the flux limit of these surveys.
Therefore, its efficiency for Bower et al. like
transients is not degraded.
Bright ($>1$\,Jy), short-lived transients,
have been reported by the
Nasu 1.4 GHz survey (e.g., Matsumura et al. 2009; see \S\ref{sec:Introduction}).
Croft et al. (2010; 2011) argued that these transients are not real
because their implied event rate cannot be reconciled with their own survey
unless this population has
sharp cut off at flux densities below 1 Jy.
We note that Croft et al. adopted the Nasu transients
areal density
reported in Matsumura et al. (2009).
However, this areal density is inconsistent with the rate reported
in Kida et al. (2008) which is roughly two orders of magnitude lower.
Furthermore, based on the Nasu survey parameters reported
in Matsumura et al. (2009) we estimate that the areal density
of the Nasu transients is roughly two orders of magnitude lower
than that stated in their
paper\footnote{This inconsistency was already mentioned by
Bower \& Saul (2011). However, they reach a different conclusion than we do.
More information about the Nasu survey is needed in order to resolve this issue.}.
Here we present an estimate of the rate of transients from
the Nasu observations:
Matsumura et al. (2009) reported that they discovered (at that time)
nine transients over a period of two years (730\,days).
Assuming that they have
four pairs of antennas ($N_{ant}$) each looking at a different position
(near the local zenith $\delta\approx 37$\,deg)
with a field of view of $W=0.4$\,deg and scanning the sky at the
sidereal rate, the total
sky area scanned by their system after two years
is $\approx 3.4\times10^{5}$\,deg$^{2}$ ($\cong N_{ant}W \times 360\cos(37^{\circ})\times730$).
Therefore,
the total areal density of their transients is
the number of transients divided by the total scanned sky area
and divided by 1.61, which is
$\approx1.7\times10^{-5}$\,deg$^{-2}$.
Where the 1.61 factor is due to the fact that their sensitivity is not
uniform within the beam and, and is calculated using Equation~\ref{Eq:NbUni} in
Appendix~\ref{Ap:Rate}.
Assuming the transients duration is 1\,day than their rate is
$\approx 0.02$\,deg$^{-2}$\,yr$^{-1}$.
We note that our derived Nasu transients rate
is roughly consistent with the upper limit on the rate
given by Kida et al. (2008).
Therefore,
we speculate that there was a confusion between areal density and transient
rate in the Matsumura et al. (2009) paper.
We conclude that if our estimate for the areal density
for the Nasu transients is correct than
the ATATS survey results cannot decisively rule out the reality
of the Nasu transients.
A comprehensive survey at 0.8\,GHz was reported
by Bannister et al. (2010).
They surveyed 2776\,deg$^{2}$ south of $\delta=-30^\circ$ over a 22-year period.
Out of about 30,000 sources they identified 53 variables and 15 transient
sources.
Recently, Bower \& Saul (2011) reported a transient search in
the fields of the VLA calibrators in which no transients were found
(Summarized in Table~\ref{Tab:PreviousSurveys} and Figure~\ref{Fig:ArealDen_Flux_SurveySummary}).
Another related work by Bell et al. (2011) searched for radio
transients in the fields of the VLA phase calibrators at 1.4\,GHz, 4.9\,GHz and 8.5\,GHz.
Based on their survey parameters (Table~\ref{Tab:PreviousSurveys})
we estimate that their $95\%$ confidence surface density upper limit on transients
brighter than 8\,mJy
in 1.4\,GHz, 4.9\,GHz and 8.5\,GHz
are 0.19\,deg$^{-2}$, 0.13\,deg$^{-2}$ and 0.50\,deg$^{-2}$, respectively.
These values are corrected for the beam non-uniformity factor
(of 1.61), mentioned earlier.
We assumed that Bell et al. (2011) searched for transients within the full width at half power
of the beam.
For clarity purposes,
in Figure~\ref{Fig:ArealDen_Flux_SurveySummary} we show only the 4.9\,GHz limit.
\subsection{Surveys at frequencies above 2\,GHz}
There have been several additional transient surveys
carried out at frequencies
above 1.4 GHz.
At 3.1 GHz Bower et al.
(2010) report a marginal detection of one possible transient ($4.3\sigma$) in a
10\,deg$^2$ survey of the Bo\"otes extragalactic field. In a five year
catalog of radio afterglow observations of 75 $\gamma$-ray bursts,
Frail et al. (2003) found several strong variables at 5 and 8.5
GHz, but no new transients apart from the radio afterglows themselves.
Two surveys at 5\,GHz have specifically targeted the Galactic plane.
Taylor and Gregory (1983) and Gregory and Taylor
(1986) used the NRAO 91-m telescope to image an
approximately 500\,deg$^2$ region from Galactic longitude $l=40$\,deg to $l=220$\,deg
with Galactic latitude $\vert{b}\vert\leq 2$\,deg in 16 epochs over a 5-year period.
They identified one transient candidate which underwent a 1\,Jy flare
but for which follow-up VLA observations showed no quiescent radio
counterpart (Tsutsumi et al.~1995). They also claimed tentative evidence for a
separate Galactic population of strong variables comprising 2\% of their
sources. Support for this comes from Galactic survey of Becker et al.
(2010), who find about one half of their variable source
sample (17/39), or 3\% of all radio sources in the Galactic plane, undergo
strong variability on 1-year and 15-year baselines.
We note that the surface density of radio sources in
the Galactic plane is only slightly higher ($\approx 20$\%)
than at high Galactic
latitudes (Helfand et al. 2006; Murphy et al. 2007).
One of the largest variability survey of its kind was carried out using the
7-beam receiver on the NRAO 91-m telescope (Scott 1996; Gregory et al. 2001).
The sky from 0$^\circ \leq \delta\leq 75^\circ$
was surveyed over two 1-month periods in 1986 November and 1987 October
(Condon, Broderick \& Seielstad 1989;
Becker et al. 1991;
Condon et al. 1994;
Gregory et al. 1996).
The final catalog, made by combining both the 1996 and 1997 epochs,
contained 75,162 discrete sources with flux densities $>18$ mJy.
Long term variability information was available for the majority
of the sources
by comparing the mean flux densities between the 1986 and 1987 epochs.
Scott (1996) carried out a preliminary analysis of the long-term
measurements and identified 146 highly variable sources, or $<$1\% of
the cataloged radio sources.
Eight possible transients
in the Scott (1996; Table 5.1) list appear in either 1986 or 1987
but are undetected in the other epoch ($<2$$\sigma$). Two sources are previously identified
variables from the Gregory and Taylor (1986) survey, while six are flagged as
possible false positives due to confusion by nearby bright sources. One source
(B150958.3$+$103541) was $9\pm6$\,mJy in 1986 and $75\pm7$\,mJy in 1987
but it is in both the FIRST and NVSS source catalogs. There are therefore no
long-term transients identified in the Scott (1996) survey.
In order to estimate the flux limit above which the Scott (1996)
comparison between the 1986 and 1987 surveys
is complete, we compared the source numbers near the celestial equator,
as a function of flux in the two publicly available catalogs
from 1987 and the combined 1986/1987 catalog.
We found that at flux densities below about 40\,mJy
the number of sources, as a function of flux,
in the 1987 catalog is
rising slower than that for the one of the deeper combined catalog.
Therefore, we estimate that near the terrestrial equator,
the 1987 catalog is complete
above a flux density of about 40\,mJy.
However, these catalogs were made from observations
in which each point on the sky was observed
$\approx 4/\cos(\delta)$ times taken within a few days.
This degrades the sensitivity of the comparison carried out by Scott (1996),
for $\ltorder 1$\,day transients,
by about $\sqrt{4}$.
Therefore, we conclude that the Scott (1996) survey is sensitive to short term ($\ltorder 1$\,day)
transients brighter than about 80\,mJy ($=40\sqrt{4}$).
Finally, assuming Scott (1996) did~not find any transients in two epochs,
we put a $2$-$\sigma$ upper limit on the areal density
of $\ltorder 1$\,day transients
brighter than 80\,mJy, of $9.5\times10^{-5}$\,deg$^{-2}$.
In summary, despite the heterogeneous nature of these GHz surveys, it
is clear that the radio sky is relatively quiet compared
with the $\gamma$-rays sky.
The fraction of strong variables
among the persistent radio source population is $0.1$--$3$\% from flux densities of 0.1\,mJy
to 1\,Jy.
However, the exact percentage of strong variables is still uncertain because of the
different criteria used by various surveys.
We note that
within this flux density range, radio source populations are dominated by AGN
roughly above 1\,mJy and star forming galaxies dominates the source counts
below 1\,mJy (Condon 1984; Windhorst 1985).
The transient areal densities detected by these various surveys,
as well as our survey,
are shown graphically in Figure~\ref{Fig:ArealDen_Flux_SurveySummary}.
Also shown in this figure are the persistent sources
areal densities at different frequencies.
This plot is further discussed in \S\ref{Disc}.
\begin{figure*}
\centerline{\includegraphics[width=16cm]{f1.eps}}
\caption{Cumulative areal density of radio sources and transients as a function
of flux density for various surveys.
Different colors represent different frequencies as specified in the legend.
95\% confidence upper limits from various transient surveys are shown
as right-angle corners, while measured areal densities are marked as filled circles.
All the error bars represent 2-$\sigma$ confidence intervals.
We note that the Nasu survey rate is based on our estimate using the
survey description in their papers (see \S\ref{Prev}).
Because some of the parameters
of this survey are unknown to us, we increase the error bars for
this survey to include a factor of two uncertainty.
For the Levinson et al. survey we mark the
areal density of the single transient found in this
survey (a supernova in NGC~4216)
and also the 95\% confidence upper limit
assuming there are no Bower et al. transients in the survey.
For the de Vries et al. (2004) search we use a flux limit of 2\,mJy
since they used the FIRST survey in which each epoch is
composed of about four observations of the same field taken within $\delta{t}\approx$days.
Therefore, this degrade their survey flux limit sensitivity to Bower et al. transients by a factor
of about $\sqrt{4}$.
The right-hand-side y axis shows the transients rate
assuming a transient duration of 0.5\,days.
Some surveys are excluded from this plot.
For example,
Becker et al. (2010) restricted their catalog to sources
detected in at least two out of three epochs,
or that have a confirmed detection at 1.4\,GHz.
Therefore, such surveys are not included here.
Also shown are the areal densities of persistent sources at 1.4\,GHz
and 4.9\,GHz based on the FIRST and GB87 surveys,
respectively (dashed lines).
\label{Fig:ArealDen_Flux_SurveySummary}}
\end{figure*}
\section{Survey Observations}
\label{SurveyObs}
We designed a survey to look for transients and variable sources near
the Galactic plane, with typical time scales of days to two years
at milli-Jansky flux levels. We were specially interested in finding
``Bower et al. transients'', conduct multi-wavelength follow up of these
events and finding counterparts and studying their spectral
evolution.
\subsection{Survey Design}
\label{Design}
We used the VLA to observe 141 pointings along the Galactic
plane. In order to minimize telescope motions we selected all the
pointings in four regions. The median longitude ($l$) and latitude
($b$) of the four regions are: $l=22.6$\,deg, $b=-6.7$\,deg;
$l=56.6$\,deg, $b=-5.5$\,deg;
$l=89.7$\,deg, $b=-7.8$\,deg; $l=106.0$\,deg, $b=-6.5$\,deg.
Within each region we
selected 26--42 pointings within 2.3\,deg from the median position of
each region. Each pointing was selected to have no NVSS sources
brighter than 1\,Jy within 3\,deg, no NVSS sources brighter than
300\,mJy within 1\,deg, and no source brighter than 100\,mJy within
the field of view as defined by the half power radius.
We also rejected fields which distance from known Galactic supernova remnants (SNR; Green 2001)
is within twice the diameter of the SNR.
The typical distance between pointings in each
region is about $20'$. The final 141 pointings are listed in
Table~\ref{Tab:ListOfFields}.
\begin{deluxetable}{llll}
\tablecolumns{4}
\tabletypesize{\scriptsize}
\tablewidth{0pt}
\tablecaption{List of survey pointings}
\tablehead{
\colhead{Field Name} &
\colhead{RA} &
\colhead{Dec} &
\colhead{$N_{{\rm ep}}$} \\
\colhead{} &
\colhead{deg} &
\colhead{deg} &
\colhead{}
}
\startdata
1851$-$1327 & $282.94699$ & $-13.45500$ & 16\\
1852$-$1309 & $283.09963$ & $-13.15668$ & 16\\
1853$-$1233 & $283.40439$ & $-12.56005$ & 16\\
1853$-$1251 & $283.25209$ & $-12.85836$ & 16\\
1853$-$1318 & $283.40625$ & $-13.30508$ & 16
\enddata
\tablecomments{List of all 141 fields that were observed as part of this survey. The number of epochs per pointing is marked in $N_{{\rm ep}}$. This table is published in its entirety in the electronic edition of the {\it Astrophysical Journal}. A portion of the full table is shown here for guidance regarding its form and content.}
\label{Tab:ListOfFields}
\end{deluxetable}
\subsection{Observations}
\label{Obs}
These 141 fields were observed on 11 epochs using the VLA
in the Summer of 2008 and on five epochs using the Expanded VLA (EVLA)
during the Summer of 2010.
All observations were made in the compact D configuration.
For the 2008
observations, we added together two adjacent 50 MHz bandwidths
centered at 4835 and 4885\,MHz with full polarization.
For the 2010 observations we added together two adjacent 128\,MHz
sub bands centered at 4896 and 5024\,MHz with full polarization.
In 2008 care was taken to ensure that the local sidereal start time
was the same for each 3-hr epoch (20:30 LST).
Therefore, each field
was observed at the same hour angle and subsequently the synthesized
beam stayed the same for each epoch, varying only when antennas were
taken out of the array. The 2010 observations were taken during EVLA
shared-risk science commissioning, and so some scans lost due to
correlator errors and the last two epochs began one hour earlier than
our 2008 local sidereal start time.
We integrated each pointing for about 50\,s on average. The maximum
integration time was 58.5\,s and the minimum was 43.3\,s. Additionally,
during each 3-hr observing run we carried out all necessary
calibrations. Amplitude calibration was achieved with observations of
3C\,286 and 3C\,147 at the start and end of each epoch, respectively.
Phase calibration was checked every 20-25\,min by switching to a bright
point source within a few degrees of the targeted region.
We used the following four phase calibrators (one per each region):
J1911$-$201, J1925$+$211, J2202$+$422, and J2343$+$538.
The total
calibration and
antenna move-time overhead was about $30$\% of the observing time.
This overhead on move time
could have been lowered had we used the fast slew methods from the
NVSS and FIRST surveys (Condon et al. 1998, Becker et al. 1995) with
a resulting increase in the number of square degrees of sky surveyed
per hour. However, since we recently found that this method could
introduce spurious transients (Ofek et al. 2010), we adopted a less
efficient but more robust observing method.
In Table~\ref{Tab:ListOfEpochs} we list the time of the UTC midpoint
\begin{deluxetable*}{lllccccc}
\tablecolumns{8}
\tablecaption{Observing Epochs}
\tablehead{
\colhead{Epoch} &
\colhead{Date} &
\colhead{Time Elapsed} &
\colhead{$\langle$rms Noise$\rangle$}&
\colhead{Observed} &
\colhead{Number} &
\colhead{Gain Corr.} &
\colhead{Cosmic error} \\
\colhead{} &
\colhead{UTC} &
\colhead{days} &
\colhead{$\mu$Jy} &
\colhead{Fields} &
\colhead{of Sources} &
\colhead{} &
\colhead{$\%$}
}
\startdata
1 & 2008 Jul 15.40 & 0.00 & 243 & 141 & 343 & 1.029 & 5.3 \\
2 & 2008 Jul 18.39 & 2.99 & 181 & 141 & 155 & 0.987 & 4.6 \\
3 & 2008 Jul 19.39 & 3.99 & 229 & 141 & 363 & 1.033 & 4.6 \\
4 & 2008 Aug 10.33 & 25.93 & 178 & 141 & 166 & 0.971 & 1.0 \\
5 & 2008 Aug 11.33 & 26.93 & 183 & 141 & 164 & 0.980 & 1.1 \\
6 & 2008 Aug 14.32 & 29.92 & 174 & 141 & 162 & 1.002 & 0.4 \\
7 & 2008 Aug 16.31 & 31.91 & 173 & 139 & 151 & 1.023 & 1.7 \\
8 & 2008 Aug 18.29 & 33.89 & 178 & 141 & 155 & 0.924 & 1.7 \\
9 & 2008 Aug 25.29 & 40.89 & 196 & 141 & 183 & 1.016 & 2.4 \\
10& 2008 Aug 28.28 & 43.88 & 178 & 141 & 157 & 0.956 & 3.5 \\
11& 2008 Aug 30.28 & 45.88 & 186 & 141 & 170 & 1.033 & 5.3 \\
12& 2010 Jul 16.42 & 731.02 & 105 & 141 & 216 & 1.020 & 1.9 \\
13& 2010 Jul 18.43 & 733.03 & 111 & 134 & 199 & 1.022 & 0.8 \\
14& 2010 Jul 22.35 & 736.95 & 108 & 109 & 158 & 0.992 & 0.6 \\
15& 2010 Jul 23.32 & 737.92 & 104 & 141 & 200 & 1.008 & 0.7 \\
16& 2010 Jul 25.31 & 739.91 & 116 & 140 & 217 & 1.003 & 2.1
\enddata
\tablecomments{List of the 16 epochs. The dates indicate the
observations mid-time. In practice, in all the instances in which we find
that the cosmic error is smaller than 3\% we replaced it by 3\% (see \S\ref{PostSurveyCat}).}
\label{Tab:ListOfEpochs}
\end{deluxetable*}
for each epoch with some additional information. The shortest
variability timescale sampled was 24\,hr and the longest was
$\cong2$\,yr. By design, the cadence of the 2008 survey was chosen to
probe variability timescales between
a day and a month. A longer 2\,yr timescale was also
sampled by comparing deep images made from the 2008 and 2010 campaigns
(see below).
\section{Data reduction and calibration}
\label{Red}
In 2008 the {\it uv} data was streamed directly to a disk in real time,
and a pipeline was ran after each one of the four regions were
observed. We used the data reduction pipeline provided in the
Astronomical Image Processing System (AIPS)
package\footnote{http://www.aips.nrao.edu/}.
For each epoch, the pipeline first flagged and calibrated the {\it uv}
data. It then imaged a 30-arcmin-wide field around all 141 pointings,
deconvolving down to three times the rms noise, and restoring the image
with a robust weighted beam. No self-calibration was done. The VLA
data rates ($\approx 30$\,Mbytes\,hr$^{-1}$) and the D-configuration image
requirements (512 pixels, 3.6$^{\prime\prime}$\,pixel$^{-1}$) were so modest
that the entire pipeline reduction and the variable source analysis
(see \S\ref{RealTimeTran}) was completed before the VLA finished observing the
next region ($\approx$40 min). The real-time analysis
capability was not available in 2010 but the data were also calibrated
within AIPS following standard practice.
As the experiment progressed we built up reference images, made by
summing all previous epochs. These deeper images proved useful in the
real-time search for transient sources (\S\ref{RealTimeTran}). After the
survey was completed, a final set of images was made separately for
each yearly campaign using the data from the 11 epochs in 2008
and the five epochs in 2010. We also summed the 2008 and
2010 deep images to create 16-epoch Master images for the entire
experiment. In summary, there were three final image datasets; the
Single epoch images, the Yearly images (for 2008 and 2010 separately)
and the final Master images made from all available data.
Our final survey parameters are given in Table~\ref{Tab:SurveyPar}.
The effective survey area was calculated using the full width at
half-power ($9.3'$ at 4.86 GHz; see however Appendix~\ref{Ap:Rate})
but the searches for transients in real time and for
variability were made over a larger area -- out to the 15\% response
point of the primary beam ($15'$ diameter).
For the analysis, no correction was made
for the primary beam attenuation, in order to maintain uniform noise
statistics over the entire field.
The synthesized beam
and rms noise estimates for different epochs and different pointings
varied by factors close to unity. The values in
Table~\ref{Tab:ListOfEpochs} are averages for each epoch over all
pointings, while Table~\ref{Tab:SurveyPar}
gives the mean rms values (over all fields) in the Master images and the
2008 and 2010 combined images.
\begin{deluxetable}{ll}
\tablecaption{Survey Parameters}
\tablehead{
\colhead{Property} &
\colhead{Value}\\
}
\startdata
Frequency & 4.9 GHz \\
Observing time & 48 hrs \\
Survey Area & 2.66 deg$^2$ \\
Angular resolution & 15$^{\prime\prime}$ \\
Repeats & 16 \\
Timescales & 1\,day--2\,years\\
No. of fields & 141 \\
mean exposure Time per field & 50 s \\
mean rms per epoch (2008) & 190 $\mu$Jy \\
mean rms per epoch (2010) & 109 $\mu$Jy \\
mean rms per 11 epochs (2008) & 72 $\mu$Jy \\
mean rms per 5 epochs (2010) & 56 $\mu$Jy \\
mean rms in Master images & 46.7$\mu$Jy
\enddata
\tablecomments{}
\label{Tab:SurveyPar}
\end{deluxetable}
Throughout the paper we state explicitly if we use corrected
or uncorrected fluxes.
``Corrected fluxes'' are corrected
for beam attenuation and for the CLEAN bias
by adding additional $+0.3$\,mJy
(e.g., Becker et al. 1995; Condon et al. 1998).
In order to maintain uniform statistics we chose to search all images
to the same depth, rather than compute a new threshold for each image
individually. In practice, this led to some false positives for
noisier than average epochs and fields with bright point sources.
Indeed, Table~\ref{Tab:ListOfEpochs} indicates that the nosiest
epochs contains larger number of sources.
\section{Real-Time Transient Search}
\label{RealTimeTran}
We employed two distinct analysis strategies for transient and
variable source identification. The first, which is discussed in this section,
was a real-time analysis.
The main motivation was to rapidly identify any short-lived
sources and mark them for immediate follow-up
at other wavelengths. The second was a post-survey analysis which was
carried out after all the epochs had been observed.
The main goals of
this second phase were to carry out a more in-depth search for
Bower et al. transients (\S\ref{PostSurveyCat})
and to characterize the variability properties
of the persistent source population (\S\ref{PostSurveyTran}).
For the real-time identification
the images (\S\ref{Red}) were searched visually for any new or
strongly varying sources by comparing them with individual epochs,
and by comparing them with a reference image
made by summing all previous epochs.
Any candidate
variable source which we identified was subject to a more detailed light
curve and position fitting analysis before deciding to trigger radio, optical
and/or X-ray follow-up observations.
The followup visible light observations were carried out using the
robotic Palomar $60''$ telescope (P60; Cenko et al. 2006)
and the Keck-I~10-m telescope. The UV and X-ray observations
were conducted by the {\em Swift} satellite (Gehrels et al. 2004).
We note that prior to and during the VLA campaign we obtained
visible light reference images for most of our fields
using the P60 telescope.
We identified two possible transients that were deemed interesting
enough for multi-wavelength follow-up.
However, followup VLA observations and a careful post observing
re-analysis (\S\ref{PostSurveyTran})
showed that these are not real transients.
One source, J213438.01$+$414836.0,
was a sidelobe artifact,
while the second source, J230424.68$+$530414.7
is a long term variable that had crossed our single-epoch noise
threshold on 2008 July 19 and is clearly seen in the 2008 and 2010 deep
coadds.
We note that both sources were observed using
the P60 telescope about 2\,hr and 1\,hr
after the radio observations were obtained, respectively.
Furthermore {\it Swift}-XRT observations of the
first source were obtained about five days after it was found.
These fast response observations
demonstrate our near real time followup capabilities.
\section{Post survey source catalog}
\label{PostSurveyCat}
The next phase of our analysis occurred after
the conclusion of the observations.
We generated source catalogs in order to search for
short-lived transients and to carry out a variability study of all
identified sources. The AIPS task {\it SAD} (search and destroy) was used for source
finding.
We found that
false positive sources came from one of two main reasons: slightly resolved
sources and sidelobe contamination.
Extended sources were
identified by requiring that their integrated flux density was within a
factor of two of their peak flux density.
False sources created by
scattered power from the snapshot sidelobe response was only a
significant problem for the six fields with sources whose flux density
exceeded 20\,mJy. We flagged any variables or transients from these
fields for visual inspection.
Three catalogs were created. The first catalog is the ``Single epoch
catalog'', generated by running {\it SAD} on a 15-arcmin diameter
region for each single-epoch
image individually (Table~\ref{Tab:SingleEpochCat}).
A second ``Master
catalog'' (Table~\ref{Tab:MasterCat})
was created by running {\it SAD} on the final master
images made from all available data (16 epochs), while a third catalog,
``Yearly catalog'', was generated on the yearly images (for 2008 and
2010 separately).
\begin{deluxetable*}{lllllllllllll}
\tablecolumns{13}
\tabletypesize{\scriptsize}
\tablewidth{0pt}
\tablecaption{Single epoch catalog}
\tablehead{
\colhead{Epoch} &
\colhead{Source} &
\colhead{Field Name} &
\colhead{J2000 RA} &
\colhead{J2000 Dec} &
\colhead{$f_{\nu,{\rm p}}^{{\rm cor}}$} &
\colhead{$f_{\nu,{\rm p}}^{{\rm uncor}}$} &
\colhead{$\sigma_{{\rm p}}$} &
\colhead{$f/f_{{\rm p}}$} &
\colhead{Major} &
\colhead{Minor} &
\colhead{PA} &
\colhead{$\Delta_{{\rm C}}$} \\
\colhead{} &
\colhead{} &
\colhead{} &
\colhead{deg} &
\colhead{deg} &
\colhead{mJy} &
\colhead{mJy} &
\colhead{mJy} &
\colhead{} &
\colhead{$''$} &
\colhead{$''$} &
\colhead{deg} &
\colhead{$'$}
}
\startdata
1 & 1 & $1851-1327$ & 282.955970& $-13.502335$& 329.63& 25.39 & 0.26 & 1.03 & 23.67 & 13.67 & 23.7 & 2.89\\
1 & 2 & $1851-1327$ & 282.937502& $-13.432227$& 3.29& 1.23 & 0.26 & 1.00 & 23.49 & 13.31 & 29.7 & 1.48\\
1 & 3 & $1851-1327$ & 282.855004& $-13.349517$& 8.60& 1.12 & 0.26 & 0.64 & 23.66 & 8.48 & 13.6 & 8.30\\
1 & 4 & $1853-1233$ & 283.399554& $-12.561197$& 7.59& 6.59 & 0.26 & 0.95 & 21.88 & 13.15 & 20.3 & 0.29\\
1 & 5 & $1853-1233$ & 283.474902& $-12.562245$& 4.74& 3.10 & 0.26 & 0.89 & 22.97 & 11.74 & 21.5 & 4.13
\enddata
\tablecomments{Catalog of 2953 sources detected in Single epochs.
{\it Columns description:} Epoch is the epoch number (see Table~\ref{Tab:ListOfEpochs}), and source is a serial source index in epoch.
$f_{\nu,{\rm p}}^{{\rm cor}}$, $f_{\nu,{\rm p}}^{{\rm uncor}}$, and $\sigma_{{\rm p}}$
are the corrected peak flux density, uncorrected peak flux density,
and the error in the uncorrected peak flux density, respectively.
Corrected flux are corrected for beam attenuation and the
CLEAN bias.
$f/f_{{\rm p}}$ is the integrated flux divided by the peak flux.
Major and Minor are the major and minor axes object size, while PA is the position angle
of the major axis.
Finally, $\Delta_{{\rm C}}$ is the distance of the source from the beam center.
This table is published in its entirety in the electronic edition of the {\it Astrophysical Journal}.
A portion of the full table is shown here for guidance regarding its form and content.
\label{Tab:SingleEpochCat}}
\end{deluxetable*}
The Master catalog has the merit of being able to identify persistent
radio sources approximately $\sqrt{N}$-times fainter than any
individual epoch (where $N=16$), but it is $\sqrt{N}$-times less
sensitive to a short-lived transient that might be identified in a single-epoch image.
For our Single epoch catalogs we used
a flux density cutoff of 1\,mJy in 2008 and 0.76\,mJy in 2010,
and the number of sources in each epoch are given in Table~\ref{Tab:ListOfEpochs}.
Our Master catalog consisted of 464 sources which
are listed in Table~\ref{Tab:MasterCat}, with a flux density
cutoff of $0.28$\,mJy. The Yearly catalog had flux density cutoff of
0.5\,mJy in 2008 and $0.35$\,mJy in 2010. These cutoffs corresponds to
about a 5-7$\sigma$ threshold, depending on the rms noise for
individual fields.
The electronic version of the Master catalog also contains
the peak flux of each source.
This was measured in
the Single epoch images
at the position of the sources
found in the Master image.
We used the Master catalog to perform a second order amplitude
calibration that would tie together the flux density scale for all
epochs. Normally self-calibration could be used to find additional
gain variations within a radio observation but our survey was designed to
avoid pointings with bright point sources.
Our approach assumes that each VLA epoch (all the observations
in each epoch were taken within 3\,hours) shares the same ``gain''
correction, and we solved for these ``nightly'' gain corrections by
fitting, using least squares minimization, the equation
\begin{equation}
m_{ij} = z_{i} + \bar{m}_{j},
\label{Eq:RelPhot}
\end{equation}
where $m_{ij}$ is the ``magnitude'': $-2.5\log_{10}{f_{ij}}$, $f_{ij}$
is the peak specific flux of the $j$-th source in the $i$-th epoch, $z_{i}$ is
the gain correction for the $i$-th epoch (in units of magnitudes)
and $\bar{m}_{j}$ is a
nuisance parameter representing the best fit mean magnitude of the
$j$-th source.
We note that, as explained in Appendix~\ref{Ap:ZP},
magnitudes have convenient statistical properties.
The final multiplicative gain corrections are
$10^{-0.4z_{i}}$.
This method is described in detail in
Appendix~\ref{Ap:ZP}, and the best fit multiplicative gain corrections are
listed in Table~\ref{Tab:ListOfEpochs}.
Similarly, we also derived the yearly gain corrections for the
Yearly epochs.
These gain
corrections are $1.004$ and $0.996$ for 2008 and 2010, respectively.
The flux errors reported by {\it SAD} do~not include
any systematic error terms.
Typically, VLA calibration is assumed to be good
to a level of $3\%$ or better (e.g., Condon et al. 1998).
In order to check if some epochs are nosier we estimated
the ``cosmic errors'', $\epsilon_{{\rm cos}}$, using the following scheme.
We measured
the standard deviation in the flux of the four phase calibrators
observed on each night, after normalizing their flux by their mean
flux over all the epochs taken at the same year. The cosmic errors
estimated using this method are listed for each epoch in
Table~\ref{Tab:ListOfEpochs}.
This estimate is based on a small number of sources
and these sources may be variable.
Therefore, this should be regarded as a rough estimate.
In some instances, the cosmic errors we estimated
were smaller than $3\%$ and in those cases we replaced
the cosmic errors for these epochs by $3\%$.
We note that if indeed the
cosmic errors in some cases are smaller than 3\%,
then our strategy of adopting a larger cosmic errors
may reduce the number of variables
found in our survey.
For the Yearly catalogs, we used the
mean cosmic error terms of the individual epochs in each year.
These are $0.028$ and $0.012$ for 2008 and 2010, respectively.
Equipped with the gain corrections and an estimate for the cosmic errors
we next corrected the flux measurements of all the sources using the
gain correction factors and added in quadrature the cosmic errors to the
peak flux errors.
The new fluxes and errors were used in all the plots
and the calculation of the light curves statistical properties.
\begin{deluxetable*}{rrrrlllrlrllrll}
\tablecolumns{15}
\tabletypesize{\scriptsize}
\tablewidth{0pt}
\tablecaption{Master catalog}
\tablehead{
\multicolumn{11}{c}{Current search} &
\multicolumn{2}{c}{NVSS} &
\multicolumn{1}{c}{USNO} &
\multicolumn{1}{c}{2MASS} \\
\colhead{J2000 RA} &
\colhead{J2000 Dec} &
\colhead{$f_{{\rm p}}^{{\rm cor}}$} &
\colhead{$f_{{\rm p}}^{{\rm uncor}}$} &
\colhead{$\sigma_{{\rm p}}$} &
\colhead{$N_{{\rm obs}}$} &
\colhead{$N_{{\rm det}}$} &
\colhead{$\chi^{2}$} &
\colhead{StD/$\langle f\rangle$} &
\colhead{$\chi^{2}_{{\rm Y}}$} &
\colhead{StD/$\langle f\rangle_{Y}$} &
\colhead{Dist} &
\colhead{$\alpha$\tablenotemark{a}} &
\colhead{Dist} &
\colhead{Dist} \\
\colhead{deg} &
\colhead{deg} &
\colhead{mJy} &
\colhead{mJy} &
\colhead{mJy} &
\colhead{} &
\colhead{} &
\colhead{} &
\colhead{} &
\colhead{} &
\colhead{} &
\colhead{$''$} &
\colhead{} &
\colhead{$''$} &
\colhead{$''$}
}
\startdata
283.675661&$-13.518385$& 7.24& 2.48& 0.05& 15& 15& 50.23& 0.17& 2.54& 0.06& 1.0& 1.03& \nodata& \nodata\\
283.682476&$-13.444363$& 2.52& 1.85& 0.05& 15& 15& 52.34& 0.18& 0.05& 0.02& 3.8& 0.17& \nodata& \nodata\\
284.193460&$-13.325061$& 47.83& 32.01& 0.10& 14& 14& 70.30& 0.08& 0.48& 0.02& 0.9& -0.07& \nodata& 0.6\\
284.387195&$-12.213881$& 11.03& 6.74& 0.06& 15& 15& 142.60& 0.14& 2.85& 0.05& 4.4& -0.54& \nodata& \nodata\\
283.976980&$-12.165968$& 93.19& 57.18& 0.13& 15& 15& 468.40& 0.18& 68.27& 0.25& 0.9& 0.08& \nodata& \nodata\\
283.949467&$-11.356239$& 7.65& 4.69& 0.05& 15& 14& 53.20& 0.10& 0.84& 0.03& 2.9& 1.14& \nodata& \nodata\\
284.592955&$-10.949209$& 13.07& 10.66& 0.06& 15& 15& 48.94& 0.08& 3.98& 0.06& 4.2& 1.26& \nodata& \nodata\\
284.497521&$-10.459850$& 2.70& 2.11& 0.05& 15& 14& 46.02& 0.14& 1.71& 0.06& 10.0& 0.83& \nodata& \nodata\\
298.720898&$ 16.755257$& 10.92& 8.39& 0.04& 16& 16& 118.76& 0.10& 5.02& 0.08& 1.0& -0.80& \nodata& \nodata\\
299.583716&$ 18.171732$& 6.93& 5.82& 0.04& 16& 16& 143.31& 0.14& 15.03& 0.12& 2.3& -0.20& \nodata& \nodata\\
299.271523&$ 18.226378$& 4.81& 3.65& 0.05& 16& 16& 75.27& 0.12& 0.67& 0.04& \nodata& \nodata& \nodata& \nodata\\
299.353219&$ 18.473931$& 41.09& 6.89& 0.05& 16& 16& 149.27& 0.13& 13.66& 0.13& 0.4& 0.14& \nodata& \nodata\\
300.337088&$ 19.936532$& 7.71& 5.50& 0.04& 16& 16& 55.62& 0.08& 0.12& 0.01& 2.2& 0.49& \nodata& \nodata\\
300.030673&$ 20.034572$& 23.81& 4.23& 0.04& 16& 16& 70.42& 0.11& 0.69& 0.02& 0.6& 0.39& \nodata& \nodata\\
324.719526&$ 42.022609$& 4.34& 3.83& 0.05& 16& 16& 154.53& 0.16& 20.39& 0.17& \nodata& \nodata& \nodata& \nodata\\
325.523312&$ 42.023491$& 12.85& 2.83& 0.04& 16& 16& 356.39& 0.20& 30.28& 0.20& 1.0& 0.88& \nodata& \nodata\\
325.348791&$ 42.085546$& 7.61& 2.23& 0.04& 16& 16& 148.57& 0.15& 8.95& 0.12& 0.3& 1.42& \nodata& \nodata\\
326.519678&$ 42.141397$& 30.49& 8.09& 0.04& 16& 16& 54.95& 0.06& 2.98& 0.05& 0.5& 1.09& \nodata& \nodata\\
325.966278&$ 42.884463$& 7.55& 2.34& 0.05& 16& 16& 46.73& 0.15& 0.45& 0.02& \nodata& \nodata& \nodata& \nodata\\
326.391328&$ 43.460876$& 7.24& 4.02& 0.04& 16& 16& 46.97& 0.10& 0.32& 0.01& 1.8& 1.00& \nodata& \nodata\\
341.141889&$ 51.541980$& 15.60& 12.43& 0.05& 16& 16& 160.78& 0.10& 0.53& 0.03& 1.6& 0.54& 0.3& \nodata\\
342.654997&$ 52.100799$& 24.51& 13.77& 0.05& 16& 16& 195.40& 0.11& 18.74& 0.13& 1.8& 0.28& 0.6& 0.5\\
341.820302&$ 52.137525$& 4.11& 2.64& 0.04& 16& 16& 49.55& 0.12& 0.43& 0.03& 1.9& 0.66& 0.4& 1.1\\
343.612547&$ 52.339910$& 9.04& 7.77& 0.05& 16& 16& 168.64& 0.12& 8.37& 0.10& 1.2& -0.38& \nodata& \nodata\\
344.074222&$ 52.431561$& 4.50& 2.92& 0.04& 16& 16& 353.19& 0.28& 35.95& 0.24& \nodata& \nodata& 1.5& \nodata\\
342.837121&$ 52.495055$& 3.85& 3.53& 0.04& 16& 16& 102.16& 0.16& 4.92& 0.07& 5.6& 0.39& \nodata& \nodata\\
346.017848&$ 52.735617$& 10.19& 2.48& 0.04& 16& 16& 181.08& 0.22& 5.30& 0.10& 4.3& -1.12& \nodata& \nodata\\
345.504754&$ 52.912710$& 2.13& 1.69& 0.04& 16& 14& 91.42& 0.25& 1.59& 0.06& \nodata& \nodata& \nodata& \nodata\\
345.245407&$ 53.186497$& 33.34& 24.81& 0.05& 16& 16& 46.53& 0.06& 0.86& 0.02& 0.8& 0.91& \nodata& \nodata\\
346.421593&$ 53.202481$& 8.26& 3.30& 0.05& 16& 16& 71.60& 0.15& 0.18& 0.01& 6.4& 0.35& 0.1& \nodata\\
\hline
283.976980&$-12.165968$& 93.19& 57.18& 0.13& 15& 15& 468.40& 0.18& 68.27& 0.25& 0.9& 0.08& \nodata& \nodata\\
298.658519&$ 18.200598$& 3.28& 0.38& 0.04& 16& 0& 41.38& 0.55& 18.72& 0.68& \nodata& \nodata& 0.8& 0.3\\
299.358645&$ 18.324890$& 3.02& 0.87& 0.05& 16& 5& 26.51& 0.22& 23.36& 0.37& 2.0& 1.03& \nodata& \nodata\\
299.811372&$ 18.369894$& 1.82& 0.94& 0.04& 16& 6& 62.65& 0.29& 22.02& 0.34& \nodata& \nodata& \nodata& \nodata\\
323.729857&$ 41.317576$& 0.91& 0.55& 0.05& 16& 0& 28.52& 0.34& 20.48& 0.53& \nodata& \nodata& 1.6& 2.0\\
324.109846&$ 41.657383$& 1.55& 0.45& 0.05& 16& 0& 70.45& 0.64& 23.78& 0.67& \nodata& \nodata& \nodata& \nodata\\
324.719526&$ 42.022609$& 4.34& 3.83& 0.05& 16& 16& 154.53& 0.16& 20.39& 0.17& \nodata& \nodata& \nodata& \nodata\\
325.523312&$ 42.023491$& 12.85& 2.83& 0.04& 16& 16& 356.39& 0.20& 30.28& 0.20& 1.0& 0.88& \nodata& \nodata\\
342.654997&$ 52.100799$& 24.51& 13.77& 0.05& 16& 16& 195.40& 0.11& 18.74& 0.13& 1.8& 0.28& 0.6& 0.5\\
344.074222&$ 52.431561$& 4.50& 2.92& 0.04& 16& 16& 353.19& 0.28& 35.95& 0.24& \nodata& \nodata& 1.5& \nodata
\enddata
\tablenotetext{a}{The spectral power-law slope,
defined by $f_{\nu}\propto \nu^{\alpha}$, as measured from
the NVSS 1.4\,GHz specific flux and our 4.9\,GHz corrected flux density.}
\tablecomments{Catalog of the 464 sources detected in the Master images and their properties.
We note that {\it SAD} detected additional sources which are not listed here.
These sources were identified as noise artifacts by subsequent inspection
of the images and were removed from the catalog.
{\it Columns description:}
$f_{\nu,{\rm p}}^{{\rm cor}}$, $f_{\nu,{\rm p}}^{{\rm uncor}}$, and $\sigma_{{\rm p}}$
are described in Table~\ref{Tab:SingleEpochCat}.
$N_{{\rm obs}}$ is the number of epochs in which the source position was observed,
while $N_{{\rm det}}$ is the number of detections in the Single epoch images.
Subscript ``Y'' in $\chi^{2}$ and StD$/\langle f\rangle$ indicates that these
values are calculated for the Yearly epochs.
Dist is the distance between the radio position and the NVSS, USNO-B1 and 2MASS (Skrutskie et al. 2006)
nearest counterparts.
A portion of the full table, containing the 30 Single epoch variable sources
(\S\ref{VarSingle}) and the ten Yearly variable sources (\S\ref{VarYear}) are shown here for guidance regarding its form and content.
The two variable lists are separated by horizontal line.
This table is published in its entirety in the electronic edition of the {\it Astrophysical Journal}.
The electronic table also contains additional columns as the Field name,
sidelobes flag, integrated flux divided by peak flux, major axis, minor axis, and position angle of the sources,
distance from the beam center,
$V_{{\rm R}}$ and $V_{{\rm F}}$ for the yearly fluxes, the NVSS flux, the $B$ and $R$ magnitudes
of the USNO-B
counterparts, $J$, $H$ and $K$-magnitudes of the 2MASS counterparts
and all the 16 peak fluxes and errors measurements in the Single epochs and the peak flux
measurements in the Yearly images.
We note that the distance threshold for USNO-B and 2MASS counterparts was set to $2''$,
and $15''$ for NVSS.
\label{Tab:MasterCat}}
\end{deluxetable*}
\section{Post survey transient search}
\label{PostSurveyTran}
Our final transient search utilized the catalogs
presented in \S\ref{PostSurveyCat}.
Specifically, we matched all the sources in the individual epochs to
sources in the Master catalog using a $4''$ matching radius.
Sources in individual epochs which do not have counterpart
in the Master catalog are transient candidates.
However, because we used a single flux density threshold,
while the noise varies in each epoch and field,
most of the faint sources are probably noise artifacts.
We used Figure~\ref{Fig:Flux_Ndet} in
order to choose a reasonable flux density limit for our transient search.
This figure shows the number of detections, in the Single epoch catalogs,
of sources which are detected in the Master catalog and their field
was observed 16 times.
\begin{figure}
\centerline{\includegraphics[width=8.5cm]{f2.eps}}
\caption{Detection repeatability as a function of flux.
The plot shows, for each source
that its field was observed 16 times
and that was detected in the Master catalog,
the number of {\it SAD} detections in individual epochs.
The number of detections is shown against the uncorrected peak flux
of the source as measured in the Master image.
Point sources are represented by filled circles,
while open squares are for resolved sources.
The vertical line shows the 1.5\,mJy cut we used
to define our completeness limit.
\label{Fig:Flux_Ndet}}
\end{figure}
This plot suggests that the probability that a faint source
detected in the master image
will be detected with uncorrected flux density above 1.5\,mJy,
in only one epoch, is low.
In fact it suggests that we could use an even lower threshold.
However, given that this is based on small number statistics
and given the variable quality of different images,
we used an higher flux limit cutoff of 1.5\,mJy.
Following this analysis,
we searched for sources detected in a single epoch
that do~not have a counterpart in the Master catalog,
have uncorrected flux density $>1.5$\,mJy,
and a distance from beam center smaller than $4.65'$
(i.e., half power beam radius).
In total we found 50 sources. However, a close inspection
of these sources shows that most of them are not real.
Of the 50 candidates, 46 are in fields which contain sources brighter
than 10\,mJy, and they are clearly the results of sidelobes.
Of the remaining four candidates,
three sources are also most probably not real.
One candidate was found next to a slightly resolved
5\,mJy source.
A second object is a known 2\,mJy source for which
the centroid position puzzlingly shifted slightly in one of the epochs,
a third candidate is a sidelobe seen in several epochs,
and the fourth candidate is probably a real transient.
This transient candidate, J$213622.04+415920.3$
is described next.
\subsection{The transient candidate J$213622.04+415920.3$}
We found a single source, J$213622.04+415920.3$,
that was detected only in
the first epoch and may be a real transient.
Given that this source was detected in the first epoch,
before we constructed a reference image,
it was not followed up in real time.
The main properties of this transient candidate are summarized in
Table~\ref{Tab:TranProp}.
The peak flux measurements at the position
of the source at all epochs show that the source
was indeed visible only in the first epoch.
Moreover, this source is not detected
in the Master or Yearly images.
The peak flux at this position
in the Master image is $74\pm42$\,$\mu$Jy,
and in the 2008 (2010) combined images is $161\pm67$\,$\mu$Jy ($-14\pm49$\,$\mu$Jy).
\begin{deluxetable}{ll}
\tablecaption{Properties of the transient candidate J$213622.04+415920.3$}
\tablehead{
\colhead{Property} &
\colhead{Value}\\
}
\startdata
Field name & 2136$+$4158 \\
R.A. (J2000.0) & $21^{h}36^{m}22.^{s}04 \pm 1.2''$ \\
Dec. (J2000.0) & $+41^{\circ}59'20.''3 \pm 1.2''$ \\
Detection date & 2008 Jul 15.4147 \\
Uncorrected peak flux & $1.61\pm0.28$\,mJy \\
Corrected peak flux & $2.36\pm0.41$\,mJy \\
Distance from beam center & $2.77'$
\enddata
\tablecomments{We note that the coordinates in this table are based on Gaussian fit.
The coordinates of this source in the Single epoch Table (Table~\ref{Tab:SingleEpochCat})
were derived using SAD and they are somewhat different
RA$=21^{h}36^{m}21.^{s}965$, Dec$=+41^{\circ}59'21.''52$ (J2000.0).
Note that the flux is corrected also for the CLEAN bias.}
\label{Tab:TranProp}
\end{deluxetable}
In order to test if the source is variable during the 46\,s VLA integration,
we split the data into two 23-s images. We found that the flux of the source
did~not change significantly between the first and second parts of the exposure.
The field of J$213622.04+415920.3$ was observed with the P60
telescope in $i$-band about two days after the transient was detected.
This image is shown in Figure~\ref{Fig:VLA_J213621+415921}.
\begin{figure}
\centerline{\includegraphics[width=8.5cm]{f3.eps}}
\caption{P60 $i$-band image of the field of J$213622.04+415920.3$.
The image, with exposure time of 540\,s, was taken on 2008 Jul 17.45.
A $2.4''$ radius circle marks the transient candidate location.
This radius roughly corresponds to the $2$-$\sigma$ error circle.
\label{Fig:VLA_J213621+415921}}
\end{figure}
We find two sources near the transient position.
One source is found $\cong3.2''$ North-East of the transient location
and has an $i$-band magnitude of $20.98\pm0.23$.
The other is found $\cong2.3''$
South of the transient position and has a magnitude of $19.46\pm0.20$.
The magnitudes are calibrated relative to USNO-B1.0 I-band
magnitude (Monet et al. 2003).
Given the relatively large angular distances between the transient position
and the nearest visible light sources,
they are probably not associated.
We note that given the stellar density in this region
the probability to find a source within a $2.4''$ radius
from a random position is about $15\%$.
Finally, we do~not find any counterpart to this transient
in the SIMBAD, NED or HEASARC databases.
The search method described in the beginning of this section
may miss transients which are bright enough to be present
in the Master catalog.
Therefore, we also searched for sources which are detected in the Master catalog
and detected in only one of the Single epoch catalogs with uncorrected flux density
above 1.5\,mJy.
No such sources were found.
\section{Variability Analysis}
\label{Variability}
We investigated the variability of all the sources in the Master catalog
using their peak flux densities measured
in the Single epoch images (\S\ref{VarSingle}),
and the Yearly images (\S\ref{VarYear}).
We used various statistics to assess the
variability and its significance, which we list in Table~\ref{Tab:MasterCat}.
For each
source we calculated its standard deviation (StD) over all the
flux measurements,
the StD over the mean flux (StD/$\langle f\rangle$) and the $\chi^{2}$ given by
\begin{equation}
\chi^{2}=\sum_{i}^{N}{ \frac{(f_{i} - \langle f\rangle )^{2}}{\sigma_{i}^{2}+(f_{i}\epsilon_{{\rm cos},i})^{2}} },
\label{Eq:chi2}
\end{equation}
where $N$ is the number of measurements which is 2 for the Yearly
catalogs and 16 for the
Single epoch catalogs,
$i$ is the epoch index, $f_{i}$ is the gain corrected peak flux in the $i$-th epoch,
$\sigma_{i}$ is its associated error,
$\epsilon_{{\rm cos}}$ is the cosmic error,
and $\langle f\rangle$ is the mean flux of the source over all the epochs.
Note that in Table~\ref{Tab:MasterCat}, $\chi^{2}$ is measured over the individual epochs,
while $\chi^{2}_{Y}$ is measured over the two yearly epochs.
In the first case, the number of degrees of freedom ($dof$) is 15, while in
the second case it is one.
Some previous surveys have defined ``strong variables'' as
exceeding some pre-defined variability measure. There are many
definitions of fractional variability in the literature
and for comparison with other surveys
we list in the electronic version of
Table~\ref{Tab:MasterCat} the two following indicators
\begin{equation}
V_{{\rm R}} = \frac{\max\{f_{i}\}}{\min\{f_{i}\}},
\label{VR}
\end{equation}
and
\begin{equation}
V_{{\rm F}} = \frac{\max\{f_{i}\}-\min\{f_{i}\}}{\max\{f_{i}\}+\min\{f_{i}\}}.
\label{VF}
\end{equation}
We note that these two quantities are related through
$V_{{\rm F}}=(V_{{\rm R}}-1)/(V_{{\rm R}}+1)$.
However, as neither of these indicators account for
measurement errors,
they cannot be used reliably on their own at low
fluxes where the measurement errors are very large.
Most importantly, estimators involving the $\min$ or $\max$ functions
strongly depend on the
number of measurements.
This fact complicates direct comparison between different surveys.
For example, a light curve whose fluxes are drawn from
a log normal random distribution, with $N_{{\rm ep}}=3$ survey,
and StD/$\langle f\rangle=0.5$ has $\langle V_{R}\rangle\cong2.44$,
while for $N_{{\rm ep}}=16$ it will have $\langle V_{R}\rangle\cong5.62$.
Therefore, although
Becker et al. (2010) and Taylor and Gregory (1983)
defined strong variables identically, i.e. $V_{R}\geq $3 ($V_{F}\geq 0.5$),
any direct comparison of these two surveys is difficult since
the number of epochs in these surveys were 3 and 16, respectively.
In the future, in order to cope with this problem we suggest to use the StD/$\langle f\rangle$ as a fractional
variability estimator.
Moreover in Appendix~\ref{Ap:StatVR} we provide a conversion table
for $V_{R}$ as a function of $N_{{\rm ep}}$ and StD/$\langle f\rangle$.
\subsection{Short time scale variability}
\label{VarSingle}
In order to explore
radio variability on short time scales (e.g., days to weeks),
we constructed 16 epochs light curves for all sources
with peak flux densities larger than 1.5\,mJy ($\approx6\sigma$).
Figure~\ref{Fig:VarEp16_StDvsChi2} shows the
StD$/\langle f\rangle$ vs. the $\chi^{2}$ of all sources in
the Master catalog.
\begin{figure}
\centerline{\includegraphics[width=8.5cm]{f4.eps}}
\caption{
StD$/\langle f\rangle$ vs. the $\chi^{2}$, where the
uncorrected peak flux of the sources (in the Master catalog)
is marked by symbol size.
The dashed line corresponds to $\chi^{2}>45.5$, which corresponds to $4\sigma$ assuming 15 degrees of freedom.
\label{Fig:VarEp16_StDvsChi2}}
\end{figure}
This figure suggests that a large fraction of at least the bright radio sources
with flux densities larger than about 10\,mJy are variables at the level
of $\gtorder5$\%.
In total, we find that $30\%$ (30 out of 98) of the sources in our survey, which are
brighter than 1.5\,mJy, are variable
(at the 4-$\sigma$ level\footnote{Assuming Gaussian noise, $4\sigma$ corresponds to a probability of $\cong 1/15,000$ while the number of measurements in our experiment (number of epochs multiplied by the number of sources) is $\approx 7400$.}).
The light curves of these 30 variable sources are presented in
Figure~\ref{Fig:Var_LC}, and their flux measurements and basic properties
are listed in Table~\ref{Tab:MasterCat}.
This is considerably larger than the
fraction of variables reported in some of the other ``blind'' surveys
listed in Table~\ref{Tab:PreviousSurveys}
(e.g. Gregory \& Taylor 1986; de Vries et al. 2004; Becker et al. 2010).
A possible explanation for this apparent discrepancy is that the
sources in these surveys were extracted from mosaic images in which
each point in the survey footprint was observed multiple times
during several days. Therefore, the fluxes they reported are average
fluxes over several days time scales (column $\delta{t}$ in Table~\ref{Tab:PreviousSurveys}).
Such measurements will tend to
average out variability on time scales which are shorter than $\delta{t}$.
\begin{figure*}
\centerline{\includegraphics[width=16cm]{f5.eps}}
\caption{Light curves of the 30 variables which have $\chi^{2}>45.5$.
For scaling purposes we show only the 11 observations taken during 2008.
In each panel we give the J2000.0 Right Ascension and Declination of the source.
The individual flux measurements are given in the electronic version
of Table~\ref{Tab:MasterCat}.
\label{Fig:Var_LC}}
\end{figure*}
In order to test this hypothesis we carried out a structure function analysis.
We calculated the mean discrete auto-correlation function, $C(\tau)$,
of all the 30 variable sources, as a function of the time lag $\tau$.
We first normalized each source light curve by subtracting its mean and than
dividing it by its (original) mean.
We treated all these light curves as a single light curve
by concatenating them with gaps, which are larger than the time span
of each light curve, in between light curves.
Then we followed the prescription of Edelson \& Krolik (1988) for calculating
the discrete auto-correlation function.
The errors were calculated using a bootstrap technique with 100 realizations
for the measurements in each time lag
(e.g., Efron 1982; Efron \& Tibshirani 1993).
The mean auto-correlation function is presented in Figure~\ref{Fig:ACF_VarLC} (black circles).
The auto-correlation at lag ``zero'' is not one. This is because at lag zero
we used a lag window of 0 to $+2$\,days.
Therefore it does~not contain only zero lag data.
The auto-correlation function reaches zero correlation at $\tau\approx10$\,days.
\begin{figure}
\centerline{\includegraphics[width=8.5cm]{f6.eps}}
\caption{The mean discrete auto-correlation function of all the 30 variable sources (black circles).
The gray circles show the same but for all the non-variable sources
(see text).
Each light curve was normalized by subtracting its mean and dividing it by its mean.
The vertical error bars represent the $1$-$\sigma$ errors calculated
using the bootstrap method. The horizontal ``error bars'' represent the full width
of the lags contained within each bin.
The horizontal dotted line marks zero correlation.
\label{Fig:ACF_VarLC}}
\end{figure}
Next, we calculated the structure function, $SF(\tau)$,
of these light curves defined by
\begin{equation}
SF(\tau) = \sqrt{2S^{2}(1-C[\tau])} \pm \frac{2S^{2}\Delta{C[\tau]}}{2\sqrt{2S^{2}(1-C[\tau])}},
\label{Eq:SF}
\end{equation}
where $S$ is the standard deviation of the normalized light curves
and $\Delta{C[\tau]}$ is the bootstrap error in the discrete auto-correlation function.
The second term on the right hand side of Eq.~\ref{Eq:SF} represents
the error in the structure function.
The structure function is presented in Figure~\ref{Fig:SF_VarLC}.
Also shown in Figures~\ref{Fig:ACF_VarLC} and \ref{Fig:SF_VarLC} are
the auto-correlation and structure functions,
respectively, for
``non-variable'' sources (gray symbols).
In this context our selection criteria for non-variable sources
are specific flux larger than 2\,mJy and $\chi^{2}<25.3$.
This $\chi^{2}$ value corresponds to 2-$\sigma$ confidence, assuming 15 degrees of freedom.
The structure function of the variable sources, after subtracting the
non-variable source
structure function is rising rapidly
from zero to $\approx0.05$ on a time scale of the
order of one day and than rises to a level
of $\approx0.12$ at lags of $\approx10$\,days
at which it stays roughly constant.
Though, we cannot rule out it is slowly rising on $\tau>10$\,days time scales.
\begin{figure}
\centerline{\includegraphics[width=8.5cm]{f7.eps}}
\caption{The mean structure function of the normalized light curves
as a function of lag, $\tau$.
Symbols like in Figure~\ref{Fig:ACF_VarLC}.
Following Lovell et al. (2008) we fit the structure function
(after subtracting the non-variable sources structure function)
with the function $m_{{\rm s}}\tau/(\tau+\tau_{{\rm char}})$.
The heavy gray line represent the best fit function,
and the best fit parameters are $m_{{\rm s}}=0.119\pm0.003$,
$\tau_{{\rm char}}=0.83_{-0.23}^{+0.27}$\,day ($\chi^{2}/dof=26.6/6$).
\label{Fig:SF_VarLC}}
\end{figure}
This analysis suggests that a large component of the variability
happens on time scales shorter than about one day.
A plausible explanation is that the short time scale variability
is due to refractive scintillations and it is discussed in \S\ref{VarDisc}
(e.g., Rickett 1990).
This level of
variability was likely missed by some previous surveys (Table~\ref{Tab:PreviousSurveys})
due either to their choice of frequency ($\nu$), cadence (N$_{ep}$) or the
observing time span ($\delta{t}$). On the other hand, 5\,GHz surveys of
flat spectrum active galactic nuclei (AGN) find that their majority
shows significant variability, and that the
number of these variable sources increase at low Galactic latitudes
(e.g., Spangler et al. 1989;
Ghosh \& Rao 1992;
Gaensler \& Huntstead 2000;
Ofek \& Frail 2011).
The source population in the flux density range of our
survey is known to be dominated by AGN and a significant fraction
($\sim50$\%) of these are compact, flat-spectrum AGN and hence
expected to show short-term flux density variations (de Zotti et al. 2010).
\subsection{Variability on years time scale}
\label{VarYear}
By comparing our 2008 and 2010 catalogs we were able to
probe the variability of sources brighter than $\gtorder 0.5$\,mJy
on two year time scale.
However, this comparison is limited to only two epochs.
Each of the two epochs in our two year time-scale variability analysis
is composed of multiple epochs taken ($\approx\delta{t}$) days to weeks apart.
Therefore, in this analysis short-term variability
is averaged out.
For example, in 5\,GHz, variability due to scintillations will
typically have time scale which is shorter than a few days.
Figure~\ref{Fig:Flux2008vsFlux2010} shows
a comparison of the 2008 and 2010 peak flux densities for all
radio sources in the Master catalog.
The dashed lines represent the mean noise 4 and 8-$\sigma$ confidence
variability contours.
\begin{figure*}
\centerline{\includegraphics[width=16cm]{f8.eps}}
\caption{A comparison of the 2008 and 2010 peak flux densities for all
radio sources in the Master catalog. Equal fluxes are indicated by a
solid line.
The dashed lines are the
4$\sigma$ and 8$\sigma$ confidence levels contours
based on the mean noise properties
of the images and are calculated assuming the noise has a constant
component, which is the rms of the two yearly
images (Table~\ref{Tab:SurveyPar})
added in quadrature with a flux dependent component
which is given by the cosmic errors discussed in \S\ref{PostSurveyCat}.
The average rms noise
for the summed 2008 images was 72\,$\mu$Jy\,beam$^{-1}$ and for 2010
images it was 56\,$\mu$Jy\,beam$^{-1}$.
The shape and color coding represent the $\chi^{2}$ (one $dof$)
of individual objects
based on the actual noise rather than the average noise (see legend).
Open triangles mark limits were the direction of the limit
is given by the direction of the tip of the triangle (bottom or left).
\label{Fig:Flux2008vsFlux2010}}
\end{figure*}
Based on the $\chi^{2}_{Y}$ (Table~\ref{Tab:MasterCat})
we find ten
sources (2\% of all the sources) which have variability with confidence level larger than
about 4$\sigma$ (i.e., $\chi^{2}_{Y}>16$; assuming one degree of freedom).
These variable candidates are listed in Table~\ref{Tab:MasterCat} (below the horizontal line).
We note that the $\chi^{2}_{Y}$ calculation takes into account
the gain correction factors and the cosmic errors described in \S\ref{PostSurveyCat}.
We visually inspected the images of all ten variable candidates
and verified they are not sidelobe artifacts.
J$213626.36+413926.6$ is the only significant strong variable (i.e.
with $\chi^{2}/dof>16/1$ and $V_{{\rm F}}>0.5$) in our two-year comparison.
In 2010 it has a peak flux density (uncorrected for primary beam attenuation) of
$627\pm50$\,$\mu$Jy ($12.5\sigma$),
while in 2008 there is a nominal detection of
the source at this position with a peak flux density of $197\pm75$\,$\mu$Jy ($2.6\sigma$).
The source flux was well below the detection threshold
for every one of the 11 epochs in 2008 but it is visible in all five
epochs in 2010.
We note that this source is one of the two faint sources seen above
the dashed line on the right side of Figure~\ref{Fig:VarEp16_StDvsChi2}.
For this reason we do not classify this as a transient source
but it appears instead to be a persistent radio source that has
tripled its flux density over a two year interval. There is no known
cataloged source at this position in the NVSS catalog (Condon et al. 1998)
nor in the SIMBAD or NED databases.
Our Master catalog contains 317 sources brighter than J$213626.36+413926.6$.
Therefore, we roughly estimate that the fraction of strong variables
with two epochs $V_{{\rm F}}>0.5$ is $0.32_{-0.26}^{+0.73}\%$.
However, we note that this measurement is based on averaging out
variability on time scales shorter than a few weeks.
\section{Discussion}
\label{Disc}
We present a 16 epoch, Galactic plane survey, for
radio transients and variables at 5\,GHz.
We detected one possible transient and many variable
sources.
The transient areal density and rate based on this
single detection are derived in \S\ref{TranRate}.
In \S\ref{VarDisc} we discuss our variability
study and compare it with previous surveys.
\subsection{Transient areal density and rate}
\label{TranRate}
The analysis of our data revealed a single radio transient candidate.
The area encompassed within a single half power beam (radius of $r_{{\rm HP}}=4.65'$)
in which we searched for transients
is $0.00601$\,deg$^{2}$ and the area we targeted within the half power radius of
all 141 fields is $2.66$\,deg$^{2}$.
Given that each field was observed on average 15.70 times
(sum of column four in Table~\ref{Tab:ListOfFields}
divided by the number of fields),
the total area covered by our survey over all the epochs is $41.2$\,deg$^{2}$.
Our survey used an uncorrected flux limit of 1.5\,mJy for transients.
However, the sensitivity within the field of view of a single beam
imaging is not uniform, and degrades by a factor of two at the half power radius.
In order to calculate the transient areal density
from these parameters
we need to assume something about the source number count function.
We parametrize the source number count function as a power law
of the form
\begin{equation}
\kappa(>f)=\kappa_{0}(f/f_{0})^{-\alpha},
\label{Eq:CountFun}
\end{equation}
where $f$ is specific flux, $\kappa(>f)$ is the sky surface density of sources
brighter than $f$,
$\kappa_{0}$ is the sky surface density of sources brighter than $f_{0}$,
and $\alpha$ is the power law index of the source number count function.
It is well known that for homogeneous source distribution in
an Euclidean universe
and arbitrary luminosity function $\alpha=3/2$.
In Appendix~\ref{Ap:Rate} we derive a simple relation
for the number of sources that are expected
to be detected in a beam with
a power sensitivity that falls like a Gaussian as a function
of $f_{0}$, $\kappa_{0}$, $\alpha$, $r_{{\rm HP}}$,
the search radius $r_{{\rm max}}$, and the specific flux
limit at the beam center $f_{{\rm min},0}$.
Based on Equation~\ref{Eq:kappa},
and assuming $\alpha=3/2$, we find that
the transient areal density at 1.8\,mJy
(corrected for the CLEAN bias) is
\begin{equation}
\kappa(>1.8\,{\rm mJy}) = 0.039_{-0.032,-0.038}^{+0.13,+0.18}\,{\rm deg}^{-2},
\label{Eq:SurveyArealDen}
\end{equation}
where the errors correspond to one and two sigma confidence intervals,
calculated using the prescription of Gehrels (1986).
If our detected transient is not real then our
survey poses a 95\% confidence upper limit
on the transient rate of $\kappa(>1.8\,{\rm mJy})<0.15$\,deg$^{-2}$.
Translation of our areal density to transient rate
depends on the transient duration $t_{{\rm dur}}$ and it is
\begin{equation}
\Re(>1.8\,{\rm mJy}) = (28_{-23,-27}^{+65,+132}) \Big( \frac{t_{{\rm dur}}}{0.5\,{\rm day}} \Big)^{-1}\,{\rm deg}^{-2}\,{\rm yr}^{-1}.
\label{Eq:SurveyTranRate}
\end{equation}
Note that this translation is correct only if $t_{{\rm dur}}$ is smaller than
the time between epochs.
Figure~\ref{Fig:ArealDen_Flux_SurveySummary}
presents a summary of the radio transient and persistent source areal density
as observed by various searches and at different frequencies.
This figure is largely based on Table~\ref{Tab:PreviousSurveys}
and the areal density reported here.
As shown in this figure, the areal density derived in this
work is consistent with the expectation based on
the Bower et al. (2007) transient sky surface density.
Moreover, it is roughly consistent with the sky
surface density of the Nasu survey transients.
We note that the Nasu sky surface density
is based on our limited knowledge of this project
(see \S\ref{Prev}).
This comparison assumes that the transient areal density
on the celestial sphere is uniform.
This figure implies that the areal density of radio
transients in the sky is roughly 2--3 orders of magnitude below
the persistent radio source sky surface density.
This is in contrast to the visible light sky
in which the fraction of transients (excluding solar system minor planets)
among persistent sources is roughly $10^{-4}$ down
to the limiting magnitude of surveys like
the Palomar Transient Factory (Law et al. 2010; Rau et al. 2010).
It is interesting to compare this figure with some recent predictions.
Nakar \& Piran (2011) predict that compact binary mergers,
regardless whether they are associated with short-duration gamma-ray bursts
(e.g., Nakar 2007),
will produce radio afterglows with a duration of several months.
They suggest that the two-months long
radio transient RT19870422 detected by Bower et al. (2007)
may be a binary merger radio afterglow.
Moreover, they find that the rate inferred from this
event is consistent with the predicted rate
of binary merger events.
Giannios \& Metzger (2011) suggested that tidal flare events
may produce radio transients with durations of month to years,
with a 5\,GHz peak flux of 1\,mJy at a distance of 1\,Gpc (i.e., $z\cong0.2$).
The total comoving volume\footnote{Assuming WMAP fifth year cosmological
parameters (Komatsu et al. 2009).}
enclosed within a luminosity distance of 1\,Gpc is $2.4\times10^{9}$\,Mpc$^{3}$
or $5.9\times10^{4}$\,Mpc$^{3}$\,deg$^{-2}$.
Bower et al. (2007) did~not find any
transients with a duration of two months
which are associated with the nucleus of a galaxy.
This is translated to a 95\% confidence upper limit on
the rate of radio tidal flare events, brighter than 1\,mJy,
of $\sim0.1$\,deg$^{-2}$\,yr$^{-1}$.
Therefore, in the context of the Giannios \& Metzger (2011)
predictions,
we can put an upper limit of $\sim7\times10^{-6}$\,Mpc$^{3}$\,yr$^{-1}$
on the rate of tidal flare event radio afterglows.
This is in rough agreement with
the predicted tidal flares rate (Magorrian \& Tremaine 1999; Wang \& Merritt 2004).
Finally, we note that our detection rate is consistent with
an old NS origin as suggested in Ofek et al. (2010),
and with their expected surface density at low Galactic latitude
based on the Ofek (2009) simulations.
\subsection{Comparison of variability with previous surveys}
\label{VarDisc}
We analyzed the variability of the sources detected in our survey on
days--weeks time scales and two years period. On short time
scales we find that considerable fraction of the bright point sources
are variable. At the 10--100\,mJy range it seems that more than half
the sources are variable on some level ($>4\sigma$).
Furthermore, we find that
30\% of the sources in our survey, which are brighter than
$\approx1.5$\,mJy,
are variable (at the 4-$\sigma$ level). This is considerably higher
than the fraction of variables reported in some other surveys (e.g. Gregory
\& Taylor 1986; de Vries et al. 2004; Becker et al. 2010).
We suggest that a possible explanation for this apparent discrepancy is
that the sources in these surveys were extracted in a way that
washed out short time scale variability (see \S\ref{VarSingle}).
This is supported by the fact that our two years time scale
variability study, in which each epoch is composed by averaging
multiple observations, shows smaller fraction of variables.
Moreover, our structure function analysis shows that a big fraction of
the variability component happens on time scales shorter than about
a few days.
We note that these fast variation of radio sources
is known for a long time, and was found
by Heeschen (1982; 1984).
Moreover, a large fraction of variable sources was previously reported by
some other efforts, which did~not average out short time
scales variability (e.g., Lovell et al. 2008).
We speculate that the fast rise of the structure function on
$\approx10$\,day
time scales is due to scintillations in the interstellar
medium (ISM).
These time scales are consistent
with those expected theoretically from refractive scintillations in 5\,GHz
(e.g., Blandford et al. 1986; Hjellming \& Narayan 1986).
Moreover, similar rise times were reported by other efforts
(e.g., Qian et al. 1995).
Unlike diffractive scintillation which may
produce strong variability
(StD/$\langle f\rangle$ of $80\%$; e.g., Goodman 1997; Frail, Waxman \& Kulkarni 2000),
refractive scintillations can easily explain the observed
amplitude of $\approx13\%$.
We note that a comparison of models with observations
suggests that most of the radio source variability below
5\,GHz is due to scintillations, while above 5\,GHz
there is an intrinsic variability component
(e.g., Hughes, Aller \& Aller 1992; Mitchell et al. 1994; Qian et al. 1995)
Therefore, we can not rule out the possibility that some of
the variability we detected in our survey is intrinsic to the sources.
After averaging out variations on days time scale, our two year
variability analysis indicates that about $0.3\%$ of the sources above
0.5\,mJy are strong variables ($V_{{\rm F}}>0.5$ for $N_{{\rm ep}}=2$),
and that only a small
fraction $\approx3\%$ of the sources are variables at some level. This
finding supports the hypothesis that the main reason for low-amplitude
radio variability at 5\,GHz is due to scintillations.
It is well known that the fraction of radio variable sources
increases toward the Galactic plane
(e.g., Spangler et al. 1989; Ghosh \& Rao 1992;
Gaensler \& Hunstead 2000; Lovell et al. 2008; Becker et al. 2010; Ofek \& Frail 2011),
plausibly due to Galactic scintillations.
However, there are some
claims as yet unconfirmed that the number of
{\it intrinsically} strong variables, at 5\,GHz, increases toward the Galactic plane
(Becker et al. 2010).
Specifically, Becker et al. (2010) suggested that there is a
separate Galactic population
of {\it strong} variables.
As noted in \S\ref{Prev}, they found more
than half of their variables, on timescales of
years, varied by more than 50\% in the 1-100 mJy flux density range.
Moreover, they found that these strong variables were concentrated at low Galactic
latitudes and toward the inner Galaxy.
In contrast, we find a much smaller fraction
of strong variables (see \S\ref{VarYear}).
However, Becker et
al. (2010) observed sources within one degree of the Galactic plane,
while our survey sampled Galactic latitudes $\vert b\vert \cong6$--$8^\circ$.
\section{Summary}
\label{Sum}
We present a VLA 5\,GHz survey to search for radio transients
and explore radio variability in the Galactic plane.
Our survey represent the first attempt to discover radio transients in
near real time and initiate multi-wavelength followup.
Our real time search identified two possible transients.
However, followup observations and our post-survey analysis
showed that these candidates are not transient sources.
Nevertheless, in one case, we were able to initiate visible light observations
of the transient candidate field only one hour after the candidate
was detected.
Our post survey analysis reveals one possible transient source
detected at the 5.8-$\sigma$ level.
Our P60 images of this transient field, taken two days after
the transient detection, do~not reveal any visible light source
brighter than $i$-band magnitude of 21
associated with the transient within $2''$.
The transient has a time scale longer than 1\,min.
However, we cannot put an upper limit on its duration
since it was detected on the first epoch of our survey.
Based on this single detection we find a transients brighter than 1.8\,mJy areal density
of $\approx0.04$\,deg$^{-2}$.
The transient surface density found in this paper is compared with other surveys
in Figure~\ref{Fig:ArealDen_Flux_SurveySummary}.
Our transient areal density is consistent with the one reported
by Bower et al. (2007), corrected for the flux limit.
This is also roughly consistent, up to a possible spectral correction
factor, with the rates reported by Levinson et al. (2002) and Kida et al. (2007).
Finally, based on existing evidence, we cannot rule out
the hypothesis that these transients are
originating from Galactic isolated old NSs (Ofek et al. 2010).
Finally, we present a comprehensive variability analysis of our data,
with emphasis on proper calibration of the data and estimating
systematic noise.
Our findings suggest that short time scale variability
among 5\,GHz point sources is common.
In fact above 1.5\,mJy at least 30\% of the point sources
are variable with variability exceeding our 4-$\sigma$ detection level.
This is consistent with the Lovell et al. (2008) results,
and is plausibly explained by refractive scintillations in the ISM.
\acknowledgments
We are grateful to Barry Clark for his
help with scheduling the VLA observations.
EOO is supported by an Einstein fellowship and NASA grants.
Support for program number HST-GO-11104.01-A was provided by NASA through
a grant from the Space Telescope Science Institute, which is
operated by the Association of Universities for Research in
Astronomy, Incorporated, under NASA contract NAS5-26555.
AG acknowledges support
by the Israeli Science Foundation, an EU Seventh
Framework Programme Marie Curie IRG fellowship, the
Benoziyo Center for Astrophysics, and the Yeda-Sela fund
at the Weizmann Institute.
|
\section{Overview}
In recent years, tremendous progress has been made towards a more complete understanding of the scattering amplitudes in $\mathcal{N}=4$ super Yang-Mills theory~\cite{origN=4} (hereafter simply $\Nsym$). Lovingly referred to as the ``harmonic oscillator'' of quantum field theory, $\Nsym$ has more symmetry than any other gauge theory, especially in its so-called planar limit~\cite{origtHooft,origMald}. Although the theory's S-matrix has been under investigation for nearly 30 years~\cite{origGreenSchwBrink}, the last five have been particularly exciting. Numerous ground-breaking discoveries have been made (like the application of the AdS/CFT correspondence~\cite{origMald} to gluon scattering at strong coupling~\cite{origAldayMald}, a hidden dual superconformal symmetry of the planar theory~\cite{DHKSdualconf}, and a dual description of the leading singularities of the S-matrix as integrals over periods in a Grassmann manifold~\cite{dualSmat} to name just a few) and there is no reason to believe that we have learned everything $\Nsym$ has to teach us.
One of our main goals in this work is to further develop existing tools for the calculation of one-loop $\Nsym$ amplitudes to all orders in the dimensional regularization~\cite{origtHooftVelt} parameter. This parameter, $\e$, is introduced to cut off the IR divergences that appear in massless gauge theory calculations (we encourage readers less familiar with the structure of IR divergences in gauge theory to peruse Appendix \ref{dimregs}). We will illustrate our methods by considering examples where our results find useful application. At times we will develop aspects of $\Nsym$ S-matrix theory that appear to be of purely academic interest, but, in fact, a significant part of the computational machinery discussed in this paper can be applied to calculations in any quantum field theory. When techniques are applicable only in $\Nsym$ we will try to emphasize this. Before delving into the details of the problems we want to solve, a few words of historical introduction are in order.
$\mathcal{N}=4$ SYM is a very special four dimensional quantum field theory and its S-matrix has a number of unusual properties, many of which were unknown until very recently. We begin by reviewing some of its better-known features. The field content of the model consists of a gauge field $A_{\mu}$, four Majorana fermions $\psi_i$, three real scalars $X_p$, and three real pseudo-scalars $Y_q$. All fields are in the adjoint representation of a compact gauge group, $G$. The Lagrange density of $\mathcal{N}=4$ is given by \cite{myfirst}\footnote{Here we use the conventions of Peskin and Schroeder~\cite{PeskinSchroeder} for the Lagrange density, which differ somewhat from the conventions of~\cite{myfirst}. Throughout this paper, when not explicitly defined, the reader may assume that our conventions for perturbation theory coincide with those of Peskin and Schroeder.}
\bea
\label{LagDens}
\mathcal{L} &&= \textrm{tr} \bigg\{ -\frac{1}{2} F_{\mu \nu} F^{\mu \nu} + \bar{\psi_i} \slashed{D} \psi_i + D^{\mu} X_p D_{\mu} X_p + D^{\mu} Y_q D_{\mu} Y_q \\ \nonumber
&& - i g \bar{\psi_i} \alpha^p_{i j}[X_p,\psi_j]+g \bar{\psi}_i \gamma_5 \beta^q_{i j} [Y_q,\psi_j] \\ \nonumber
&& +\frac{g^2}{2}\bigg([X_l,X_k][X_l,X_k]+[Y_l,Y_k][Y_l,Y_k]+2[X_l,Y_k][X_l,Y_k]\bigg )\bigg\},
\eea
where the $4 \times 4$ matrices $\alpha^p$ and $\beta^q$ are given by \footnote{$\sigma_0$ is the $2 \times 2$ identity matrix.}
\bea
&& \alpha^1 = \left(\begin{array}{cc} i \sigma_2 & 0 \\ 0 & i \sigma_2 \end{array}\right),~~\alpha^2 = \left(\begin{array}{cc} 0 & - \sigma_1 \\ \sigma_1 & 0 \end{array}\right),~~\alpha^3 = \left(\begin{array}{cc} 0 & \sigma_3 \\ -\sigma_3 & 0 \end{array}\right), \\ \nonumber
&& \beta^1 = \left(\begin{array}{cc} -i \sigma_2 & 0 \\ 0 & i \sigma_2 \end{array}\right),~~\beta^2 = \left(\begin{array}{cc} 0 & -i \sigma_2 \\ -i \sigma_2 & 0 \end{array}\right),~~\beta^3 = \left(\begin{array}{cc} 0 & \sigma_0 \\ -\sigma_0 & 0 \end{array}\right).
\eea
Once the gauge group and coupling constant $g$ are fixed, the theory is uniquely specified. It turns out that in scattering amplitude calculations it is somewhat more typical to pair up the scalars and pseudoscalars and work with three complex scalar fields. The presence of four supercharges means that there is an $SU(4)$ R-symmetry acting on the fields. This symmetry acts on the state space as well and dictates selection rules for $\Nsym$ scattering amplitudes.
One of the first remarkable discoveries made about the $\Nsym$ model is that the
scale invariance of the classical Lagrange density (\ref{LagDens}) remains a symmetry at
the quantum level~\cite{SohniusWest}, implying that the $\beta$ function vanishes to
all orders
in perturbation theory. It follows~\cite{HoweStelleWest,HoweST,BrinkLN1,BrinkLN2,Lemes} that the
theory is UV finite in
perturbation theory (it turns out that the $\beta$ function remains
zero non-perturbatively
as well, but this is trickier to prove~\cite{Seiberg}). The classical superconformal
invariance
of the classical Lagrange density (\ref{LagDens}) (see Appendix B for a brief
discussion of
the $\Nsym$ superconformal group) continues to be a quantum mechanical
symmetry of
all correlation functions of gauge-invariant operators.
Most of the work on $\Nsym$ scattering amplitudes focuses on the massless, superconformal $\Nsym$ model described above but we note in passing that it is also possible to construct an $\Nsym$ model with both massive and massless fields~\cite{Fayet}. One can give some of the scalar fields in (\ref{LagDens}) vacuum expectation values (VEVs) at the cost of superconformal invariance and some of the generators of $G$. Formally, the fact that the six scalar fields can acquire VEVs without breaking supersymmetry implies that the theory has a six-dimensional moduli space of vacua. The $\Nsym$ model where some, but not all, of the gauge group generators are broken by scalar VEVs is called the Coulomb phase of the theory. While most of the literature has focused on the massless, conformal phase of the theory, the Coulomb phase is also quite interesting and is starting to attract the attention it deserves~\cite{myfirst,Higgsreg1,Higgsreg2,Higgsreg3,Higgsreg4}. Unfortunately, a proper discussion of the Coulomb phase is beyond the scope of this paper and we focus our attention exclusively on the S-matrix in the conformal phase of the theory, using dimensional regularization to regulate the IR divergences.
To better understand what makes $\Nsym$ so much simpler than garden-variety quantum field theories, it is instructive to compare the form of the one-loop virtual corrections in $\Nsym$ to, say, those in ordinary Yang-Mills theory. To be concrete, consider the four-gluon scattering amplitude in both models. Na\"{i}vely, one might think that the final results in the $\Nsym$ model are naturally expressed in terms of the Feynman integral basis for pure Yang-Mills, modulo UV divergent contributions. In fact, the basis of Feynman integrals that one needs for one-loop $\Nsym$ calculations form an even smaller subset. To understand this point, we must take a closer look at the integral basis for four-point scattering in pure Yang-Mills theory~\cite{PassarinoVeltman}, pictured in Figure \ref{4pttops}.
\FIGURE{
\resizebox{0.95\textwidth}{!}{\includegraphics{4ptints.eps}}
\caption{The possible integral topologies that could enter into the calculation of the one-loop four-gluon amplitude in pure Yang-Mills theory. The external particles are all taken to be outgoing and they are understood to be labeled clockwise beginning with the decorated leg.}
\label{4pttops}}
The integrals in each line of Figure \ref{4pttops} have a different topology. From top to bottom, we have scalar box integrals, triangle integrals, and bubble integrals. For instance, the 1st integral in the top row is defined as
\bea
-i(4 \pi)^{2-\e}\mintd{p}{4-2\e} {1\over p^2 (p-k_1)^2 (p-k_1-k_2)^2(p+k_4)^2}
\eea
and the third integral in the bottom row is defined as
\bea
-i(4 \pi)^{2-\e}\mintd{p}{4-2\e} {1\over p^2 (p-k_1-k_3)^2}
\eea
After a moment's thought it is clear that the $D = 4 - 2\e$ bubble integrals are UV divergent and the $D = 4 - 2\e$ triangle and box integrals are UV finite (let us stress that they are {\bf NOT} IR finite (see {\it e.g.} Appendix \ref{dimregs})).\footnote{It's worth pointing out that, in general, the basis integrals in spacetime dimensions other than four can appear as well. As explained in the \ref{GUD} the UV/IR behavior of such integrals is rather different.}
If we follow the above line of reasoning, we might guess that the one-loop four-gluon amplitude in $\Nsym$ is built out $D = 4 - 2\e$ triangles and boxes, but not $D = 4 - 2\e$ bubbles. Remarkably, this turns out not to be the case; the one-loop four-gluon $\Nsym$ amplitude is built out of box integrals only. What is even more remarkable is that, with the caveat that we drop contributions $\Ord(\e)$ and higher, this conclusion holds~\cite{1001lessons} for $n$-gluon scattering amplitudes\footnote{As we will be clear later, it is now known that this conclusion holds for one-loop $\Nsym$ scattering amplitudes with arbitrary external states.}.
For our purposes, the result in the above paragraph will not suffice; we are interested in studying $\Nsym$ amplitudes to all orders in $\e$ and we therefore need to modify the integral basis. Actually, this is not too hard. It has been clear at least since the work of~\cite{oneloopdimreg} that all one has to do is add scalar pentagon integrals to the basis. Then one can express any one-loop $\Nsym$ scattering amplitude in terms of pentagons and boxes to all orders in $\e$. These ideas will be explained in much more detail in Section \ref{revcomp} after the necessary framework has been developed.
Another main theme of this paper is a novel relation between one-loop scattering amplitudes in $\Nsym$ gauge theory and tree-level scattering amplitudes in open superstring theory\footnote{Tree-level amplitudes of massless particles in open superstring constructions compactified to four dimensions have a universal form~\cite{ST1}.}. With a bit of inspiration, the relationships to be discussed can be derived from the existing string theory literature. To the best of our knowledge, however, they are unknown at the time of this writing. We shall test our relationships explicitly in the simplest non-trivial case to establish confidence that they are correct.
What do we mean by ``the simplest non-trivial case?'' It turns out that there is a natural organizing principle for the S-matrix of $\Nsym$. If we label all external momenta as outgoing, as is conventional, then $\Nsym$ amplitudes can be organized according to a natural isomorphism between their little-group transformation properties and their $SU(4)_R$ transformation properties. In particular, we can assign a set of $SU(4)_R$ indices to each external state according to whether it is a positive helicity gluon, negative helicity fermion etc. Then there is a natural map~\cite{DHKSdualconf,SUSYBCFW,origElvang} from each external state of the theory to a subset\footnote{The map given in (\ref{operatormap}) is obviously not unique. Any consistent permutation of the flavor and $SU(4)_R$ labels for a given type of field would define an equally valid map.} of $\{1,2,3,4\}$. In what follows $g^\pm(p_i)$ is a positive or negative helicity gluon of momentum $p_i$, $\phi_a^{\pm}(p_i)$ is a positive or negative helicity fermion of flavor $a$ and momentum $p_i$, and $S^\pm_a(p_i)$ is a complex scalar of flavor $a$ and momentum $p_i$. A scalar has no helicity so the assignment of ``$+$'' and ``$-$'' is arbitrary (but useful) for scalar particles.
\bea
&&g^+(p_i)\leftrightarrow p_i
\nn
\phi_1^+(p_i) \leftrightarrow p_i^1 ~~~~~~~~~ \phi_2^+(p_i) &&\leftrightarrow p_i^2 ~~~~~~~~~ \phi_3^+(p_i) \leftrightarrow p_i^3 ~~~~~~~~~ \phi_4^+(p_i) \leftrightarrow p_i^4
\nn
S_1^+(p_i) \leftrightarrow p_i^{12} ~~~~~~~~~~~~~~ &&S_2^+(p_i) \leftrightarrow p_i^{23} ~~~~~~~~~~~~~ S_3^+(p_i) \leftrightarrow p_i^{13}
\nn
S_1^-(p_i) \leftrightarrow p_i^{34} ~~~~~~~~~~~~~~ &&S_2^-(p_i) \leftrightarrow p_i^{14} ~~~~~~~~~~~~~ S_3^-(p_i) \leftrightarrow p_i^{24}
\nn
\phi_1^-(p_i) \leftrightarrow p_i^{234} ~~~~~~~~~ \phi_2^-(p_i) &&\leftrightarrow p_i^{134} ~~~~~~~~~\phi_3^-(p_i) \leftrightarrow p_i^{124} ~~~~~~~~~ \phi_4^-(p_i) \leftrightarrow p_i^{123}
\nn
&& g^-(p_i) \leftrightarrow p_i^{1234}
\label{operatormap}
\eea
The only {\it a priori} non-zero scattering amplitudes are those that respect the R-symmetry; it must be possible to collect $k$ complete copies of $\{1,2,3,4\}$, where $k$ is a non-negative integer.
To make this rather abstract discussion more concrete, we consider examples for $k = 0,~1,~{\rm and}~2$. For $k = 0$ we have, for example, the all-positive helicity amplitude $\mathcal{A}\left(p_1,p_2,p_3,p_4\right)$. For $k = 1$ a good example is the four-positive helicity fermion amplitude $\mathcal{A}\left(p_1^1,p_2^2,p_3^3,p_4^4\right)$. Finally, an example for $k = 2$ is the four-point amplitude with a positive-negative helicity gluon pair and a positive-negative helicity fermion pair $\mathcal{A}\left(p_1,p_2^{1234},p_3^{1},p_4^{234}\right)$. In fact, it turns out that supersymmetry forces all scattering amplitudes with $k = 0~ {\rm or}~1$ to be equal to zero (see Appendix \ref{SWI} for a discussion of the supersymmetric Ward identities responsible for this). This implies that the first non-zero amplitudes have $k = 2$. Such amplitudes are called MHV amplitudes for historical reasons\footnote{MHV stands for maximally helicity violating. The $n$-point MHV amplitude describes, for example, a scattering experiment where two negative helicity gluons go in and $n-4$ positive helicity gluons and 2 negative helicity gluons come out. Such an outcome violates helicity as much as is possible at tree-level in QCD.}.
At the outset of the author's investigations, Stieberger and Taylor had recently discovered~\cite{ST1} a relation between the one-loop gluon $\Nsym$ MHV amplitudes and the tree-level gluon open superstring MHV amplitudes for which they had no explanation. Our work demystifies the relation they found and generalizes it as much as possible. Since Stieberger and Taylor showed explicitly that all MHV amplitudes satisfy the relation, it is of some interest to look at the simplest uncalculated example as an explicit test of our proposed generalization of the simpler Stieberger-Taylor relation. In other words, we ought to calculate the all-orders-in-$\e$ one-loop six-gluon\footnote{It is straightforward to check that at least six external particles need to participate in order to get an NMHV amplitude.} next-to-MHV (NMHV) amplitudes in $\Nsym$. Fortunately, Stieberger and Taylor have already tabulated all independent six-gluon NMHV amplitudes in open superstring theory~\cite{ST4} compactified to four dimensions. The existence of these results will make it significantly easier to check our proposed relations. Furthermore, our relations shed some light in a non-obvious way on an old result in pure Yang-Mills. In a nutshell, we are able to explain why $A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,\cdots,k_n)$ vanishes when $n > 4$ and three of the gluons are replaced by photons.
The precise statement of our relations between gauge and string theory is somewhat technical and we postpone further discussion of it to Section \ref{gsrel}. Suffice it to say that the gauge theory side of our relation requires one-loop $\Nsym$ amplitudes calculated to all orders in $\e$. At the outset of our investigations, it was not clear precisely what computational strategy was most appropriate. Therefore the entirety of Section \ref{gluoncomp} and part of Section \ref{supercomp} will be devoted to the explicit calculation of all-orders-in-$\e$ one-loop amplitudes in $\Nsym$.
Before we can discuss what is perhaps our nicest result for all-order one-loop $\Nsym$ amplitudes, we have to review several exciting recent developments in the theory of $\Nsym$ scattering amplitudes. For what follows the planar limit of the $\Nsym$ theory will be indispensable. The planar limit will be defined more carefully in Section \ref{revcomp}, but for now we discuss the simple example of four-gluon scattering to get across the main idea and to illustrate why taking this limit simplifies the S-matrix. We remind the reader that the building blocks for one-loop four-point scattering amplitudes in $\Nsym$ are the three box integrals in the top row of Figure \ref{4pttops}. Only the first of the three can be drawn in a plane without self-intersections. Operationally, the planar limit is reached if one computes the complete $\Nsym$ amplitude and then throws away all basis integrals that cannot be drawn in a plane without self-intersections. The reduction in the number of basis integrals (three to one) is modest at the one-loop four-point level. However, as you add more loops and legs, working in the planar limit dramatically reduces the complexity of the final results.
We now specialize to the planar limit and discuss some of the remarkable features of the $\Nsym$ S-matrix in this limit. Particularly exciting is the fact that, in the planar limit, it is possible to completely solve the perturbative S-matrix (up to momentum independent pieces) for the scattering of either four gluons or five gluons (and, by $\Nsym$ supersymmetry, all four- and five-point amplitudes). Starting from the work of~\cite{ABDK}, Bern, Dixon, and Smirnov (BDS) made an all-loop, all-multiplicity proposal for the finite part of the MHV amplitudes in $\Nsym$. In this paper~\cite{BDS}, BDS explicitly demonstrated that their ansatz was valid for the four-point amplitude through three loops. Subsequent work demonstrated that the BDS ansatz holds for the five-point amplitude through two-loops~\cite{TwoLoopFive} and that the strong coupling form of the four-point amplitude calculated via the AdS/CFT correspondence (the unfamiliar reader should consult Appendix \ref{ADS/CFT} for a brief description of this important result) has precisely the form predicted by BDS~\cite{origAldayMald}.
In fact,~\cite{origAldayMald} sparked a significant parallel development. Motivated by the fact that the strong coupling calculation proceeded by relating the four-point gluon amplitude to a particular four-sided light-like Wilson loop, the authors of~\cite{DKS4pt} were able to show that the finite part of the four-point light-like Wilson loop at one-loop matches the finite part of the planar one-loop four-gluon scattering amplitude. The focus of~\cite{DKS4pt} was on the planar four-gluon MHV amplitude, but it was shown in~\cite{BHTWLn} that this MHV amplitude/light-like Wilson loop correspondence holds for all one-loop MHV amplitudes in $\Nsym$. As will be made clear in Section \ref{WL/MHV}, an arbitrary $n$-gluon light-like Wilson loop should be conformally invariant in position space.\footnote{Strictly speaking, the conformal symmetry is anomalous due to the presence of divergences at the cusps in the Wilson loop. If one regulates these divergences and subtracts the conformal anomaly, then the finite part of what remains will be conformally invariant.} What was not at all obvious before the discovery of the amplitude/Wilson loop correspondence is that $\Nsym$ scattering amplitudes must be (dual) conformally invariant in {\it momentum} space.
It turns out that this hidden symmetry (referred to hereafter as dual conformal invariance) has non-trivial consequences for the $\Nsym$ S-matrix. Assuming that the MHV amplitude/light-like Wilson loop correspondence holds to all loop orders, the authors of~\cite{DHKSward} were able to prove that dual conformal invariance fixes the (non-perturbative) form of all the four- and five- point gluon helicity amplitudes (recall that non-MHV amplitudes first enter at the six-point level) in $\Nsym$. Up to trivial factors, they showed that the functional form of the (dual) conformal anomaly coincides with that of the BDS ansatz. Subsequently, work was done at strong coupling~\cite{BM,BRTW} that provides evidence for the assumption made in~\cite{DHKSward} that the MHV amplitude/light-like Wilson loop correspondence holds to all orders in perturbation theory. Quite recently, the symmetry responsible for the correspondence was understood from a perspective that bears on the results seen at weak coupling as well~\cite{DrummondFerro}.
The idea is that, due to the fact that non-trivial conformal cross-ratios can first be formed at the six-point level, one would na\"{i}vely expect the four- and five-point amplitudes to be momentum-independent constants to all orders in perturbation theory. It is well-known, however, that gluon loop amplitudes have IR divergences. These IR divergences explicitly break the dual conformal symmetry and it is precisely this breaking which allows four- and five- gluon loop amplitudes to have non-trivial momentum dependence. In fact, the arguments of~\cite{DHKSward} allowed the authors to predict the precise form that the answer should take and they found (up to trivial constants) complete agreement with the BDS ansatz to all orders in perturbation theory.
At this stage, it was unclear whether the appropriate hexagon Wilson loop would still be dual to the six-point MHV amplitude at the two-loop level. This question was decisively settled in the affirmative by the work of~\cite{BDKRSVV} on the scattering amplitude side and~\cite{DHKS2nd2L6pt} on the Wilson loop side. Another issue settled by the authors of~\cite{BDKRSVV} and~\cite{DHKS2nd2L6pt} was the question of whether the BDS ansatz fails at two loops and six points. It had already been pointed out by Alday and Maldacena in~\cite{AMlargen} that the BDS ansatz must fail to describe the analytic form of the $L$-loop $n$-gluon MHV amplitude for sufficiently large $L$ and $n$, but it had not yet been conclusively proven until the appearance of~\cite{BDKRSVV} and~\cite{DHKS2nd2L6pt} that $L = 2$ and $n = 6$ was the simplest possible example of BDS ansatz violation. The difference between the full answer and the BDS ansatz is called the remainder function and it is invariant under the dual conformal symmetry.
Since full two-loop six-point calculations are extremely arduous, one might hope that there is a smoother route to proving the above fact. In fact, Bartels, Lipatov, and Sabio Vera~\cite{BLSV1,BLSV2} derived an approximate formula for the imaginary part of the two-loop remainder function in a particular region of phase-space and multi-Regge kinematics. For some time, this formula was the subject of controversy, due to subtleties associated with analytical continuation of two-loop amplitudes. In~\cite{mysecond} the present author confirmed the controversial result\footnote{To appreciate the subtlety here the reader may wish to read the discussion at the end of~\cite{mysecond,BNST} as it relates to the erratum at the end of~\cite{DDG}.} of BLSV for the imaginary part of the remainder by explicitly continuing the full results of~\cite{DHKS2nd2L6pt} into the Minkowski region of phase-space in question.
The authors of~\cite{DHKS2nd2L6pt}, Drummond, Henn, Korchemsky, and Sokatchev (DHKS), recently discovered an even larger symmetry of the planar S-matrix. In~\cite{DHKSdualconf} DHKS found that there is actually a full dual $\Nsym$ superconformal symmetry acting in momentum space, which they appropriately christened dual superconformal symmetry. One of the main ideas utilized in ~\cite{DHKSdualconf} is that all of the scattering amplitudes with the same value of $n$ are unified into a bigger object called an on-shell $\Nsym$ superamplitude. This superamplitude can be further expanded into $k$-charge sectors and we will often refer to the $k$-charge sectors of a given superamplitude as superamplitudes as well. For example, the $n = 6$, $k = 2$ superamplitude would contain component amplitudes like $\mathcal{A}\left(p_1^{1234},p_2^{1234},p_3,p_4,p_5,p_6\right)$ and $\mathcal{A}\left(p_1^{1},p_2^{1234},p_3^{234},p_4,p_5,p_6\right)$ among others.
In~\cite{DHKSdualconf} DHKS made an intriguing conjecture for the ratio of the $k = 3$ and $k = 2$ six-point superamplitudes. They argued that the $k = 3$ superamplitude is naturally written as the $k = 2$ superamplitude times a function invariant under the dual superconformal symmetry. DHKS explicitly demonstrated that their proposal holds in the one-loop approximation. This ratio function has recently been the subject of intensive investigation and there are strong arguments in favor of it~\cite{BHMPYangian2}. Nevertheless, it would be nice to see explicitly that the two-loop ratio function is invariant under dual superconformal symmetry and this was done quite recently by Kosower, Roiban, and Vergu for an appropriately defined even part of the ratio function~\cite{KRV}. It turns out that the all-orders-in-$\e$ one-loop six-point $\Nsym$ NMHV superamplitude is necessary to explicitly test the dual superconformal invariance of the ratio function at two loops in dimensional regularization. In Section \ref{WL/MHV} we review the dual superconformal symmetry and explain how the all-orders-in-$\e$ one-loop formula we present in Section \ref{supercomp} can be rewritten to manifest this hidden symmetry as much as possible.
To summarize, the structure of this article is as follows. In Section \ref{revcomp} we review the modern computational techniques prerequisite to the topics discussed later in the paper. In Section \ref{gluoncomp} we discuss a new, efficient approach to the calculation of all-orders-in-$\e$ one-loop $\Nsym$ amplitudes, with the one-loop six-point gluon NMHV amplitude as our main non-trivial example. In Section \ref{gsrel} we discuss a novel relation between one-loop $\Nsym$ gauge theory and tree-level open superstring theory and illustrate its usefulness by solving an old puzzle in pure Yang-Mills. In Section \ref{supercomp} we discuss $\Nsym$ on-shell supersymmetry and extend our results for all-orders-in-$\e$ one-loop six-gluon NMHV amplitudes to the full one-loop $\Nsym$ NMHV superamplitude. In Section \ref{WL/MHV} we elaborate on the light-like Wilson loop/MHV amplitude correspondence, on dual superconformal invariance and on the relevance of the results of Section \ref{supercomp} to testing the dual superconformal invariance of the two-loop NMHV ratio function. Finally, in Section \ref{sum}, we summarize the main results of our work. In addition, we provide several appendices where we discuss important topics that deserve some attention but would be awkward to include in the main text. In Appendix \ref{dimregs} we discuss dimensional regularization, its usefulness in the regularization of IR divergences, and the structure of these divergences in planar $\Nsym$ gauge theory at the one-loop level. In Appendix \ref{sconf} we give a brief introduction to the $\Nsym$ superconformal group and present in considerable detail the $\Nsym$ superconformal and dual superconformal algebras. Appendix \ref{sconf} contains the complete conformal and dual superconformal algebras and corrects various misprints existing in the literature. In Appendix \ref{SWI} we give a brief introduction to the supersymmetric Ward identities, consequences of supersymmetry for the S-matrix that result in linear relations between many of the components of $\Nsym$ superamplitudes. Finally, in Appendix \ref{ADS/CFT}, we explain the AdS/CFT correspondence in general terms and then apply it to the calculation of the strong coupling form of the four-gluon amplitude in $\Nsym$.
\section{Review of Computational Technology}
\label{revcomp}
In this section we review some of the tools that make state-of-the-art gauge theory computations possible. In \ref{color} we discuss color decompositions, useful procedures that allow one to isolate the independent color structures that appear in the final results at the very beginning of a calculation. In \ref{planar} we define the planar limit of Yang-Mills theory. In \ref{SH} we introduce the spinor helicity formalism, a very convenient way of dealing with the external wave-functions of fermions and gauge bosons. In \ref{BCFW} we introduce the BCFW recursion relation and discuss its main applications. In \ref{GU4} we introduce the four dimensional generalized unitarity method at the one-loop level in the context of $\Nsym$ and discuss the integral basis, valid through $\Ord(\e^0)$, needed to use it. Finally, in \ref{GUD} we generalize the results of \ref{GU4} to $4 - 2\e$ dimensional spacetime ($D$ dimensional generalized unitarity).
\subsection{Color Decompositions}
\label{color}
In non-Abelian gauge theories one has to deal with the color, helicity, and kinematic degrees of freedom separately. Otherwise the resulting expressions for loop-level virtual corrections in scattering processes become much too complicated. In this work, we will only need to deal with color adjoints (all fields in $\Nsym$ live in the adjoint representation of the gauge group, which we take to be $SU(\Nc)$) and therefore restrict ourselves to discussing methods applicable to the situation where all of the external particles are in the adjoint representation. This should be contrasted to the situation in QCD. There one needs to deal with external particles that are color fundamentals as well (the fermions). Decoupling the color degrees of freedom in QCD is possible as well (see {\it e.g.}~\cite{BDKW} for an example at the one-loop level). The resulting decomposition, however, is much less simple.
To begin, we illustrate the concept of color decomposition by analyzing a contribution to four-gluon scattering at tree-level. In what follows, we deviate from the standard normalization (described in textbooks like~\cite{PeskinSchroeder}) and replace
$${\rm Tr}\{T^a T^b\} = {\D^{a b} \over 2}$$
with
$${\rm Tr}\{T^a T^b\} = \D^{a b} ~{\rm .}$$
If this alternative normalization convention is not adopted the color decomposition results in objects that have an annoying $2^n$ out front for an $n$-point scattering process (see {\it e.g.}~\cite{Yasui} for some sample calculations with the standard conventions). We will see that the usual Feynman rules can be split up into simpler rules that only contribute to specific color structures. As a first example, we consider the $s$-channel Feynman diagram drawn in Figure \ref{4gluecolor}. For this diagram we need only the gluon propagator and the three-gluon vertex and we work in 't Hooft-Feynman gauge using the conventions of~\cite{PeskinSchroeder}.
\FIGURE{
\resizebox{0.75\textwidth}{!}{\includegraphics{4gluesch.eps}}
\caption{The $s$-channel diagram contributing to the tree-level four-gluon amplitude.}
\label{4gluecolor}}
\vspace{.75 in}
Denote the graph of Figure \ref{4gluecolor} by $\mathcal{A}_S$. We then have
\bea
\a_S &=& \left(g f^{a_1 a_2 b}\big[\pol^{h_1}(k_1)\cdot \pol^{h_2}(k_2)(k_1-k_2)_\mu+\pol_\mu^{h_2}(k_2)\pol^{h_1}(k_1)\cdot (2 k_2+k_1)+\pol_\mu^{h_1}(k_1)\pol^{h_2}(k_2)\cdot(-2 k_1-k_2)\big]\right)\times
\el
\times \left({-i g^{\mu \nu} \delta^{b c} \over (k_1 + k_2)^2}\right) \times
\el
\times \left(g f^{a_3 a_4 c}\big[\pol^{h_3}(k_3)\cdot \pol^{h_4}(k_4)(k_3-k_4)_\nu+\pol_\nu^{h_4}(k_4)\pol^{h_3}(k_3)\cdot (2 k_4+k_3)+\pol_\nu^{h_3}(k_3)\pol^{h_4}(k_4)\cdot(-2 k_3-k_4)\big]\right) \,{\rm ,}\nn
\label{raw4glue}
\eea
where we have kept the polarizations of the gluons arbitrary. For now we ignore the dependence of $\a_S$ on everything except color. This leaves us with
\be
f^{a_1 a_2 b}f^{a_3 a_4 b}
\label{initcolorcont}
\ee
To push further, we exploit a few simple facts about the Lie algebra of $SU(\Nc)$. The structure constants of the group can be written as
\be
f^{a b c} = -{i \over \sqrt{2}} {\rm Tr}\{T^{a}T^{b}T^{c}-T^{a}T^{c}T^{b}\}
\ee
and the structure constants with one index contracted up with a fundamental representation generator matrix can be rewritten by using the algebra itself:
\be
- {i \over \sqrt{2}}[T^a,~T^b] = f^{a b c}T^c \,{\rm .}
\ee
These relations allow us to write (\ref{initcolorcont}) completely in terms of the $T^a$. After massaging the color factors into the desired form
\bea
f^{a_1 a_2 b}f^{a_3 a_4 b} &=& -{i \over \sqrt{2}} {\rm Tr}\{T^{a_1}T^{a_2}T^{b}-T^{a_1}T^{b}T^{a_2}\} f^{a_3 a_4 b}
\el
= -{1\over 2}{\rm Tr}\{T^{a_1}T^{a_2}[T^{a_3},T^{a_4}]-T^{a_1}[T^{a_3},T^{a_4}]T^{a_2}\}
\elale
-{1\over 2}{\rm Tr}\{T^{a_1}T^{a_2}T^{a_3}T^{a_4}-T^{a_1}T^{a_2}T^{a_4}T^{a_3}-T^{a_1}T^{a_3}T^{a_4}T^{a_2}+T^{a_1}T^{a_4}T^{a_3}T^{a_2}\} \nn
\eea
we can now go back to eq. (\ref{raw4glue})
\bea
\a_S &=& i g^2 {\rm Tr}\{T^{a_1}T^{a_2}T^{a_3}T^{a_4}-T^{a_1}T^{a_2}T^{a_4}T^{a_3}-T^{a_1}T^{a_3}T^{a_4}T^{a_2}+T^{a_1}T^{a_4}T^{a_3}T^{a_2}\} \times
\el
\times {1\over 2}\big[\pol^{h_1}(k_1)\cdot \pol^{h_2}(k_2)(k_1-k_2)_\mu+\pol_\mu^{h_2}(k_2)\pol^{h_1}(k_1)\cdot (2 k_2+k_1)+\pol_\mu^{h_1}(k_1)\pol^{h_2}(k_2)\cdot(-2 k_1-k_2)\big]
\times
\el
\times {g^{\mu \nu} \over (k_1 + k_2)^2} \big[\pol^{h_3}(k_3)\cdot \pol^{h_4}(k_4)(k_3-k_4)_\nu+\pol_\nu^{h_4}(k_4)\pol^{h_3}(k_3)\cdot (2 k_4+k_3)+\pol_\nu^{h_3}(k_3)\pol^{h_4}(k_4)\cdot(-2 k_3-k_4)\big]\nn
\label{fincoldemo}
\eea
and we see explicitly that we've achieved our goal. This amplitude contributes to four different color structures and four different {\it color-ordered partial amplitudes}. A color-ordered partial amplitude is defined as the collection of all terms from all the different diagrams that have the same color structure. The reason that this decomposition is so useful in practice is that the color structures are independent and therefore, by construction, each color-ordered partial amplitude must be gauge invariant. Since a color structure is only defined up to cyclic permutations (because the trace is cyclicly symmetric), we choose representatives for them with the convention that $T^{a_1}$ is always the first generator matrix to appear in any trace structure. To make sure that the notation is clear, we remind the reader that $a_1$ denotes the color label, valued in $\{1,\cdots,\Nc\}$, of the first gluon. However, due to cyclic symmetry, what we decide to call the first gluon is completely arbitrary. Finally, we note that this technique can easily be used to expand the color factors of the four-gluon vertex as well.
Although the $n$-point generalization of the above tree-level color decomposition is typically explained in a field theory context, for us it is useful to follow the route that was taken historically. In~\cite{Manganocolor,BerendsGiele88} it was pointed out that an elegant way to derive color decomposition formulae in field theory is to first derive the formulae in an open superstring theory and then take a particular limit (the infinite string tension limit) in which the unwanted modes in the superstring theory decouple and $\Nsym$ gauge theory falls out. This approach benefits us because we will need the tree-level color decomposition for open superstring theory in Section \ref{gsrel} when we discuss our non-trivial relations between one-loop $\Nsym$ amplitudes and tree-level open superstring amplitudes.
It has been known since the early days of superstring theory, that one can write a open superstring theory amplitude as
\bea
&& \a^{tree}_{str}\left(k_1^{h_1},~k_2^{h_2},~\cdots,~k_n^{h_n}\right) =
\el
g^{n-2} \sum_{\sigma \in S_n/\mathcal{Z}_n} {\rm Tr}\{T^{a_{\sigma(1)}}T^{a_{\sigma(2)}}~\cdots T^{a_{\sigma(n)}} \} A^{tree}_{str}\left(k_{\sigma(1)}^{h_{\sigma(1)}},~k_{\sigma(2)}^{h_{\sigma(2)}},~\cdots,~k_{\sigma(n)}^{h_{\sigma(n)}}\right)
\elale
g^{n-2} \sum_{\sigma \in S_n/\mathcal{Z}_n} {\rm Tr}\{T^{a_{\sigma(1)}}T^{a_{\sigma(2)}}~\cdots T^{a_{\sigma(n)}} \} A^{tree}\left(k_{\sigma(1)}^{h_{\sigma(1)}},~k_{\sigma(2)}^{h_{\sigma(2)}},~\cdots,~k_{\sigma(n)}^{h_{\sigma(n)}}\right) + \Ord(\alpha'^2)\,{\rm ,}\nn
\label{fintreecols}
\eea
where partial amplitudes are associated to appropriate Chan-Paton factors and one sums over all distinguishable permutations\footnote{String amplitudes can be expressed as a power series in the inverse string tension, $\alpha'$, and (\ref{fintreecols}) is the zeroth order term. More discussion of the perturbation theory of the massless modes in open superstring is given in \ref{gsrel}.}. In the infinite string tension limit, the formula above reduces to an analogous one for $U(\Nc)$ $\Nsym$ gauge theory:
\bea
&& \a^{tree}\left(k_1^{h_1},~k_2^{h_2},~\cdots,~k_n^{h_n}\right) =
\el
g^{n-2} \sum_{\sigma \in S_n/\mathcal{Z}_n} {\rm Tr}\{T^{a_{\sigma(1)}}T^{a_{\sigma(2)}}~\cdots T^{a_{\sigma(n)}} \} A^{tree}\left(k_{\sigma(1)}^{h_{\sigma(1)}},~k_{\sigma(2)}^{h_{\sigma(2)}},~\cdots,~k_{\sigma(n)}^{h_{\sigma(n)}}\right) \,{\rm .}\nn
\label{fintreecol}
\eea
At this stage, the alert reader may be wondering whether the $U(\Nc)$ written above should really be $SU(\Nc)$. Actually, this is not the case. Locally, one can always write $U(\Nc) \simeq SU(\Nc)\times U(1)$ and, since $U(1)$ is Abelian, we are effectively in $SU(\Nc)$ because any partial amplitude containing the $U(1)$ particle (photon) must vanish. This cancellation is simply due to the fact that such a particle has no way to couple to states that live in the adjoint of $SU(\Nc)$. In fact, this vanishing yields the following identity
\be
0 = A^{tree}\left(k_1^{h_1},~k_2^{h_2},~\cdots,~k_n^{h_n}\right)+ A^{tree}\left(k_2^{h_2},~k_1^{h_1},~\cdots,~k_n^{h_n}\right)+\cdots+ A^{tree}\left(k_2^{h_2},~k_3^{h_3},~\cdots,~k_1^{h_1}\right)
\label{photdecoup}
\ee
for the partial amplitudes, commonly referred to as the photon decoupling identity. Eq. (\ref{photdecoup}) is nothing but the expression for an (vanishing) amplitude with a photon with momentum $k_1$ and helicity $h_1$ and $n-1$ $SU(\Nc)$ adjoint particles. The reason that there is a sum with $k_1^{h_1}$ inserted in all possible positions is that the photon is not ordered with respect to the $n-1$ adjoint particles and therefore we have to symmetrize with respect to the insertion of the photon. Finally, we remark that there is another feature of (\ref{fintreecol}) that should bother the alert reader: it looks like there are $(n-1)!$ independent partial amplitudes that need to be computed to determine the full amplitude. In practice, the situation is much better; recently, the authors of~\cite{BCJ} showed that, in fact, there are really only $(n-3)!$ independent partial amplitudes.
To summarize, the claim is that, at tree-level, $\Nsym$ amplitudes\footnote{All that we have really assumed is that our gauge group is $SU(\Nc)$ and that all of the fields transform in the adjoint representation.} organized in terms of their color structures look identical to open superstring amplitudes organized in terms of their Chan-Paton~\cite{ChanPaton} factors. It is also very important to note that all we had to assume about the superstring construction to get this to work at tree-level is that it approaches $\Nsym$ field theory in a particular limit. In a nutshell, this is why the tree-level S-matrix of massless modes in superstring theory is model independent. This model independence disappears at loop level and one has to work much harder. Nevertheless an analogous program can be carried out at one-loop~\cite{BKcolor} and it has been shown that one can use string constructions to derive useful one-loop color decomposition formulae in field theory. It is to this topic that we now turn.
Unfortunately, there are many more color structures at one-loop than at tree-level and this complicates things. Since a description of the necessary heterotic string construction would take us much too far afield, we refer the interested reader to~\cite{BKcolor} and simply state the main result of their work. In $\Nsym$, one-loop scattering amplitudes can be decomposed into color-ordered partial amplitudes using the formula
\begin{changemargin}{-.4 in}{0 in}
\bea
&& \a^{1-{\rm loop}}_{\Nsym}\left(k_1^{h_1},~\cdots,~k_n^{h_n}\right) =
g^{n-2}{g^2 \Nc \mu^{2\e} e^{-\gamma_E \e}\over (4 \pi)^{(2-\e)}} \sum_{\sigma \in S_n/\mathcal{Z}_n} {\rm Tr}\{T^{a_{\sigma(1)}}~\cdots T^{a_{\sigma(n)}} \} \times
\el
\times A^{1-{\rm loop}}_{1;\,\Nsym}\left(k_{\sigma(1)}^{h_{\sigma(1)}},~\cdots,~k_{\sigma(n)}^{h_{\sigma(n)}}\right)+ g^{n-2}{g^2 \mu^{2\e} e^{-\gamma_E \e} \over (4 \pi)^{2-\e}} \sum_{m = 2}^{[{n \over 2}]+1}\Bigg( \sum_{\sigma \in~ S_n/(\mathcal{Z}_{m-1}\times\mathcal{Z}_{n-m+1})} {\rm Tr}\{T^{a_{\sigma(1)}}~\cdots T^{a_{\sigma(m-1)}}\}\times
\el \times{\rm Tr}\{T^{a_{\sigma(m)}}~\cdots T^{a_{\sigma(n)}} \} A^{1-{\rm loop}}_{2;\,\Nsym}\left(k_{\sigma(1)}^{h_{\sigma(1)}},~\cdots,~k_{\sigma(n)}^{h_{\sigma(n)}}\right)\Bigg)\,{\rm ,}\nn
\label{finloopcol}
\eea
\end{changemargin}
where
\be
\Big[{n \over 2}\Big] \equiv {\rm Floor}\left({n \over 2}\right)\,{\rm .}
\ee
In this work we will very often use the notation $A^{1-{\rm loop}}_{1}$ as a somewhat abbreviated version of $A^{1-{\rm loop}}_{1;\,\Nsym}$. The appearance of $e^{-\gamma_E \e}/(4 \pi)^{2-\e}$ in (\ref{finloopcol}) is necessary for technical reasons and will be explained in Subsection \ref{GU4}. The factor $\mu^{2\e}$, the unit of mass, is explained in Appendix \ref{dimregs} and is important in theories where the interesting observables are infrared finite. Also, in contrast to what happened at tree-level, we have both single-trace color structures and double-trace color structures. Actually, as we will see in the next Subsection, the double-trace structures will be irrelevant for us because they are sub-leading in the number of colors, $\Nc$. In any case, at one-loop, there are analogs of the photon decoupling relations at tree-level and these allow one to express the coefficients of the double-trace color structures in terms of the coefficients of the single-trace color structures~\cite{BKcolor}. The coefficients of the single-trace structures are commonly referred to as the leading color-ordered partial amplitudes, again because of their dominance at large $\Nc$.
\subsection{Planar Limit}
\label{planar}
Long ago, 't Hooft observed that non-Abelian gauge theories simplify dramatically~\cite{origtHooft} in a particular limit, in which one fixes the combination $\lambda = 2 \Nc ~g^2$, eliminates $g$ in favor of $\Nc$ and $\lambda$, and then takes $\Nc$ to infinity ($\lambda$ is referred to as the 't Hooft coupling in his honor). One thing 't Hooft conjectured was that large $\Nc$ gauge theory ought to have a stringy description. This idea was given new life by Maldacena in his ground-breaking work~\cite{origMald} on the near-horizon geometry of $AdS_5 \times S_5$ (see Appendix \ref{ADS/CFT}). In brief, Maldacena showed that type IIB superstring theory in an $AdS_5 \times S_5$ background is dual to a $\Nsym$ SYM gauge theory. Maldacena's duality was incredibly novel because it related planar $\Nsym$ at strong coupling to classical type IIB superstring theory at weak coupling. In this paper, we will see that unexpected simplicity also emerges in the planar limit of {\it weakly} coupled $\Nsym$.
For our purposes, the advantage of working in the planar limit is that the simple tree-level color decomposition formula of eq. (\ref{fintreecol}) actually generalizes to multi-loop amplitudes. This is not hard to guess at the one-loop level from the decomposition formula (\ref{finloopcol}). In this formula, the single-trace color structures have an explicit factor of $\Nc$ out front that the double-trace structures do not. It follows that the single-trace structures dominate in the large $\Nc$ limit. To be explicit, the planar $L$-loop color decomposition formula is
\bea
&&{\cal A}_1^{L-{\rm loop}}\left(k_1^{h_{\sigma(1)}}, k_2^{h_{\sigma(2)}}, ~\ldots, k_n^{h_{\sigma(n)}}\right) =
\el
g^{n-2}\left({g^2 \Nc \mu^{2\e} e^{-\gamma_E \e}\over (4 \pi)^{2-\e}}\right)^L
\sum_{\sigma \in~ S_n/Z_n}
\, \Tr\{T^{a_{\sigma(1)}} T^{a_{\sigma(2)}}
\ldots T^{a_{\sigma(n)}}\}
A_{1}^{L-{\rm loop}}\left(k_1^{h_{\sigma(1)}}, k_2^{h_{\sigma(2)}}, ~\ldots, k_n^{h_{\sigma(n)}}\right)\,{\rm .} \nn
\label{LLoopDecomposition}
\eea
Clearly, this is going to be much easier to work with than a full $L$-loop color decomposition.
Although $\Nsym$ supersymmetry by itself is very powerful and puts highly non-trivial constraints on the perturbative S-matrix, $\Nsym$ supersymmetry together with the planar limit is even more powerful. In section \ref{WL/MHV} we will discuss a so-called hidden symmetry of the $\Nsym$ S-matrix that emerges in the large $\lambda$ limit. This symmetry, called dual superconformal invariance is like a copy of the ordinary superconformal invariance of the $\Nsym$ that acts in {\it momentum} space\footnote{The Lagrange density of $\Nsym$ is manifestly superconformally invariant in position space. We encourage the reader unfamiliar with superconformal symmetry to peruse Appendix \ref{sconf}.}.
\subsection{Spinor Helicity Formalism}
\label{SH}
In this Subsection we review the spinor helicity formalism, a computational framework suitable for research-level helicity amplitude computations in non-Abelian gauge theories. Although the formalism has been fully developed for more than twenty years~\cite{Calkul1,Calkul2,Calkul3,Calkul4,Calkul5,Calkul6}, it has only recently begun to replace the traditional techniques in mainstream textbooks (see {\it e.g.}~\cite{Srednicki}). The underlying assumption of the method is that the correct way to deal with the S-matrix is to compute helicity amplitudes. This is a useful approach in practice because many helicity amplitudes are protected by supersymmetry or related by discrete symmetries (entering either from parity invariance or the color decomposition).
The spinor helicity formalism is designed to streamline the computation of helicity amplitudes and to allow one to express the results obtained in as simple a form as possible. In the spinor helicity formalism, {\it everything} that enters into the computation of a helicity amplitude is expressed in terms of the same set of building blocks. For example, all dot products involving the polarization vectors of the gluons (or photons) in the scattering process under consideration are expressed in terms of spinor products. This is also true for invariants built out of scalar products of external four-momenta. Another nice feature of the method is directly related to the way in which external polarization vectors are dealt with in the spinor helicity framework. In a moment, we will see that on-shell gauge invariance is built into the spinor helicity polarization vectors and this can be exploited to cancel numerous terms at the beginning of the calculation. In this subsection we closely follow the conventions and discussion of~\cite{Dixon96rev}, though we will ultimately rewrite everything in more modern notation.
Consider free particle solutions to the massless Dirac equation. The positive and negative energy solutions, $u^\pm(p)$ and $v^\mp(p)$ are equivalent up to phase conventions~\cite{Srednicki}. In particular, it is possible to make a choice where $v^\mp(p) = u^\pm(p)$ and we can eliminate $v^\mp(p)$ in favor of $u^\pm(p)$. In this language, a momentum invariant, $s_{ij} \equiv (p_i+p_j)^2$, is easily expressed in terms of spinor products as
\be
s_{i j} = \bar{u}^-(p_i)u^+(p_j)\bar{u}^+(p_j)u^-(p_i) \,{\rm .}
\label{sijsh}
\ee
The above identity can be verified by turning the right-hand side of (\ref{sijsh})into a trace over gamma matrices and projection operators (see {\it e.g.}~\cite{PeskinSchroeder} if unfamiliar). The polarization vectors $\e^+(p_i)$ and $\e^-(p_i)$ are expressed as
\bea
&&\pol^+_\mu(p_i) = {\bar{u}^-(q_i)\gamma_\mu u^-(p_i)\over \sqrt{2}~\bar{u}^-(q_i)u^+(p_i)}
\label{polvecs+} \\
&&
\pol^-_\mu(p_i) = {\bar{u}^-(p_i)\gamma_\mu u^-(q_i)\over \sqrt{2}~\bar{u}^+(p_i)u^-(q_i)}\,{\rm ,}
\label{polvecs-}
\eea
where the four-momentum $q_i$ is called the reference momentum associated to $p_i$. Though it is probably not clear how we arrived at eqs. (\ref{polvecs+}) and (\ref{polvecs-}) (see~\cite{TUPhoton,XZC} and, more recently,~\cite{BFLWard} for details), it should be plausible that one can build a $\bf (1/2, 1/2)$ wavefunction out of a $\bf (1/2, 0)$ wavefunction and a $\bf (0, 1/2)$ wavefunction\footnote{Here we are using the representation theory of $SU(2) \times SU(2)$ to label the states of definite helicity. This is possible because the complexified proper orthochronous Lorentz group and $SU(2) \times SU(2)$ have isomorphic Lie algebras. The complexification is necessary to obtain the desired isomorphism between the Lie algebras but doesn't affect the labeling of representations. See {\it e.g.}~\cite{WiedemannKirstenMuller} for more details.}. In fact, using the massless Dirac equation, it is straightforward to show that eqs. (\ref{polvecs+}) and (\ref{polvecs-}) are solutions of the Fourier transformed Yang-Mills field equations in the appropriate limit. Furthermore, they are normalized in the standard way
\be
\pol^{+}(p_i)\cdot\left(\pol^+(p_i)\right)^* = \pol^{+}(p_i)\cdot \pol^-(p_i) = \left(\pol^{-}(p_i)\right)^*\cdot\pol^-(p_i) = -1
\ee
and satisfy the completeness relation appropriate for a light-like axial gauge~\cite{Dixon96rev}:
\be
\sum_{\lambda = \pm} \pol^{\lambda}_\mu(p_i, q_i) \left(\pol^{\lambda}_\nu(p_i, q_i)\right)^* = -g_{\mu \nu} + {p_{i\,\mu} q_{i\,\nu}+p_{i\,\nu} q_{i\,\mu}\over p \cdot q}\,.
\ee
The reference momentum associated to $p_i$ is present because the polarization vector of a gluon state is only defined up to a term proportional to $p_i$; it is always permissible to add a term $\alpha \,p_\mu$ to $\pol^\pm_\mu(p)$ since the Ward identities of gauge theory guarantee that any such term will drop out of gauge invariant quantities.
The arbitrariness introduced into the definitions of the polarization vectors ($\pol^\pm(p_i) \equiv \pol^\pm(p_i, q_i)$) by the $q_i$ can be effectively exploited because of the following identities:
\bea
&&\bar{u}^+(p_j) \slashed{\pol}^-(p_i, p_j) = \bar{u}^-(p_j)\slashed{\pol}^+(p_i, p_j) = 0
\el
\slashed{\pol}^+(p_i, p_j) u^+(p_j) = \slashed{\pol}^-(p_i, p_j) u^-(p_j) = 0
\el
\pol^+(p_i, p_j)\cdot\pol^-(p_j, q) = \pol^+(p_i, q)\cdot \pol^-(p_j, p_i) = 0
\el
\pol^+(p_i, q) \cdot \pol^+(p_j, q) = \pol^-(p_i, q)\cdot \pol^-(p_j, q) = 0
\el
\pol^+(p_i, q) \cdot q = \pol^-(p_i, q) \cdot q = 0\,{\rm .}
\label{polidents}
\eea
Thankfully, the traditional, clunky notation for spinors is no longer used. It makes much more sense to define
\be
u^+(p_i) \equiv |i~\rangle ~~~~~~ u^-(p_i) \equiv |i~] ~~~~~~ \bar{u}^+(p_i) \equiv [i~| ~~~~~~ \bar{u}^-(p_i) \equiv \langle i~| \,{\rm .}
\ee
Then the eqs. (\ref{sijsh}), (\ref{polvecs+}), and (\ref{polvecs-}) discussed above can be rewritten as
\bea
&&s_{i j} = \spa{i}.j\spb{j}.i \label{modsij}\\
&&\pol^+_\mu(p_i) = {\spab{q_i}.\gamma_\mu.i \over \sqrt{2}\, \spa{q_i}.i} \label{+p}\\
&&\pol^-_\mu(p_i) = {\langle i |\, \gamma_\mu \,| q_i ] \over \sqrt{2}\, \spb{i}.{q_i}}
\label{-p}
\eea
It is also conventional to suppress explicit slashes: $p^\mu_j \spab1.\gamma_\mu.2$ is written as $\spab1.j.2$. Finally, for reference, we translate the Fierz identity, the Schouten identity, and the identity $\sum_{i = 1}^n k_i = 0$ (in an $n$ particle scattering process) into modern spinor language:
\bea
&&\spab{i}.\gamma^\mu.j \spab{k}.\gamma_\mu.\ell = 2\,\spb{j}.k \spa{\ell}.j
\el
\spa{i}.j \spa{k}.\ell + \spa{i}.k \spa{\ell}.j + \spa{i}.\ell \spa{j}.k = 0
\el
\sum_{i = 1}^n \spb{j}.i \spa{i}.k = 0
\label{shidents}
\eea
To gain some familiarity with the ideas of this Subsection, we calculate the tree-level four-gluon amplitude, $A^{tree}\left(k_1^{1234},k_2^{1234},k_3,k_4\right)$. We started evaluating the $s$-channel diagram in (\ref{fincoldemo}). Actually, this is the {\it only} diagram we have to evaluate since the $t$-channel and contact diagrams are identically zero if we choose $q_1 = q_2 = k_4$ and $q_3 = q_4 = k_1$, as can easily be verified using eqs. (\ref{polidents}). This is a remarkable feature of the spinor helicity formalism. Many terms that would be present in a traditional calculation vanish if one makes a judicious choice for the reference momenta. In fact, it is now known (and will be made clear in the next subsection) that one {\it never} needs to deal with four-gluon vertices in the evaluation of gluonic tree amplitudes. Continuing with our computation, we evaluate the $s$-channel diagram using eqs. (\ref{modsij}), (\ref{+p}), and (\ref{-p}) and rewrite all invariants in spinor language
\begin{changemargin}{-.5 in}{0 in}
\bea
&&A^{tree}\left(k_1^{1234},k_2^{1234},k_3,k_4\right) = {i\over 2}\big[\pol^{-}(k_1)\cdot \pol^{-}(k_2)(k_1-k_2)_\mu+\pol_\mu^{-}(k_2)\pol^{-}(k_1)\cdot (2 k_2+k_1)
\el
+\pol_\mu^{-}(k_1)\pol^{-}(k_2)\cdot(-2 k_1-k_2)\big] {g^{\mu \nu} \over (k_1 + k_2)^2} \big[\pol^{+}(k_3)\cdot \pol^{+}(k_4)(k_3-k_4)_\nu+\pol_\nu^{+}(k_4)\pol^{+}(k_3)\cdot (2 k_4+k_3)
\el+\pol_\nu^{+}(k_3)\pol^{+}(k_4)\cdot(-2 k_3-k_4)\big]
\elale
{i\over 2 \spa1.2\spb2.1} \big[ \pol^-(k_1)\cdot\pol^-(k_2)\pol^+(k_3)\cdot\pol^+(k_4)(k_1-k_2)\cdot(k_3-k_4)
\el+\pol^-(k_1)\cdot\pol^-(k_2)(k_1-k_2)\cdot\pol^+(k_4)(k_3 + 2 k_4)\cdot \pol^+(k_3)
+\pol^-(k_1)\cdot\pol^-(k_2)(k_1-k_2)\cdot\pol^+(k_3)(-2 k_3-k_4)\cdot\pol^+(k_4)
\el+\pol^+(k_3)\cdot\pol^+(k_4)(2 k_2+k_1)\cdot\pol^-(k_1)(k_3-k_4)\cdot\pol^-(k_2)
+\pol^-(k_2)\cdot\pol^+(k_4)(2 k_2+k_1)\cdot \pol^-(k_1)(2 k_4+k_3)\cdot\pol^+(k_3)
\el+\pol^-(k_2)\cdot\pol^+(k_3)(2 k_2+k_1)\cdot \pol^-(k_1)(-2 k_3-k_4)\cdot\pol^+(k_4)
\el
+\pol^+(k_3)\cdot\pol^+(k_4)(-2 k_1-k_2)\cdot\pol^-(k_2)(k_3-k_4)\cdot\pol^-(k_1)
\el
+\pol^-(k_1)\cdot\pol^+(k_4)(-2 k_1-k_2)\cdot\pol^-(k_2)(2 k_4+k_3)\cdot\pol^+(k_3)
\el
+\pol^-(k_1)\cdot\pol^+(k_3)(-2 k_1-k_2)\cdot\pol^-(k_2)(-2 k_3-k_4)\cdot\pol^+(k_4)\big]
\elale {i\over 2 \spa1.2\spb2.1} \big[-4\pol^-(k_2)\cdot\pol^+(k_3)k_2\cdot\pol^-(k_1)k_3\cdot\pol^+(k_4)\big]
\elale {i\over 2 \spa1.2\spb2.1} \left[-4 {\spab2.{\gamma_\mu}.4\over \sqrt{2}\,\spb2.4}{\spab1.{\gamma^\mu}.3\over \sqrt{2}\,\spa1.3}{\spab1.2.4\over\sqrt{2}\,\spb1.4}{\spab1.3.4\over\sqrt{2}\,\spa1.4} \right] \,{\rm .}
\label{4ptsimp}
\eea
\end{changemargin}
In the above, all but one (~$\pol^-(k_2) \cdot \pol^+(k_3)$~) of the dot products of polarization vectors vanished as a consequence of our choice of reference momenta. Using eqs. (\ref{shidents}), we arrive at the final result
\be
A^{tree}\left(1^{1234},2^{1234},3,4\right) = {i \spa1.2^4\over \spa1.2 \spa2.3 \spa3.4 \spa4.1}{\rm .}
\ee
In fact, this result generalizes to arbitrarily many gluons~\cite{ParkeTaylor} and we take this opportunity to define the $n$-gluon tree-level MHV amplitude (Parke-Taylor formula):
\be
A_{n;\,\spa{i}.{j}}^{\textrm{\scriptsize{MHV}}}\equiv A^{tree}\left(1,...,i^{1234},...,j^{1234},...,n\right)=i { \langle i j \rangle^4 \over \langle 1 2 \rangle \langle 2 3 \rangle...\langle n 1 \rangle} \,{\rm .}
\ee
The identities of eqs. (\ref{shidents}) are useful in simple situations (like the calculation above) but, in practice, it usually makes more sense to simplify spinor strings using complex deformations of spinor variables and the analyticity properties of scattering amplitudes. This will be discussed more in the next subsection.
\subsection{BCFW Recursion}
\label{BCFW}
In this Subsection we review the recursion relation of Britto, Cachazo, Feng, and Witten (BCFW recursion), a powerful tool used primarily for the calculation of tree-level helicity amplitudes in massless gauge theories. Shortly after BCFW recursion was developed in~\cite{origBCFW} it was realized, probably by the authors of~\cite{allNMHV} and perhaps also by the authors of the original paper, that BCFW recursion is also useful when confronted with the problem of simplifying messy linear combinations of rational functions of spinor products. This is because, in many situations of practical interest, physical linear combinations of spinor products are tightly constrained by their singularity structure.
Although, it is possible~\cite{SUSYBCFW,DHsuperBCFW} to write down a manifestly $\Nsym$ supersymmetric version of BCFW recursion, we present the recursion relation in its original incarnation. Suppose that all tree-level $(n-1)$ and lower-point gluon helicity amplitudes are known. Britto, Cachazo, Feng, and Witten showed that one can write any $n$-gluon tree amplitude in terms of particular deformations of the known lower-point gluon amplitudes. The algorithm will be easy to understand once we explain the concept.
The main idea is that scattering amplitudes are analytic functions of all their inputs. To exploit this analyticity it makes sense to complexify all four-momenta in the problem before going any further. Now, imagine factorizing the amplitude we wish to calculate on a collinear or multi-particle pole\footnote{If these notions are not familiar, see~\cite{Dixon96rev} for an elementary discussion.}, say the one associated to the invariant $K^2 \equiv (k_i +\cdots+k_{i+j})^2$:
\bea
&&A^{tree}\left(k_1^{h_1},\cdots,k_i^{h_i},\cdots,k_{i+j}^{h_{i+j}},\cdots,k_n^{h_n}\right) \stackrel{K^2 \rightarrow~ 0}{\longrightarrow} \nn
&& \sum_h A^{tree}\left(k_i^{h_i},\cdots,k_{i+j}^{h_{i+j}},K^h\right){-i\over K^2}~A^{tree}\left(-K^{-h},k_{i+j}^{h_{i+j}},\cdots,k_n^{h_n},k_1^{h_1},\cdots,k_{i-1}^{h_{i-1}}\right) \,{\rm .}\nn
\label{factor}
\eea
This is the picture that one should have in mind. Intuitively, BCFW recursion is based on the observation that the set of all such limits of a particular $n$-point gluon tree amplitude actually carry all the information necessary to reconstruct the complete tree amplitude. In general, one should only expect this approach to work for tree amplitudes; we shall see later that factorization properties are typically not sufficient to fix the analytical structure of amplitudes at the one-loop level and higher.
To try and realize this intuitive picture of reversing collinear and multi-particle factorization limits more concretely, BCFW found it convenient to consider a particular analytical continuation applicable to general gluon tree amplitudes (under appropriate assumptions). Consider the following deformation of the holomorphic spinor associated to $k_\ell$ and the anti-holomorphic spinor associated to $k_m$:
\bea
\lambda_\ell &&\rightarrow \lambda_\ell(z) = \lambda_\ell + z \lambda_m\nn
\tilde{\lambda}_m &&\rightarrow \tilde{\lambda}_m(z) = \tilde{\lambda}_m - z \tilde{\lambda}_\ell \,{\rm ,}
\label{BCFWs}
\eea
where $z$ is a complex parameter. At the level of spinors it is not even obvious that this complex deformation is well-defined. The corresponding relations for $k_\ell$ and $k_m$
\bea
k_\ell^\mu &&\rightarrow k_\ell^\mu(z) = k_\ell^\mu + {z \over 2}\spab{m}.\gamma^\mu.\ell \nn
k_m^\mu &&\rightarrow k_m^\mu(z) = k_m^\mu - {z \over 2}\spab{m}.\gamma^\mu.\ell
\eea
make it clear that the BCFW shift (eq. (\ref{BCFWs})) preserves overall momentum conservation
\be
\sum_{i=1}^n k_i = 0
\ee
and, furthermore, a small calculation using (\ref{shidents}) makes it clear that
\be
k_\ell(z)^2 = k_m(z)^2 = 0 \,{\rm .}
\ee
So the BCFW deformation is well-defined after all.
We now evaluate the integral
\be
{1\over 2 \pi i} \oint_C dz {A^{tree}\left(k_1^{h_1},\cdots,k_m^{h_m}(z),\cdots,k_\ell^{h_\ell}(z),\cdots,k_n^{h_n}\right)\over z}
\ee
in two different ways, assuming that $C$ is a very large circle in the complex $z$-plane that encloses all poles of the integrand. Of course, we know the answer must be zero
\be
{1\over 2 \pi i} \oint_C dz {A^{tree}\left(k_1^{h_1},\cdots,k_m^{h_m}(z),\cdots,k_\ell^{h_\ell}(z),\cdots,k_n^{h_n}\right)\over z} = 0
\label{BCFW0}
\ee
by virtue of the choice of contour and Cauchy's theorem. On the other hand\footnote{Here we can proceed only under the assumption that the integrand goes to zero fast enough that $C$ can be safely taken to infinity. This assumption is justified for a large class of shifts~\cite{origBCFW}.}, we can write
\bea
&&{1\over 2 \pi i} \oint_C dz {A^{tree}\left(k_1^{h_1},\cdots,k_m^{h_m}(z),\cdots,k_\ell^{h_\ell}(z),\cdots,k_n^{h_n}\right)\over z} = A^{tree}\left(k_1^{h_1},\cdots,k_m^{h_m},\cdots,k_\ell^{h_\ell},\cdots,k_n^{h_n}\right) \el
+ \sum_\alpha {\rm Res}_{z = z_\alpha}\Bigg\{ {A^{tree}\left(k_1^{h_1},\cdots,k_m^{h_m}(z),\cdots,k_\ell^{h_\ell}(z),\cdots,k_n^{h_n}\right)\over z}\Bigg\}
\label{BCFWrel}
\eea
where $\alpha$ is indexing the poles of the amplitude in $z$ induced in particular factorization channels by the BCFW shift. Though it is not at all obvious, it can be shown~\cite{origBCFW} that it is always possible to find {\it some} shift (and associated pair $(k_m^{h_m},k_\ell^{h_\ell})$) for which (\ref{BCFWrel}) is valid (we focus on pure glue for now). Combining eqs. (\ref{BCFW0}) and (\ref{BCFWrel}), we see that the amplitude at the origin of $z$-space (which is what we want) is given by a sum of residues of the shifted amplitude divided by $z$:
\bea
&&A^{tree}\left(k_1^{h_1},\cdots,k_m^{h_m},\cdots,k_\ell^{h_\ell},\cdots,k_n^{h_n}\right) = - \sum_\alpha {\rm Res}_{z = z_\alpha}\Bigg\{ {A^{tree}\left(k_1^{h_1},\cdots,k_m^{h_m}(z),\cdots,k_\ell^{h_\ell}(z),\cdots,k_n^{h_n}\right)\over z}\Bigg\} \,{\rm .}\nn
\label{BCFWfin}
\eea
Since the physical poles that amplitudes can develop must all be of the form
\be
{1\over (k_i +\cdots+k_{i+j})^2}
\ee
for various $i$ and $j$, it is possible to develop a recursive algorithm based on eqs. (\ref{factor}) and (\ref{BCFWfin}).
Specifically, an $n$-point gluon amplitude can be expressed as a sum over factorization channels such that $k_m$ and $k_\ell$ are not both on the same side of the intermediate state\footnote{This is because no residue in $z$ can arise unless there is non-trivial $z$ dependence in $(k_i +\cdots+k_{i+j})^2$. The fact that BCFW shifts respect momentum conservation makes such non-trivial $z$ dependence impossible if both shifted particles are on the same side of the intermediate state.} in (\ref{factor}). Each factorization channel should be evaluated after the chosen BCFW shift has been made, with the value of $z$ fixed by solving the equation $(k_i +\cdots+k_{i+j})^2(z) = 0$. The technique is best illustrated through a simple example.
As such, we derive the six-gluon tree amplitude $A^{tree}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$ using the deformation
\bea
&&\lambda_4(z) = \lambda_4 + z \lambda_3\nn
&&\tilde{\lambda}_3(z) = \tilde{\lambda}_3 - z \tilde{\lambda}_4 \,{\rm .}
\eea
Possible contributions for this choice of shift are shown pictorially in Figure \ref{BCFWfig}.
\FIGURE{
\resizebox{.75\textwidth}{!}{\includegraphics{BCFW.eps}}
\caption{The three BCFW diagrams that could potentially contribute to $A^{tree}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$.}
\label{BCFWfig}}
Clearly the multi-particle channel gives zero contribution, due to the fact that gluon amplitudes with zero or one negative helicity (and all the rest positive) are protected by supersymmetry (see Appendix \ref{SWI} for a discussion). The first and second graphs in Figure \ref{BCFWfig} are non-zero and, if we label them $\a_1$ and $\a_2$, we have
\bea
\a_1 &=& \sum_h A^{tree}\left(k_2^{1234},k_3^{1234}(z),-(k_2+k_3)^{-h}(z)\right){-i\over (k_2+k_3)^2}A^{tree}\left((k_2+k_3)^h(z),k_4(z),k_5,k_6,k_1^{1234}\right)
\elale A^{tree}\left(k_2^{1234},k_3^{1234}(z),-K_{23}(z)\right){-i\over s_{23}}A^{tree}\left(K_{23}^{1234}(z),k_4(z),k_5,k_6,k_1^{1234}\right)
\elale {i \spa2.{\hat{3}}^4\over \spa2.{\hat{3}} \spa{\hat{3}}.{-\hat{K}_{23}} \spa{-\hat{K}_{23}}.2} {-i\over s_{23}} {i \spa{\hat{K}_{23}}.1^4\over \spa{\hat{K}_{23}}.{\hat{4}}\spa{\hat{4}}.5\spa5.6\spa6.1\spa1.{\hat{K}_{23}}}
\eea
and
\bea
\a_2 &=& \sum_h A^{tree}\left(k_4(z),k_5,-(k_4+k_5)^{-h}(z)\right){-i\over (k_4+k_5)^2}A^{tree}\left((k_4+k_5)^h(z),k_6,k_1^{1234},k_2^{1234},k_3^{1234}(z)\right)
\elale A^{tree}\left(k_4(z),k_5,-K_{45}^{1234}(z)\right){-i\over s_{45}}A^{tree}\left(K_{45}(z),k_6,k_1^{1234},k_2^{1234},k_3^{1234}(z)\right)
\elale {i \spb{\hat{4}}.5^4\over \spb{\hat{4}}.5 \spb5.{-\hat{K}_{45}} \spb{-\hat{K}_{45}}.{\hat{4}}}{-i\over s_{45}} {i \spb{\hat{K}_{45}}.6^4\over \spb{\hat{K}_{45}}.6 \spb6.1 \spb1.2 \spb2.{\hat{3}} \spb{\hat{3}}.{\hat{K}_{45}}} \,{\rm .}
\eea
In the above, a hatted spinor variable reminds us which spinors have been shifted. At this stage, we see why complexified momenta are necessary; if we worked with real momenta the three point amplitudes $A^{tree}\left(k_2^{1234},k_3^{1234}(z),-K_{23}(z)\right)$ and $A^{tree}\left(k_4(z),k_5,-K_{45}^{1234}(z)\right)$ in the above would vanish identically. In order to further simplify the above equations we need
\bea
\spa{a}.{\hat{K}_{23}} &=& {\spab{a}.{K_{23}}.4 \over \spb{\hat{K}_{23}}.4} ~~~~{\rm and} \nn
\spb{\hat{K}_{45}}.{b} &=& {\spab{3}.{K_{45}}.b \over \spa{3}.{\hat{K}_{23}}}
\label{BCFWid}
\eea
together with the solutions of $(k_2+k_3)^2(z)=0$ and $(k_4+k_5)^2(z)=0$:
\bea
z_{23} &=& {s_{23}\over \spab3.2.4} ~~~~{\rm and} \nn
z_{45} &=& -{s_{45}\over \spab3.5.4} \,{\rm .}
\eea
It may concern the reader that we are not able to express all quantities appearing in $\a_1$ and $\a_2$ explicitly in terms of unshifted spinors. Actually, this will turn out not to be a problem; all factors of $\spb{\hat{K}_{23}}.4$ and $\spa{3}.{\hat{K}_{23}}$ cancel out of the final result as can easily be verified by counting how many times they will appear in the numerator and in the denominator of each expression. Applying identities (\ref{BCFWid}) to $\a_1$ and $\a_2$, we finally obtain
\bea
\a_1 &=& {i \spa2.3^4 \spab1.{K_{23}}.4^4 \over \spa2.3 \spab3.{K_{23}}.4 \spab2.{K_{23}}.4 s_{23} (\spab4.{K_{23}}.4+z_{23}\spab3.{K_{23}}.4)(\spa4.5+z_{23}\spa3.5) \spa5.6 \spa6.1 \spab1.{K_{23}}.4}\nn
\a_2 &=& {i \spb4.5^4 \spab3.{K_{45}}.6^4 \over \spb4.5 \spab3.{K_{45}}.5 \spab3.{K_{45}}.4 s_{45} \spab3.{K_{45}}.6 \spb6.1 \spb1.2 (\spb2.3-z_{45}\spb2.4)(\spab3.{K_{45}}.3-z_{45}\spab3.{K_{45}}.4)} \nn
\eea
and
\bea
A^{tree}(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6) = \a_1+\a_2
\eea
which can be confirmed (numerically using {\it e.g.} S@M~\cite{S@M}) by comparing to the result given in~\cite{ManganoParke}.
Before moving on to loop-level calculations, we need to say a few words about the application of BCFW recursion to the simplification of messy rational linear combinations of spinor products. This works well when there is reason to believe that an expression for which you have an ugly formula naturally collapses down to a single term. BCFW then allows you to systematically guess the form of the allegedly simple common denominator. To better understand this we consider the following thought experiment. Suppose that instead of choosing $q_1 = q_2 = k_4$ and $q_3 = q_4 = k_1$ to evaluate $A^{tree}\left(k_1^{1234},k_2^{1234},k_3,k_4\right)$ in eq. (\ref{4ptsimp}) we instead made an unintelligent choice of reference momenta that resulted in more than one Feynman diagram making a non-zero contribution to the amplitude. Now imagine making a table of all possible BCFW shifts, numerically evaluated at a randomly chosen non-degenerate phase-space point. Initially, we would find that all shifts produce several poles in $z$. We could deduce that everything should be put over the common denominator $\spa1.2\spa2.3\spa3.4\spa4.1$ by systematically multiplying the unsimplified expression for $A^{tree}\left(k_1^{1234},k_2^{1234},k_3,k_4\right)$ by each invariant in the problem and then checking to see if our table of all BCFW shifts has a simpler $z$-pole structure or not. In short order, we would be able to deduce that
\be
A^{tree}\left(k_1^{1234},k_2^{1234},k_3,k_4\right) = {C\over \spa1.2 \spa2.3\spa3.4\spa4.1} \,{\rm .}
\ee
This is powerful because evaluating BCFW shifts numerically is a lot less labor intensive than attempting an analytic simplification. Once the denominator is determined, it is then a simple matter to use dimensional analysis and the known little group rescaling properties\footnote{The little group of the Lorentz group in four dimensions for a massless external state is $SO(2)$. Keeping in mind $k_i = \lambda_i \tilde{\lambda}_i$, we expect scattering amplitudes to transform covariantly under the rescaling $\lambda_i \rightarrow t_i \lambda,\,\tilde{\lambda}_i \rightarrow t_i^{-1} \tilde{\lambda}_i$. The precise transformation law of course depends on the helicities of the massless external particles in the scattering process: we have $A(t_i \lambda_i,\,t_i^{-1} \tilde{\lambda}_i; h_i) = \left(\prod_{i=1}^n t_i^{-2 h_i}\right) A(\lambda_i,\,\tilde{\lambda}; h_i)$.} of the amplitude to fix $C = \spa1.2^4$. This thought experiment might seem somewhat contrived, but, at least in $\Nsym$, many loop-level calculations result in objects that are naturally put over a single denominator. While more non-trivial amplitudes have constituents like $\spab2.{5+3}.6$ and $\langle6|4+5|2+3|1\rangle$, it is straightforward to try a large number of guesses in a fraction of a second using a computer. Once the denominator is determined, it is typically possible to apply the dual constraints of little group covariance and correct dimensionality to great effect. All new results in this work were simplified using some variant of this technique.
Finally, we should emphasize that arbitrary tree-level scattering processes in $\Nsym$ can be generated using BCFW recursion~\cite{SUSYBCFW,DHsuperBCFW} provided that the above discussion is supersymmetrized and described in the language of the $\Nsym$ on-shell superspace introduced in Section \ref{supercomp}.
\subsection{Generalized Unitarity in Four Dimensions}
\label{GU4}
We now turn to loop-level calculations. Most of the calculations in this paper are at the one-loop level, but the ideas reviewed in this subsection and the next, with appropriate modifications, have been applied to multi-loop calculations as well. The program of {\it generalized unitarity} pioneered by Bern, Dixon, Dunbar, and Kosower~\cite{BDDKMHV,BDDKNMHV} was developed to replace the traditional Feynman diagram based approach to loop-level calculations. For most loop-level applications it is much more efficient to use generalized unitarity diagrams because they are built out of on-shell tree amplitudes and the number of contributions scales roughly like number of topologies times the number of particle species allowed to run in the loop. This is a already a big improvement over the usual Feynman diagram expansion. As will be made clear, once you take into account the fact that each diagram is also easier to compute, the generalized unitarity approach is even more attractive.
As was made clear in the introduction, any one-loop planar $\Nsym$ amplitude can be written as
\be
A_1^{1-{\rm loop}}(k_1^{h_1},\cdots,k_n^{h_n}) = \sum_\alpha C_\alpha I^{(\alpha)}_4 + \Ord(\e)\,{\rm ,}
\label{gen1loop}
\ee
where $\alpha$ labels the specific kinematic structure of the box integral (more on our labeling scheme below) and each box integral is evaluated through $\Ord(\e^0)$. Much of the power of the generalized unitarity technique comes from (\ref{gen1loop}), so it is worth spending some time trying to understand it. It turns out that (\ref{gen1loop}) is very special to $\Nsym$. An equation similar to (\ref{gen1loop}) would hold for generic $\mathcal{N}=2$ and $\mathcal{N}=1$ gauge theories, except that triangle and bubble integrals (discussed briefly in the introduction) would have to be added to the box integrals on the right-hand side~\cite{1001lessons, BDDKMHV}. Less supersymmetry ({\it i.e.} $\mathcal{N}=1$ super Yang-Mills) makes such a relation less powerful and a little more difficult to work with (see {\it e.g.}~\cite{Forde}). For $\mathcal{N}=0$ it becomes harder still and we really need ideas from the next subsection to make an analogous construction.
Before we start, we need a convenient way to enumerate the box topologies for a planar $n$-particle scattering process. Consider, as usual, a regular $n$-gon with one external line attached at each vertex. In an approach based on Feynman diagrams this would be the highest-point Feynman integral that could possibly appear prior to integral reduction. There are
\be
\left(\begin{array}{c} n \\ n-4\end{array}\right) = {n! \over (n-4)! 4!}
\label{noboxes}
\ee
ways to collapse this $n$-gon down to a box. Consequently, it is natural to label each box in the integral basis by an $n-4$-tuple of integers corresponding to the internal lines that need to be erased to produce the box in question\footnote{Our convention will be to start counting with the propagator connecting the 1st and $n$th vertices.}. In this work we will mostly be interested in $n = 6$ for which (\ref{noboxes}) gives 15 boxes.
This formula gives the largest number of boxes that could possibly appear. Depending on the helicity configuration, certain classes of boxes may make no contribution to the sum in (\ref{gen1loop}). To be less cryptic, we write (\ref{gen1loop}) out explicitly for $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6\right)$ and $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$:\footnote{In eqs. (\ref{6ptgMHV}) and (\ref{6ptgNMHV}) $s_i \equiv s_{i\,i+1}$ and $t_i \equiv s_{i\,i+1\,i+2}$, where indices are mod 6. We will frequently use this notation in our discussions of six-point scattering. The notation can, of course, be generalized to describe a basis of kinematic invariants for arbitrary $n$. For instance, at the eight-point level, invariants like $w_{1} \equiv (k_1+k_2+k_3+k_4)^2$ will enter.}
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6\right) = {A_{n;\,\spa{1}.{2}}^{\textrm{\scriptsize{MHV}}} \over 2}\Big(- s_3 s_4 I_{4}^{(1,2)}- {s_4 s_5} I_{4}^{(2,3)} - {s_5 s_6} I_{4}^{(3,4)} -{s_1 s_6 }I_{4}^{(4,5)} - {s_1 s_2 }I_{4}^{(5,6)}
\el - {s_2 s_3 } I_{4}^{(1,6)}+ (s_3 s_6 - t_2 t_3) I_{4}^{(1,4)} + (s_1 s_4 - t_1 t_3) I_{4}^{(2,5)} + (s_2 s_5 - t_1 t_2) I_{4}^{(3,6)} + \Ord(\e)\Big)
\label{6ptgMHV}
\eea
\cme
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right) = -{i\over2} {\spa1.2 \spa2.3 \spb4.5 \spb5.6 \spab3.{(1+2)}.6 \spab1.{(2+3)}.4 t_1^3 \over s_1 s_2 s_4 s_5 (t_1 t_2-s_2 s_5)(t_1 t_3-s_1 s_4)}\Big( s_4 s_5 I_4^{(2,3)}+s_1 s_2 I_4^{(5,6)}
\el +s_6 t_1 I_4^{(3,5)} + s_3 t_1 I_4^{(2,6)}\Big)-{i\over2}\Bigg(\bigg({\spab1.{(2+3)}.4\over t_2}\bigg)^4 {\spa2.3 \spa3.4 \spb5.6 \spb6.1 \spab4.{(2+3)}.1 \spab2.{(3+4)}.5 t_2^3 \over s_2 s_3 s_5 s_6 (t_2 t_3-s_3 s_6)(t_2 t_1-s_2 s_5)}
\el+\bigg({\spa2.3 \spb5.6 \over t_2}\bigg)^4 {\spb2.3 \spb3.4 \spa5.6 \spa6.1 \spab1.{(2+3)}.4 \spab5.{(3+4)}.2 t_2^3 \over s_2 s_3 s_5 s_6 (t_2 t_3-s_3 s_6)(t_2 t_1-s_2 s_5)}\Bigg) \Big( s_5 s_6 I_4^{(3,4)}+s_2 s_3 I_4^{(6,1)}+s_1 t_2 I_4^{(4,6)}+s_4 t_2 I_4^{(1,3)}\Big)
\el-{i\over2}\Bigg(\bigg({\spab3.{(1+2)}.6\over t_3}\bigg)^4 {\spa6.1 \spa1.2 \spb3.4 \spb4.5 \spab2.{(6+1)}.5 \spab6.{(1+2)}.3 t_3^3 \over s_6 s_1 s_3 s_4 (t_3 t_1-s_1 s_4)(t_3 t_2-s_6 s_3)}
\el+\bigg({\spa1.2 \spb4.5 \over t_3}\bigg)^4 {\spb6.1 \spb1.2 \spa3.4 \spa4.5 \spab5.{(6+1)}.2 \spab3.{(1+2)}.6 t_3^3 \over s_6 s_1 s_3 s_4 (t_3 t_1-s_1 s_4)(t_3 t_2-s_6 s_3)}\Bigg) \Big( s_6 s_1 I_4^{(4,5)}+s_3 s_4 I_4^{(1,2)}+s_2 t_3 I_4^{(1,5)}
\el+s_5 t_3 I_4^{(2,4)}\Big) + \Ord(\e) \,{\rm .}
\label{6ptgNMHV}
\eea
\end{changemargin}
In the six-point MHV amplitude, all of the boxes with two adjacent external masses enter with zero coefficient and in the six-point NMHV amplitude all of the boxes with two diametrically opposed external masses enter with zero coefficient. Boxes with two external masses are special in that they have different analytic structures depending on whether the two external masses are adjacent or diametrically opposed. For historical reasons, two mass box integrals with adjacent external masses are called two mass hard and two mass box integrals with diametrically opposed external masses are called two mass easy (see Figure \ref{boxints}). The six different types of box integrals that appear in planar one-loop $\Nsym$ calculations are summarized in Figure \ref{boxints}.
\FIGURE{
\resizebox{0.65\textwidth}{!}{\includegraphics{Boxes.eps}}
\vbox{\vskip 1 in}
\caption{The one-loop basis of scalar integrals in $\Nsym$ through $\Ord(\e^0)$ consists of the: zero mass box integral, one mass box integral, two mass easy box integral, two mass hard box integral, three mass box integral, and four mass box integral.}
\label{boxints}}
Explicit formulae for these integral functions will be provided shortly, but first let us say a few more words about (\ref{gen1loop}). Clearly, the zero mass box will only appear for the special case of four particle scattering. For general $n$, planar $\Nsym$ MHV amplitudes are built out of one mass and two mass easy boxes~\cite{BDDKMHV} and planar $\Nsym$ NMHV amplitudes are built out of one mass, two mass easy, two mass hard, and three mass boxes~\cite{allNMHV}; four mass boxes don't appear until the eight-point ${\rm N}^2$MHV level. In particular, the absence of two mass easy basis integrals in $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$ does not generalize to higher $n$ NMHV amplitudes.
Before going further, we need to carefully define the one-loop Feynman integrals which enter into our perturbative calculations. In general, the Feynman integrals that enter into calculations in massless gauge theories have severe IR divergences that need to be regulated. In dimensional regularization~\cite{origtHooftVelt} one regulates the IR divergences by analytically continuing the scattering amplitude under consideration from $D = 4$ to $D = 4 - 2 \e$ and then computing its Laurent expansion about $\e = 0$ (see Appendix \ref{dimregs} for background). We make the definition
\bea
I_n^{D=4-2 \e}
&\equiv& i (-1)^{n+1}(4\pi)^{2-\e} \int
{d^{4-2\e} p \over (2\pi)^{4-2\e} }
{ 1 \over p^2 \ldots
(p-\sum_{i=1}^{n-1} K_i )^2 }
\nonumber \\ &=& i (-1)^{n+1} (4\pi)^{2-\e} \int
{d^{4} p \over (2\pi)^{4} }
{d^{-2\e} \mu \over (2\pi)^{-2\e} }
{ 1 \over (p^2 - \mu^2) \ldots (
(p-\sum_{i=1}^{n-1} K_i )^2 - \mu^2) } \,{\rm .}\nn
\label{IntDef}
\eea
The prefactor $i (-1)^{n+1} (4\pi)^{2-\e}$ cancels a factor of $i (-1)^{n} (4\pi)^{\e-2}$ that always arises in the calculation of one-loop Feynman integrals and on the second line we have explicitly separated out the integrations into four dimensional and $-2\e$ dimensional pieces. This second form will be useful in the next subsection and later on. In what follows, we give explicit results for a representative sample of basis integrals that enter into planar one-loop $\Nsym$ scattering amplitudes. For our representative of each integral species (Figure \ref{boxints}), we shall use the kinematics of the point at which the basis integral first enters into the sum in eq. (\ref{gen1loop}). For example, we use five-point kinematics for the one mass box because this integral first appears in planar one-loop five-point MHV amplitudes. A technical point to be aware of is that, in expanding an $L$-loop Feynman integral in $\e$, a factor of $e^{\gamma_E \e\,L}$ is expanded with the Feynman integral in order to prevent a proliferation of factors of $\gamma_E$ that would otherwise occur. This remark and definition (\ref{IntDef}) explain why the prefactors in eq. (\ref{finloopcol}) have the form that they do. Following~\cite{BDDKNMHV} (and checking against the more recent article~\cite{allNMHV}), the epsilon expansions through $\Ord(\e^0)$ of the basis integrals are presented below in the same order that they appeared in Figure \ref{boxints}:
\be
I_4 = {\G(1+\e)\G^2(1-\e)\over s t \,\G(1-2\e)}
\bigg\{
{2 \over \e^2} \Big[ ( -s)^{-\e}+ (-t)^{-\e} \Big]
- \ln^2\left( {-s \over - t} \right) - \pi^2 \bigg\} \,{\rm ,}
\label{box0}
\ee
\bea
I_{4}^{(5)} &=& { -2 \G(1+\e)\G^2(1-\e) \over s_1 s_2 \,\G(1-2\e)}
\bigg\{
-{1\over\e^2} \Big[ (-s_1)^{-\e} +
(-s_2 )^{-\e} - (-s_4)^{-\e} \Big]
\el + \Li_2\left(1-{s_4 \over s_1}\right)
+ \ \Li_2\left(1-{s_4 \over s_2}\right)
+{1\over 2} \ln^2\left({-s_1 \over -s_2}\right)\
+ {\pi^2\over6} \bigg\}\,{\rm ,}
\label{box1}
\eea
\bea
I_{4}^{(2,5)} &=& { -2 \G(1+\e)\G^2(1-\e) \over (t_1 t_3 -s_1 s_4) \,\G(1-2\e)}
\bigg\{
- {1\over\e^2} \Big[ (-t_1)^{-\e} + (-t_3)^{-\e}
- (-s_1)^{-\e} - (-s_4)^{-\e} \Big]
\el + \Li_2\left(1-{s_1 \over t_1}\right)
+ \Li_2\left(1-{s_1 \over t_3}\right)
+ \Li_2\left(1-{s_4 \over t_1}\right)
\el + \Li_2\left(1-{s_4 \over t_3}\right)
- \Li_2\left(1-{s_1 s_4
\over t_1 t_3}\right)
+ {1\over 2} \ln^2\left({-t_1 \over -t_3}\right) \bigg\}\,{\rm ,}
\label{box2e}
\eea
\bea
I_{4}^{(2,4)} &=& { -2 \G(1+\e)\G^2(1-\e) \over s_5 t_3 \,\G(1-2\e)}
\bigg\{
- {1\over 2\e^2} \Big[ (-s_5)^{-\e} + 2(-t_3)^{-\e}
- (-s_1)^{-\e} - (-s_3)^{-\e} \Big]
\el -{1\over2}\ln\left({-s_1 \over -s_5}\right)\ln\left({-s_3 \over -s_5}\right)+ {1\over 2} \ln^2\left({-s_5 \over -t_3 }\right) + \Li_2\left(1-{ s_1 \over t_3}\right)
+ \Li_2\left(1-{s_3 \over t_3 }\right) \bigg\}\,{\rm ,}\nn
\label{box2h}
\eea
\bea
I_{4}^{(3,5,7)} &=& { -2 \G(1+\e)\G^2(1-\e) \over (t_1 t_6-s_2 s_6) \,\G(1-2\e)}
\bigg\{
-{1\over 2\e^2} \Big[ (-t_1 )^{-\e} + (-t_6)^{-\e}
- (-s_2)^{-\e}
- (-s_6)^{-\e} \Big]
\el -{1\over2}\ln\left({-s_2 \over -t_6}\right)\ln\left({-s_4 \over -t_6}\right)-{1\over2}\ln\left({-s_4 \over -t_1}\right)\ln\left({-s_6 \over -t_1}\right)+\ {1\over2}\ln^2\left({-t_1 \over -t_6}\right)
\el + \Li_2\left(1-{s_2\over t_1 }\right)
+ \Li_2\left(1-{s_6 \over t_6}\right)
- \Li_2
\left(1-{s_2 s_6\over t_1 t_6 }\right) \bigg\}\,{\rm ,}
\label{box3}
\eea
\bea
I_{4}^{(2,4,6,8)} &=& {-\G(1+\e)\G^2(1-\e) \over w_1 w_3 \rho\,\G(1-2\e)}
\bigg\{ - \Li_2\left({1\over2}(1-\lambda_1+\lambda_2+\rho)\right)
+ \Li_2\left({1\over2}(1-\lambda_1+\lambda_2-\rho)\right)
\el - \Li_2\left(
-{1\over2\lambda_1}(1-\lambda_1-\lambda_2-\rho)\right)
+ \Li_2\left(-{1\over2\lambda_1}(1-\lambda_1-
\lambda_2+\rho)\right)
\el - {1\over2}\ln\left({\lambda_1\over\lambda_2^2}\right)
\ln\left({ 1+\lambda_1-\lambda_2+\rho \over 1+\lambda_1
-\lambda_2-\rho }\right) \bigg\} \,{\rm ,}
\label{box4}
\eea
where
\be
\rho \equiv \sqrt{1 - 2\lambda_1 - 2\lambda_2
+ \lambda_1^2 - 2\lambda_1\lambda_2 + \lambda_2^2}\, {\rm ,}
\ee
and
\bea
\lambda_1 &=& {s_1 s_3\over w_1 w_3} \, {\rm ,}
\\
\lambda_2 &=& {s_5 s_7 \over w_1 w_3} \,{\rm .}
\eea
We will never need to use these results directly to compute our amplitudes. By combining a power variant of the optical theorem for Feynman diagrams (if unfamiliar see {\it e.g.}~\cite{PeskinSchroeder}), the fact that the above integrals form a complete basis for one-loop planar $\Nsym$ scattering amplitudes through $\Ord(\e^0)$, and the fact that the above integrals are uniquely determined by how they develop residues when viewed as contour integrals in $\mathbb{C}^4$ (there is a canonical choice for the contours which shall be discussed shortly), we can deduce all of the coefficients in the sum of eq. (\ref{gen1loop}) for any given amplitude without explicitly evaluating a single Feynman integral~\cite{BDDKMHV,SharpLS}. This is the power of the generalized unitarity technique in the context of planar $\Nsym$.\footnote{As mentioned before, similar arguments can be used to greatly simplify calculations in theories with less supersymmetry. However, such calculations are usually harder because the ansatz for the amplitude (right-hand side of (\ref{gen1loop})) will be less tightly constrained and may contain many more terms.} This is a remarkable claim, so we examine it in detail. First let us be clear about one subtle point, in the above formulae all of the basis integrals were evaluated in the Euclidean region after Wick rotation because this is technically easier; in the Euclidean region the Laurent expansions of Feynman loop integrals have real coefficients. In what follows we deal with Feynman integrals before Wick rotation and therefore it makes sense to ignore the prefactor of $i (-1)^{n+1} (4\pi)^{2-\e}$ until we actually have our final answer and are ready to Wick rotate from Minkowski space to Euclidean space. At that point it can be trivially restored. In general, it is useful to think about loop calculations in two phases, the first being the determination of the coefficients in eq. (\ref{gen1loop}) and the second being the actual analytical evaluation of the basis integrals.
To better understand how generalized unitarity is superior to Feynman diagrams, we compare the two approaches for the simple example of the five-point amplitude \\$A^{1-{\rm loop}}_1(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$. This exercise is perfect for us because the actual answer is well-known and it fits on a page even if written out in gory detail:
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5) = i {A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} \left( s_{2}s_{3} I_4^{(1)}+s_{3}s_{4} I_4^{(2)}+ s_{4}s_{5} I_4^{(3)}+s_{5}s_{1} I_4^{(4)}+s_{1}s_{2} I_4^{(5)} + \Ord(\e)\right) \,{\rm ,}
\nn
&&A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5) = i {A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} \bigg( s_{2}s_{3} \int
{d^{4-2\e} p \over (2\pi)^{4-2\e} }
{ 1 \over (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
(p+k_5)^2 }
\el+s_{3}s_{4} \int
{d^{4-2\e} p \over (2\pi)^{4-2\e} }
{ 1 \over p^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
(p+k_5)^2 }+ s_{4}s_{5} \int
{d^{4-2\e} p \over (2\pi)^{4-2\e} }
{ 1 \over p^2 (p-k_1)^2(p+k_4+k_5)^2
(p+k_5)^2 }
\el+s_{5}s_{1} \int
{d^{4-2\e} p \over (2\pi)^{4-2\e} }
{ 1 \over p^2 (p-k_1)^2 (p-k_1-k_2)^2
(p+k_5)^2 }+s_{1}s_{2} \int
{d^{4-2\e} p \over (2\pi)^{4-2\e} }
{ 1 \over p^2 (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
}
\el+ \Ord(\e)\bigg) \,{\rm .}
\label{5ptapprox}
\eea
\end{changemargin}
We now argue that, by complexifying the loop momentum, $p$, and changing the contour from the usual one over all $\mathbb{R}^{1,3}$ to a particular 4-torus $T^4 \cong S^1\times S^1 \times S^1 \times S^1$ embedded in $\mathbb{C}^4$, we can isolate a single coefficient, say $i s_1 s_2 A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}/2$, on the right-hand side of (\ref{5ptapprox}).
To obtain a meaningful relation, we will of course be interested in somehow performing the same sequence of operations on the left-hand side of (\ref{5ptapprox}). This is less obvious and is where the optical theorem for Feynman diagrams comes into play. Ultimately, the left-hand side will be evaluated using a variant of the optical theorem, generalized unitarity, a name coined by in Eden, Landshoff, Olive, and Polkinghorne in their classic text~\cite{AnalyticSmatrix}.\footnote{We should point out that most of the formalism reviewed in this subsection was developed in~\cite{quadcuts}.} For now, one should simply remember that we must at some point return to the question of how to make sense of the ``raw'' expression for the amplitude (that delivered directly from Feynman diagrams) with respect to whatever contours of integration we introduce on the right-hand side of (\ref{5ptapprox}).
The idea proposed above is well-motivated. One of the main reasons loop-level computations (even in UV finite theories like $\Nsym$) are hard is that one has to worry about regulating divergences in the momentum integrals over all $p$-space. It would be nice if there was some meaningful IR-finite data that one could extract by considering the amplitude on contours in $\mathbb{C}^4$ other than $\mathbb{R}^{1,3}$. We now show how this is realized in the present example. We consider eq. (\ref{5ptapprox}) on a contour $\G_p$ defined by
\be
\G_p = \{p \in\mathbb{C}^4: ~~|p^2|<\D,\,|(p-k_1)^2|<\D,\,|(p-k_1-k_2)^2|<\D,\,|(p+k_4+k_5)^2|<\D \}
\ee
for sufficiently small $\D$. On this contour, we can evaluate the right-hand side of eq. (\ref{5ptapprox}),
\begin{changemargin}{-.6 in}{0 in}
\bea
&& i {A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} \bigg( s_{2}s_{3} \int_{\G_p}
{d^{4} p \over (2\pi i)^{4} }
{ 1 \over (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
(p+k_5)^2 }
\el+s_{3}s_{4} \int_{\G_p}
{d^{4} p \over (2\pi i)^{4} }
{ 1 \over p^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
(p+k_5)^2 }+ s_{4}s_{5} \int_{\G_p}
{d^{4} p \over (2\pi i)^{4} }
{ 1 \over p^2 (p-k_1)^2(p+k_4+k_5)^2
(p+k_5)^2 }
\el+s_{5}s_{1} \int_{\G_p}
{d^{4} p \over (2\pi i)^{4} }
{ 1 \over p^2 (p-k_1)^2 (p-k_1-k_2)^2
(p+k_5)^2 }+s_{1}s_{2} \int_{\G_p}
{d^{4} p \over (2\pi i)^{4} }
{ 1 \over p^2 (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
} \bigg) \,{\rm ,}\nn
\label{5ptaprxrhs}
\eea
\end{changemargin}
using a multidimensional generalization of Cauchy's residue theorem. Note that we have set $\e$ to zero. Very soon we will see that, on $\G_p$, eq. (\ref{5ptapprox}) is perfectly well-defined in $D = 4$. $\G_p$ is a product of four tiny circles that wrap all of the singularities of the integrand
$${ 1 \over p^2 (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
}$$
in $\mathbb{C}^4$ and, furthermore, fail to wrap all four of the singularities of any of the other four integrands. The so-called global residue theorem~\cite{GriffithsHarris} allows us to evaluate (\ref{5ptaprxrhs}) in an incredibly straightforward fashion. The first four terms in (\ref{5ptaprxrhs}) give zero contribution because a non-zero contribution can only arise if all singularities of the integrand are wrapped by the contour. The final term,
$$i{A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} s_{1}s_{2} \int_{\G_p}
{d^{4} p \over (2\pi i)^{4} }
{ 1 \over p^2 (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2
}\,{\rm ,}$$
evaluates to
\be
i{A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} s_{1}s_{2}\, {\rm det}^{-1}\left({\partial\over \partial p_\mu} ~(p-K_i)^2\right)\Bigg|_{p_\mu =\, p_\mu^*}\,{\rm ,}
\ee
where the new factor is just the Jacobian that results if one changes the integration variables from the $p_\mu$ to the propagator denominators, $(p-K_i)^2$, themselves. This Jacobian is evaluated on the solution of the four equations $(p^*-K_i)^2 = 0$. Thus, it appears that our strategy to isolate the integral coefficient $i s_1 s_2 A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}/2$ was almost successful. The only problem is that the definition of $\G_p$ does not specify a unique contour of integration; the system $(p^*-K_i)^2 = 0$ has two solutions, $p^{*\,(1)}$ and $p^{*\,(2)}$. The question of which contour is the ``right'' one to use is an important technical one that we return to later; it will be much easier to address this point once we've explained how to interpret the left-hand side of eq. (\ref{5ptapprox}) evaluated on $\G_p$.
The optical theorem for Feynman diagrams is usually presented in text books (see {\it e.g.}~\cite{PeskinSchroeder}) for individual Feynman diagrams. In its simplest incarnation, it relates the product of two tree-level diagrams integrated over an appropriate Lorentz invariant phase-space (of the external lines of the two tree amplitudes that depend on $p$) to the imaginary part of the one-loop Feynman diagram built by gluing together the two tree diagrams in the kinematic channel under consideration (called the channel ``being cut''). See Figure \ref{ordunitarity} for a cartoon depicting an $s$-channel cut of a diagram that would enter into the calculation of the one-loop virtual corrections to Bhabha scattering.\footnote{Technically speaking, a cut can be implemented by replacing of a set of propagators by delta functions that force the momenta carried by the replaced propagators on-shell.}
\FIGURE{
\resizebox{0.9\textwidth}{!}{\includegraphics{ordunitarity.eps}}
\vbox{\vskip 1 in}
\caption{An example of the optical theorem for Feynman diagrams for an $s$-channel vacuum polarization graph that contributes to the one-loop amplitude for Bhabha scattering. On the left-hand side we have twice the imaginary part of a one-loop Feynman diagram and on the right-hand side we have a product of two tree-level Feynman diagrams integrated over the Lorentz invariant phase-space of the cut propagators.}
\label{ordunitarity}}
From this point of view, the optical theorem is only useful as a cross-check on individual Feynman diagrams, not as a calculational tool. However, the generalized version of the optical theorem is much more powerful because it relates {\it on-shell} tree amplitudes integrated over an appropriate Lorentz invariant phase-space to the imaginary parts of pieces of complete one-loop amplitudes. Furthermore, one might hope that $\Nsym$ is a theory where the integral basis is such that the analytic structure of each basis element can be deduced from its imaginary part without any ambiguities. Indeed, as was shown in~\cite{BDDKMHV}, $\Nsym$ is in the class of so-called cut-constructible theories. One might guess that this is the case by looking at the explicit formulae of eqs. (\ref{box0})-(\ref{box4}).
To better understand this discussion, we introduce our variant of the optical theorem and apply it to the $I_4^{(5)}$ topology of $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$. Consider the complete set of Feynman diagrams that have gluons running in the loop\footnote{In the particular case of $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$, it turns out that, once internal lines are put on-shell, the $SU(4)_R$ symmetry does not allow non-zero contributions from fermions or scalars. This is a general phenomenon that will occur in pure-glue one-loop $\Nsym$ amplitudes whenever multiple external lines are attached to a corner of a generalized unitarity diagram and all of these lines are positive (or negative) helicity gluons.} and the topology of $I_4^{(5)}$. These diagrams are drawn in Figure \ref{BFGdiags}. It is worth pointing out that the set of Feynman diagrams in Figure \ref{BFGdiags} are only valid if we work in background field gauge. It was first understood in~\cite{BDDKMHV} that there are enormous practical advantages to working in background field gauge when one is faced with the task of computing a one-loop $\Nsym$ scattering amplitude in which all the external states are gluons. The reason for this is that the use of background field gauge reduces the degree of the loop momentum polynomial in the numerator of each Feynman integrand. In the calculation of an $n$-point scattering amplitude, one expects a loop momentum polynomial of degree $n-4$ in the numerator of an $n$-point contribution to the amplitude. For us, this means that the use of background field gauge will allow us to express $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$ in such a way that there will be at most one power of the loop momentum in any given term in the numerator of the Feynman integral with five propagator denominators (pentagon topology) and no powers of the loop momentum in the numerators of the five daughter integrals (the notion of daughter integral is explained in \ref{muint}) with four propagator denominators (box topology).
\FIGURE{
\resizebox{0.75\textwidth}{!}{\includegraphics{BFGdiags.eps}}
\vbox{\vskip 1 in}
\caption{Background field gauge Feynman diagrams that contribute to the $I_4^{(5)}$ topology of $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$.}
\label{BFGdiags}}
First let us think about the pentagon diagram. It is possible to exploit that fact that there is at most one power of the loop momentum in the numerator of the pentagon diagram to reduce the pentagon diagram to a sum over boxes. We can see this directly from results derived in~\cite{oneloopdimreg} truncated to $\Ord(\e^0)$. If we Feynman parametrize the loop momentum in the pentagon diagram using Feynman parameters $x_i$, the following two formulas can be used to write the pentagon diagram as a sum of scalar box integrals:
\bea
I_5 &=& {1\over 2} \sum_{j=1}^5 C_j I_{4}^{(j)} + \Ord(\e) \\
I_5[x_i] &=& {1\over 2}\sum_{j=1}^5 S_{i j}^{-1} I_{4}^{(j)} + \Ord(\e)
\eea
In \ref{muint} we derive these formulae and carefully define the functions $C_j$ and $S_{i j}$. The important point is that there is a piece of the pentagon diagram that has the same topology as the other two diagrams drawn in Figure \ref{BFGdiags} and it is this piece that should be grouped together with those diagrams.
At last, we are set up to explain the principle of generalized unitarity. Suppose we cut through, say, propagators one and three in the two true box diagrams and the relevant piece of the reduced pentagon. The claim is that, if we add up all three pieces, the fact that they share the same topology (and we have added up all possible contributions) guarantees that the sum will be a product of two on-shell tree amplitudes, one with external momenta $-p$, $k_1$, $k_2$, and $p-k_1-k_2$ and one with external momenta $-p+k_1+k_2$, $k_3$, $k_4$, $k_5$ and $p$ , integrated over the appropriate Lorentz invariant phase-space. Since $\Nsym$ is cut-constructible, we can easily invert this process and deduce complete one-loop integrands of a particular topology by calculating two appropriate on-shell tree amplitudes and tacking on a couple of missing propagator denominators. In fact, this discussion motivates replacing Feynman diagrams completely in favor of what we'll call generalized unitarity diagrams. The generalized unitarity diagram (singular) associated to the discussion of this paragraph is drawn in Figure \ref{I45GUdiag}.
\FIGURE{
\resizebox{0.65\textwidth}{!}{\includegraphics{I45GUdiag.eps}}
\vbox{\vskip 1 in}
\caption{The $s_1$-channel generalized unitarity diagram can be used to reconstruct the coefficient of the $I_4^{(5)}$ topology of $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$ but the information about $I_4^{(5)}$ will have to be disentangled from that pertaining to all the other boxes detected by the cut (all except $I_1^{(1)}$ and $I_1^{(3)}$). The blobs on either side of the lines carrying momenta $p$ and $p - k_1 - k_2$ (the propagators to be cut in evaluating the diagram) represent on-shell tree amplitudes.}
\label{I45GUdiag}}
Using generalized unitarity with double cuts is already very powerful, but we want to do even better. We would like to use the principle of generalized unitarity in a way that meshes well with our earlier discussion of multidimensional contour integrals and the right-hand side of eq. (\ref{5ptapprox}). What this amounts to is cutting all the propagators in all contributions with a particular box topology. For the five-point example of this subsection, we would end up with the diagram of Figure \ref{I45maxGU}.
\FIGURE{
\resizebox{0.75\textwidth}{!}{\includegraphics{I45maxGU.eps}}
\vbox{\vskip 1 in}
\caption{The maximal generalized unitarity diagram that contributes to the $I_4^{(5)}$ topology of $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$. As before, each blob represents an on-shell tree amplitude and each line in the diagram is to be cut in the evaluation of the diagram. This time, however, the cuts {\it only} detect $I_4^{(5)}$.}
\label{I45maxGU}}
In other words, we can reconstruct the integrand of the piece of the one-loop amplitude $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$ (with gluons running in the loop) and topology $I_4^{(5)}$ by first multiplying four appropriate tree amplitudes together (three three-point amplitudes and one four-point amplitude) and then tacking on the missing propagator denominators:
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_4^{(5)}} = \int {d^{4} p \over (2\pi i)^{4}} A^{tree}\left(-p, k_1^{1234}, (p-k_1)^{1234} \right)
\el A^{tree}\left((-p+k_1), k_2^{1234}, p-k_1-k_2\right) A^{tree}\left((-p+k_1+k_2)^{1234}, k_3, p+k_4+k_5\right) \times
\el \times A^{tree}\left((-p-k_4-k_5)^{1234}, k_4, k_5, p^{1234}\right){1\over p^2 (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2}+
\el
\el
\el \int {d^{4} p \over (2\pi i)^{4}} A^{tree}\left(-p, k_1^{1234}, p-k_1 \right)A^{tree}\left((-p+k_1)^{1234}, k_2^{1234}, p-k_1-k_2\right)
\el A^{tree}\left((-p+k_1+k_2)^{1234}, k_3, p+k_4+k_5\right) A^{tree}\left((-p-k_4-k_5)^{1234}, k_4, k_5, p^{1234}\right)\times
\el\times{1\over p^2 (p-k_1)^2 (p-k_1-k_2)^2 (p+k_4+k_5)^2} \nn
\label{5pt45piece}
\eea
\end{changemargin}
In this example, there are just two consistent assignments of the internal $SU(4)_R$ indices that would not obviously give a vanishing result by virtue of the SUSY Ward identities $\a(k_1,k_2,\cdots,k_n) = 0$ and $\a(k_1^{1234},k_2,\cdots,k_n) = 0$ or their parity conjugates\footnote{Parity acts on scattering amplitudes as complex conjugation (interchange of angle and square brackets in spinor products).}. In general, one should sum over all the internal configurations allowed by the $SU(4)_R$ symmetry.\footnote{In component language, what we mean is that, usually, the four Majorana fermions running in the loop and the three complex scalars running in the loop would give non-vanishing contributions.} There is, however, a more subtle constraint which forces the first term in eq. (\ref{5pt45piece}) to zero. It turns out that when there are configurations with two on-shell three-point amplitudes next to each other, the adjacent three-point amplitudes must have different $SU(4)_R$ index structures or they vanish~\cite{BDKrev}. Therefore, the product
\bea
&&A^{tree}\left(-p, k_1^{1234}, (p-k_1)^{1234} \right)A^{tree}\left((-p+k_1), k_2^{1234}, p-k_1-k_2\right)
\el A^{tree}\left((-p+k_1+k_2)^{1234}, k_3, p+k_4+k_5\right) A^{tree}\left((-p-k_4-k_5)^{1234}, k_4, k_5, p^{1234}\right) \nonumber
\eea
must vanish because $A^{tree}\left((-p+k_1), k_2^{1234}, p-k_1-k_2\right)$ and $A^{tree}\left((-p+k_1+k_2)^{1234}, k_3, p+k_4+k_5\right)$ have the same $SU(4)_R$ index structure.
All the hard work is now done and we can evaluate the reconstructed one-loop integrand of eq. (\ref{5pt45piece}) as a contour integral (with $\G_p$ as the contour) in exactly the same way that we evaluated the right-hand side of eq. (\ref{5ptaprxrhs}). We find that eq. (\ref{5pt45piece}) gives
\bea
&&A^{tree}\left(-p, k_1^{1234}, p-k_1 \right)A^{tree}\left((-p+k_1)^{1234}, k_2^{1234}, p-k_1-k_2\right)\times
\el \times A^{tree}\left((-p+k_1+k_2)^{1234}, k_3, p+k_4+k_5\right) A^{tree}\left((-p-k_4-k_5)^{1234}, k_4, k_5, p^{1234}\right)\times
\el\times {\rm det}^{-1}\left({\partial\over \partial p_\mu} ~(p-K_i)^2\right)\Bigg|_{p_\mu = \,p_\mu^*} \,{\rm ,}
\label{lhsLS}
\eea
where $p^{*\,(i)}$ represents either of the two solutions of $(p^*-K_i)^2 = 0$. Finally, we can put together the left- and right-hand sides of (\ref{5ptapprox}) evaluated on $\G_p$:
\bea
&&A^{tree}\left(-p, k_1^{1234}, p-k_1 \right)A^{tree}\left((-p+k_1)^{1234}, k_2^{1234}, p-k_1-k_2\right)\times
\el \times A^{tree}\left((-p+k_1+k_2)^{1234}, k_3, p+k_4+k_5\right) A^{tree}\left((-p-k_4-k_5)^{1234}, k_4, k_5, p^{1234}\right)\Bigg|_{p_\mu = p_\mu^*}
\elale i {A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} s_{1}s_{2}\,{\rm ,}
\label{LSsketch}
\eea
where we have cancelled a Jacobian factor from both sides. The only loose end to tie up is what to do about the fact that there are actual two distinct contours specified by $\G_p$. There is no natural reason to chose one solution of $(p^*-K_i)^2 = 0$ over the other and, to make matters worse, it appears that the equations defined by (\ref{LSsketch}) are not consistent. On one solution to $(p^*-K_i)^2 = 0$, $p^{*\,(1)} = k_1^\mu + 1/2\,\spa2.3 \spab1.{\gamma^\mu}.2/\spa1.3 $, the product of trees on the left-hand side of (\ref{LSsketch}) is $i s_1 s_2 A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}$ and on the other solution, $p^{*\,(2)} = k_1^\mu + 1/2\,\spb2.3 \spab2.{\gamma^\mu}.1/\spb1.3$, the product of trees is 0. To summarize, our two equations read
\bea
0 &=& i {s_1 s_2\over 2} A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}
\label{LS51}\\
i s_1 s_2 A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} &=& i {s_1 s_2\over 2} A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}
\label{LS52}
\eea
The reason for the lack of consistency is that we truncated our basis at $\Ord(\e^0)$. As we will see in the next section, adding a massless pentagon integral to the basis of Feynman integrals fixes this apparent inconsistency in a very nice way. For now, we appeal to symmetry to fix our problem~\cite{SharpLS}. Our two solutions, $p^{*\,(1)}$ and $p^{*\,(2)}$, are related by parity symmetry (complex conjugation of the spinor products). We want our final answer to be parity invariant and one way to ensure this is to simply add eqs. (\ref{LS51}) and (\ref{LS52}). If we do this we find that the left-hand side is equal to the right-hand side, which implies that this procedure works, at least for $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$. It turns out that this ad hoc prescription will work for general one-loop amplitudes if one doesn't care about the higher order in $\e$ pieces of the amplitude. For general multi-loop amplitudes, however, such sloppy analysis is simply not sufficient. We will also need something better for our all-orders-in-$\e$ computations at the one-loop level. It is to this that we turn in the next subsection.
\subsection{Generalized Unitarity in $D$ Dimensions}
\label{GUD}
The great thing about generalized unitarity is that it works in very general situations (the unitarity of the S-matrix is a consequence of probability conservation in quantum mechanics). In particular, generalized unitarity is compatible with dimensional regularization because dimensional regularization preserves unitarity. As was shown shortly after the seminal papers on the generalized unitarity technique were published, there is no inherent restriction to $\Ord(\e^0)$; if desired, one can compute amplitudes to all orders in $\e$ by working a little harder.~\cite{BernMorgan} We explain how this works in the context of the example of the last subsection and, in the process, explain how one needs to modify the basis of planar one-loop box integrals used so far if one is interested in computing planar $\Nsym$ amplitudes to all orders in the dimensional regularization parameter.
At the five-point level, there is an obvious candidate integral that one could try adding to the basis of scalar boxes: the massless pentagon. Let us try to prove that there is a non-vanishing pentagon contribution to $A^{1-{\rm loop}}_1(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$ that we ignored in the last subsection. Our experience in \ref{GU4} has taught us that it is a bad idea to try and think of the pentagon integrals in the problem as reductive box contributions from the start. Rather, we should make the most general ansatz of scalar basis integrals that makes sense and let the amplitude decide how it wants to be written. To this end, we perform the same generalized unitarity analysis on the left-hand side of (\ref{5ptapprox}) that we did in the last subsection but this time make an ansatz $\sum_i A_i \,I_4^{(i)}+B\,I_5$ for the right-hand side. As before, we can get a non-zero contribution from $A_4 I_4^{(5)}$ but now we will also have a non-zero contribution from $B\,I_5$:
\bea
&&B \int_{\G_p} {d^4p\over (2 \pi i)^4} {1\over p^2 (p - k_1)^2 (p - k_1 - k_2)^2 (p + k_4 + k_5)^2 (p + k_5)^2 }
\elale B\,{1\over (p + k_5)^2} {\rm det}^{-1}\left({\partial\over \partial p_\mu} ~(p-K_i)^2\right)\Bigg|_{p_\mu = \,p_\mu^*} \,{\rm .}
\eea
This makes all the difference. Instead of eqs. (\ref{LS51}) and (\ref{LS52}) we now have the system
\bea
0 &=& A_4 + B\,{1\over (p^{*\,(1)} + k_5)^2}
\label{LS5D1}\\
i s_1 s_2 A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} &=& A_4 + B\,{1\over (p^{*\,(2)} + k_5)^2}
\label{LS5D2}
\eea
which is consistent and solvable. We will follow~\cite{SharpLS} and refer to this technique as the leading singularity method (the leading singularities being the left-hand sides of the above equations). We find
\begin{equation}
A_4 = -i A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}
\frac{s_{1}s_{2}\tilde\beta_5}{\beta_5-\tilde\beta_5}, \qquad B =
i A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}\frac{s_{5}s_{1}s_{2}}{\beta_5-\tilde\beta_5} \,{\rm ,}
\label{LS5ptfin}
\end{equation}
where
\begin{equation}
\beta_{5} = \left(1+ \frac{\langle 2 3\rangle [2 5] }{\langle
1 3\rangle [1 5]} \right)^{-1}, \qquad \tilde\beta_{5} =
\left(1+ \frac{\langle 2 5\rangle [2 3] }{\langle 1 5\rangle
[1 3]} \right)^{-1}\,{\rm .}
\end{equation}
The formula for $A_4$ bears no resemblance to $i s_1 s_2 A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}/2$ and, indeed, one can check numerically (using {\it e.g.} S@M~\cite{S@M}) that they are not equal. At first sight, it appears that the leading singularity method fails to reproduce the known result. In actuality there is no contradiction because, secretly, the result obtained in the previous subsection was expressed in terms of a different basis with $4 - 2\e$ dimensional boxes and $6 - 2\e$ dimensional pentagon integrals as opposed to the $4-2\e$ dimensional box and pentagon integrals we used above. The connection between these two bases is the reduction formula derived in \ref{muint} using traditional techniques:\footnote{Note that eq. (\ref{PtoDCB}) employs the original conventions of definition (\ref{IntDef}).}
\be
I_5^{D=4-2\e} = {1\over 2}\bigg[\sum_{j=1}^5 C_j I_{4}^{(j),\,D=4-2\e} + 2\e C_0 I_5^{D=6-2 \e} \bigg]\,{\rm .}
\label{PtoDCB}
\ee
The answer looks much more compact when expressed in this basis (employed in the last subsection); the higher order in $\e$ terms are more cleanly separated from those that are present through $\Ord(\e^0)$. This is related to the fact that there will always be an explicit $\e$ out front of $I_5^{D=6-2 \e}$ and $I_5^{D=6-2 \e}$ is both UV and IR finite~\cite{oneloopdimreg}. In this paper we will refer to the basis with all elements expanded about $D = 4$ as the geometric basis and the basis with $I_5^{D=6-2 \e}$ pentagons as the dual conformal basis (this notation will be motivated in Section \ref{WL/MHV}). Simplifying (\ref{LS5ptfin}) after projecting the geometric basis onto the dual conformal basis using (\ref{PtoDCB}), we find that
\be
\tilde{A}_4 = -{s_1 s_2\over 2} A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}
\label{A45ptDCB}
\ee
as before and
\be
\tilde{B} = \e ~\pol(1,2,3,4) A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}\,{\rm ,}
\label{B5ptDCB}
\ee
where we have made the useful definition
\be
\pol(i,j,m,n) \equiv 4 i \pol_{\mu \nu \rho \sigma} k_i^\mu k_j^\nu k_m^\rho k_n^\sigma = \spb{i}.j \spa{j}.m \spb{m}.n \spa{n}.i - \spa{i}.j \spb{j}.m \spa{m}.n \spb{n}.i \, {\rm .}
\label{epstensor}
\ee
In above eq. (\ref{A45ptDCB}) is the coefficient of $I_4^{(5),\,D = 4-2\e}$ and eq. (\ref{B5ptDCB}) is the coefficient of $I_5^{D=6-2\e}$ in the conventions of definition (\ref{IntDef}).
What we have learned is that, in the case of $A^{1-{\rm loop}}_1(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$, we are able to learn all about the higher-order in $\e$ pieces of the amplitude without ever leaving four dimensions. It would be great if the leading singularity method gave us all of the pentagon coefficients for arbitrary $n$ but, unfortunately, life is not so simple. In fact, as we shall see in the next subsection, one needs to develop more machinery to calculate the pentagon integrals already at the six-point level. In a nutshell, what we need to do is to further develop the $D$ dimensional unitarity technique of Bern and Morgan~\cite{BernMorgan} to reconstruct one-loop integrands in $\Nsym$ without dropping any higher order in $\e$ pieces.
We now review the Bern-Morgan approach to $D$-dimensional integrand reconstruction to prepare the reader for the next section where we discuss simple extensions of their results. As an illustration, we consider the amplitude $A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4)$ in pure Yang-Mills theory. Following~\cite{BernMorgan}, we remind the reader of the second form for $I_4^{D = 4 - 2\e}$ where we split up the integral over the loop momentum into four dimensional and $-2\e$ dimensional pieces:
\be
I_4^{D=4-2 \e} = -i (4\pi)^{2-\e} \int
{d^{4} p \over (2\pi)^{4} }
{d^{-2\e} \mu \over (2\pi)^{-2\e} }
{ 1 \over (p^2 - \mu^2) ((p-k_1)^2 - \mu^2) ((p-k_1-k_2)^2 - \mu^2)(
(p+k_4)^2 - \mu^2) } \,{\rm .}\nn
\ee
If we consider an $s$-channel cut of the above zero mass box integral, we find the on-shell conditions
\be
p^2 = \mu^2 \qquad (p-k_1-k_2)^2 = \mu^2\,{\rm .}
\ee
It follows that, to reconstruct the complete one-loop integrand in $D$ dimensions using the principle of generalized unitarity, one should simply imagine that the lines of the tree amplitudes on either side of the unitarity cut(s) (external lines of the trees that have $p$-dependent momenta) have a mass $\mu$. Actually, the procedure of gluing trees together to form loops is a little more complicated in our approach because we do not have in hand an analog of the spinor helicity framework in $-2\e$ dimensions\footnote{It is possible that, with a bit of inspiration, we might be able to profitably make use of some combination of the formalisms worked out in~\cite{Dennen,CheungC,BCDHI}.}. Consequently, the whole process is more closely related to traditional perturbation theory. In particular, summing over internal degrees of freedom inside the loop being reconstructed is much more labor intensive than it is in four dimensions. One trick to try and avoid tedious algebra, which works better in some situations than in others, is to perform a supersymmetric decomposition of the amplitude. For example, if we rewrite a loop of gluons in the following way:
$$A_g = (A_g + 4 A_f + 3 A_s) - 4 (A_f + A_s) + A_s$$
We see that the contribution from a loop of gluons ({\it i.e.} pure Yang-Mills theory) can be derived by summing the answer in $\Nsym$ and the contribution from a loop of complex scalars and then subtracting off the contribution from four $\mathcal{N}=1$ chiral multiplets. For the present application this works beautifully because the first two terms on the right-hand side of the above equation are protected by supersymmetry and vanish (see Appendix \ref{SWI}). It follows that
\be
A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4) = A^{1-{\rm loop}}_{1;\,{\rm scalar}}(k_1,k_2,k_3,k_4)
\label{all+susydecomp}
\ee
and, in this particular case, we can avoid some numerator algebra by calculating $A^{1-{\rm loop}}_{1;\,{\rm scalar}}(k_1,k_2,k_3,k_4)$ instead of $A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4)$.
Generalized unitarity applied to $A^{1-{\rm loop}}_{1;\,{\rm scalar}}(k_1,k_2,k_3,k_4)$ gives\footnote{In what follows, we will very often be interested in amplitudes where some of the external states have definite helicity and some should be thought of as having any of the possible physical polarizations. We label external states with indeterminate polarization as $(q)_{x}$ where $q$ is the momentum carried by the external particle and $x$ denotes the particle type ($s$ or $\bar{s}$ for scalar states, $f$ or $\bar{f}$ for fermion states, and $g$ for gluon states).}
\bea
&&{1\over (4 \pi)^{2-\e}} A^{1-{\rm loop}}_{1;\,{\rm scalar}}(k_1,k_2,k_3,k_4) = \int
{d^{4} p \over (2\pi)^{4} }
{d^{-2\e} \mu \over (2\pi)^{-2\e} }
\bigg({ i \over p^2 - \mu^2} A^{tree}_{\mu^2}\left((-p)_{s},k_1,k_2,(p-k_1-k_2)_{\bar{s}}\right)
\el{i\over(p-k_1-k_2)^2 - \mu^2} A^{tree}_{\mu^2}\left((-p+k_1+k_2)_s,k_3,k_4,p_{\bar{s}}\right)
\el+ { i \over p^2 - \mu^2} A^{tree}_{\mu^2}\left((-p)_{\bar{s}},k_1,k_2,(p-k_1-k_2)_s\right) {i\over(p-k_1-k_2)^2 - \mu^2} A^{tree}_{\mu^2}\left((-p+k_1+k_2)_{\bar{s}},k_3,k_4,p_s\right)\bigg) \,{\rm .}\nn
\eea
The massive scalar amplitudes $A^{tree}_{\mu^2}\left((-p)_s,k_1,k_2,(p-k_1-k_2)_{\bar{s}}\right)$ and $A^{tree}_{\mu^2}\left((-p)_{\bar{s}},k_1,k_2,(p-k_1-k_2)_s\right)$ are equal, as are $A^{tree}_{\mu^2}\left((-p+k_1+k_2)_s,k_3,k_4,p_{\bar{s}}\right)$ and $A^{tree}_{\mu^2}\left((-p+k_1+k_2)_{\bar{s}},k_3,k_4,p_s\right)$. Using
\bea
A^{tree}_{\mu^2}\left((-p)_s,k_1,k_2,(p-k_1-k_2)_{\bar{s}}\right) &=& {i \mu^2 \spb1.2 \over \spa1.2 ((p-k_1)^2-\mu^2)}~~~~{\rm and} \\
A^{tree}_{\mu^2}\left((-p+k_1+k_2)_s,k_3,k_4,p_{\bar{s}}\right) &=& {i \mu^2 \spb3.4 \over \spa3.4 ((p+k_4)^2-\mu^2)}\,{\rm ,}
\eea
which can be derived from Feynman diagrams, we find
\bea
&&{1\over (4 \pi)^{2-\e}} A^{1-{\rm loop}}_{1;\,{\rm scalar}}(k_1,k_2,k_3,k_4) = {1\over (4 \pi)^{2-\e}} {2 \spb1.2 \spb3.4\over \spa1.2 \spa3.4}\times
\el \times \int
{d^{4} p \over (2\pi)^{4} }
{d^{-2\e} \mu \over (2\pi)^{-2\e} }
{ \mu^4 \over (p^2 - \mu^2)((p-k_1)^2-\mu^2)((p-k_1-k_2)^2 - \mu^2)((p+k_4)^2-\mu^2)} \nn
&&= {1\over (4 \pi)^{2-\e}} {2 i \spb1.2 \spb3.4\over \spa1.2 \spa3.4} I_4^{D = 4 - 2\e}[\mu^4] \\
A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4) &&= A^{1-{\rm loop}}_{1;\,{\rm scalar}}(k_1,k_2,k_3,k_4) = {2 i \spb1.2 \spb3.4\over \spa1.2 \spa3.4} I_4^{D = 4 - 2\e}[\mu^4]\,{\rm .}
\eea
A basis integral with some power of $\mu^2$ inserted in the numerator is usually referred to as a $\mu$-integral and such terms will play a central role in this work. It is often convenient to rewrite $\mu$-integrals in terms of dimensionally shifted integrals. This is easily accomplished by manipulating the $-2\e$ dimensional part of the integration measure in eq. (\ref{IntDef}). Written out, the $-2\e$ dimensional integral is
\be
\int {d^{-2\e} \mu \over (2\pi)^{-2\e} } f(\mu^2) = \int{d \Omega_{-2\e}\over (2\pi)^{-2\e}}\int_0^\infty d\mu \mu^{-2\e-1}f(\mu^2) = {1\over 2} \int {d \Omega_{-2\e}\over (2\pi)^{-2\e}}\int_0^\infty d\mu^2 (\mu^2)^{-\e-1}f(\mu^2)\, {\rm ,}
\ee
where, as usual,
\be
\int {d \Omega_{-2\e}} = {2 \pi^{-\e} \over \G(-\e)}\,{\rm .}
\ee
Now, if we replace $f(\mu^2)$ with $\mu^{2 r}$ we can absorb the extract factors of $\mu^2$ into the integration measure:
\begin{changemargin}{-.6 in}{0 in}
\be
\int {d^{-2\e} \mu \over (2\pi)^{-2\e} } \mu^{2 r}f(\mu^2) = {(2\pi)^{2 r}\int d\Omega_{-2\e}\over \int d\Omega_{2 r -2\e}} \int {d^{2 r-2\e}\mu \over (2 \pi)^{2 r-2 \e}} f(\mu^2) = -\e(1-\e)(2-\e)\cdots(r-1-\e)(4\pi)^r\int {d^{2 r-2\e}\mu \over (2 \pi)^{2 r-2 \e}} f(\mu^2)\,{\rm .}
\ee
\end{changemargin}
If $r$ is a natural number, this analysis leads to
\be
I_n^{D = 4 - 2\e}[\mu^{2 r}] = -\e(1-\e)(2-\e)\cdots(r-1-\e)I_n^{D = 2 r + 4 - 2\e}
\label{DSmu}
\ee
relating $\mu$-integrals and dimensionally-shifted integrals. Now, a very interesting phenomenon can occur, which we illustrate by applying eq. (\ref{DSmu}) to our result for $A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4)$. We first rewrite the answer
\be
A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4) = {2 i \spb1.2 \spb3.4\over \spa1.2 \spa3.4} I_4^{D = 4 - 2\e}[\mu^4] = -{2 \e (1-\e)i \spb1.2 \spb3.4\over \spa1.2 \spa3.4} I_4^{D = 8 - 2\e}
\ee
and then Feynman parametrize it:
\be
A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4) = -{2 \e (1-\e)i \spb1.2 \spb3.4\over \spa1.2 \spa3.4} \G(\e)\int_0^1 dx\int_0^{1-x}dy\int_0^{1-x-y}dz {1\over \mathcal{D}(x,y,z)^\e}\,{\rm .}
\ee
Remarkably, the $\e$ expansion of the above starts at $\Ord(\e^0)$. Explicitly, we find
\bea
A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4) &=& -{2 i \spb1.2 \spb3.4\over \spa1.2 \spa3.4} \int_0^1 dx\int_0^{1-x}dy\int_0^{1-x-y}dz + \Ord(\e)
\elale -{i \spb1.2 \spb3.4\over 3 \spa1.2 \spa3.4}+ \Ord(\e)\,{\rm .}
\eea
At first sight, this result might seem rather puzzling since, without the $\mu^4$ in the numerator, the integral $I_4^{D = 4 -2 \e}$ is UV finite and IR divergent. What has happened is that, in shifting to $D = 8 - 2\e$, we have induced a UV divergence (the integral now has the same number of powers of the loop momenta in the measure of integration as it has in the denominator) and the IR divergences effectively got regulated by the $\mu^2$ factors in the propagator denominators. The explicit $\e$ in the numerator coming from eq. (\ref{DSmu}) is canceling the induced UV pole, which is why the $\e$ expansion of $A^{1-{\rm loop}}_{1;\,\mathcal{N}=0}(k_1,k_2,k_3,k_4)$ starts at $\Ord(\e^0)$.
Although we have been focusing on scalars running in the loop we could equally well have performed the above analysis for a loop of fermions with one obvious additional complication: the need to sum over internal spin states in a Lorentz covariant way. Typical tree amplitudes with a pair of massive fermions will be built out of a string beginning with $\bar{u}^{\pm}(p)$ and ending with $u^\pm(p)$. In order to fuse together two such tree amplitudes across a unitarity cut, we simply use the spin sum identity
\be
\sum_{s} u^s(p)\bar{u}^s(p) = \slashed{p} + \mu
\label{fermspsum}
\ee
heavily used in traditional perturbation theory~\cite{PeskinSchroeder}. In \ref{effgcomp}, we treat a gluon running in the loop as well. Due to the fact there is no straightforward massive counterpart (with two spin states) to the massless gluon, treating an internal gluon line requires a little more thinking. In the last few years, $D$-dimensional unitarity has been systematized by several different groups~\cite{Rocket1,Rocket2,BadgerGUD,Ossola}.
Also, we wish to remark that there is no reason for us to restrict ourselves to double cuts; as we shall see in the next section, we can profit enormously by using quintuple cuts in $D$ dimensions to determine individual pentagon coefficients one at a time. The idea is conceptually similar to what we did with quadruple cuts and box coefficients in \ref{GU4}, though it is a bit more complicated. It turns out that the leading singularity method supplemented by $D$ dimensional quintuple cuts allows one to efficiently calculate all-orders-in-$\e$ one-loop $\Nsym$ amplitudes.
\section{Efficient Computation and New Results For One-Loop $\Nsym$ Gluon Amplitudes Calculated To All Orders in $\e$}
\label{gluoncomp}
\subsection{Efficient Computation Via $D$ Dimensional Generalized Unitarity}
\label{effgcomp}
In order to harness the power of $D$-dimensional unitarity for the application at hand, we have to extend the results of Bern and Morgan to treat cut internal gluon lines. To be clear, many other authors have thought about extending the Bern-Morgan approach to integrand reconstruction (see {\it e.g.}~\cite{Rocket1,Rocket2,BadgerGUD,Ossola}). All of them either focus on a getting numerical results or isolate terms that would be missed by four dimensional generalized unitarity. There are obviously many applications where it makes sense to follow one of these strategies. In this paper, however, we have a different goal. We further develop the Bern-Morgan approach and show that it is a very efficient way to analytically reconstruct general one-loop integrands. In fact, we expect that our approach will mesh well with the spinor integration reduction technique of~\cite{ABFKM,BFY}, which is applicable to general field theory amplitudes at one-loop. Although these references analyzed a variety of processes, they started with integrands obtained by other means in all cases except that of a complex scalar running in the loop. A general strategy for the analytical reconstruction of one-loop integrands in $D$ dimensions was not discussed. In what follows we fill in this gap.
Now, to illustrate our approach to $D$-dimensional unitarity, we offer an alternative derivation of the massless pentagon coefficient of eq. (\ref{B5ptDCB}) associated to \\$A^{1-{\rm loop}}_1(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$. All we really need to do right now is extend Bern-Morgan to the case of purely gluonic external states with a massless vector running in the loop. Later on we will also treat the case where some of the external gluons are replaced by fermions. It seems likely that so far most researchers have found it expedient to side-step the question of how to properly treat a gluon running in the loop by exploiting supersymmetry decompositions as was done in Subsection \ref{GUD}. We argue that it is no more difficult to calculate directly.
We warm up by repeating the analysis of the last subsection, but for the pentagon coefficient of $A^{1-{\rm loop}}_1(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$ using quintuple cuts. Using the massive scalar three-point vertices~\cite{BGKSmassive},
\bea
A^{tree}_{\mu^2}\left((-p)_s,k_1,(p-k_1)_{\bar{s}}\right) &&= -i \sqrt{2} p \cdot \pol^+(k_1)~~~~{\rm and}
\\ A^{tree}_{\mu^2}\left((-p)_s,k_1^{1234},(p-k_1)_{\bar{s}}\right) &&= -i \sqrt{2} p \cdot \pol^-(k_1){\rm ,}
\eea
and quintuple $D$ dimensional generalized unitarity cuts we can deduce the pentagon integral coefficient for the scalar loop contribution to the five-point MHV amplitude. In the above, the polarization vectors can be evaluated using eqs. (\ref{polvecs+}) and (\ref{polvecs-}) because we are implicitly using the four dimensional helicity scheme (see \ref{4DHS}) where the external polarization vectors are kept in four dimensions. The result of this calculation is
\begin{changemargin}{-.4 in}{0 in}
\bea
&&A_{1;\,{\rm scalar}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5} = A^{tree}_{\mu^2}\left((-p_*)_s,k_1^{1234},(p_*-k_1)_{\bar{s}}\right) \times
\el \times A^{tree}_{\mu^2}\left((-p_*+k_1)_s,k_2^{1234},(p_*-k_1-k_2)_{\bar{s}}\right) A^{tree}_{\mu^2}\left((-p_*+k_1+k_2)_s,k_3,(p_*+k_4+k_5)_{\bar{s}}\right)\times
\el \times A^{tree}_{\mu^2}\left((-p_*-k_4-k_5)_s,k_4,(p_*+k_5)_{\bar{s}}\right)A^{tree}_{\mu^2}\left((-p_*-k_5)_s,k_5,(p_*)_{\bar{s}}\right) \,{\rm ,}\nn
\label{s5ptpent}
\eea
\end{changemargin}
where $p_*^\nu$ solves the on-shell conditions:
\cmb{-.6 in}{0 in}
\bea
p_*^2 - \mu^2 = 0 \qquad (p_*-k_1)^2 - \mu^2 &=& 0 \qquad (p_*-k_1-k_2)^2 - \mu^2 = 0 \nonumber \\
(p_*+k_4+k_5)^2 - \mu^2 = 0 & \qquad & (p_*+k_5)^2 - \mu^2 = 0 \,{\rm .}
\eea
\cme
It turns out that, in this case, the solution is unique and is given by~\cite{BFY} expanding the four dimensional, massive loop momentum with respect to a basis $K_1$, $K_2$, $K_3$, and $K_4$ of four-vectors:
\be
p^\nu = L_1 K_1^\nu + L_2 K_2^\nu + L_3 K_3^\nu + L_4 K_4^\nu
\label{5ptos1}
\ee
and then solving a system of linear equations for the $L_i$ coefficients. It makes sense to choose the $K$'s to be the four-vectors in the problem; in the present example we set
\be
K_1 = k_1 + k_2 \qquad K_2 = k_1 \qquad K_3 = -k_4-k_5 \qquad K_4 = -k_5{\rm .}
\ee
Explicitly, we have
\be
\left(\begin{array}{c}L_1\\L_2\\L_3\\L_4\end{array}\right) = {1\over2}\left(\begin{array}{cccc}K_1^2 & K_1\cdot K_2 & K_1\cdot K_3 & K_1 \cdot K_4 \\K_2\cdot K_1 & K_2^2 & K_2\cdot K_3 & K_2 \cdot K_4\\K_3\cdot K_1 & K_3\cdot K_2 & K_3^2 & K_3 \cdot K_4\\K_4\cdot K_1 & K_4\cdot K_2 & K_4\cdot K_3 & K_4^2\end{array}\right)^{-1} \left(\begin{array}{c}K_1^2\\K_2^2\\K_3^2\\K_4^2\end{array}\right)\,{\rm .}
\label{5ptos2}
\ee
Now that we are warmed up, we are ready to try the quintuple cut of the fermion loop contribution. The only reason that the fermion loop contribution is more complicated is that we have to sum over internal fermion spin states using eq. (\ref{fermspsum}); the net result of the sum over internal states for the scalar loop contribution is just an overall factor of two. Although Bern and Morgan did not literally give their fermions a mass $\mu$, our procedure is easily deduced from the discussion in their paper~\cite{BernMorgan}.
To reconstruct the one-loop integrand, we need tree amplitudes with two massive fermions and a gluon:
\bea
A^{tree}_{\mu^2}\left(p_{\bar{f}},k_1,(-p-k_1)_f\right) &=& -{i \over \sqrt{2}}\bar{u}(p)\slashed{\pol}^+(k_1) u(p+k_1)\\
A^{tree}_{\mu^2}\left(p_{\bar{f}},k_1^{1234},(-p-k_1)_f\right) &=& -{i \over \sqrt{2}}\bar{u}(p)\slashed{\pol}^-(k_1) u(p+k_1)
\eea
where we don't worry about specifying the spins of the fermions because we will ultimately sum over them using (\ref{fermspsum}). For the quintuple cut of the fermion loop we find
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1;\,{\rm fermion}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5} = - \left(-{i\over \sqrt{2}}\right)^5 \bar{u}(p_*)\slashed{\pol}^+(k_5) u(p_*+k_5) \bar{u}(p_*+k_5)\slashed{\pol}^+(k_4) u(p_*+k_4+k_5)
\el \bar{u}(p_*+k_4+k_5)\slashed{\pol}^+(k_3) u(p_*-k_1-k_2) \bar{u}(p_*-k_1-k_2)\slashed{\pol}^-(k_2) u(p_*-k_1)\bar{u}(p_*-k_1)\slashed{\pol}^-(k_1) u(p_*)
\elale - \left({i\over \sqrt{2}}\right)^5 {\Tr}\Big[\slashed{\pol}^+(k_5)(\slashed{p}_*+\slashed{k}_5 + \mu)\slashed{\pol}^+(k_4) (\slashed{p}_*+\slashed{k}_4+\slashed{k}_5 + \mu)
\el \slashed{\pol}^+(k_3) (\slashed{p}_*-\slashed{k}_1-\slashed{k}_2 + \mu)\slashed{\pol}^-(k_2) (\slashed{p}_*-\slashed{k}_1 + \mu)\slashed{\pol}^-(k_1) (\slashed{p}_* + \mu) \Big] \,{\rm .}\nn
\label{f5ptpent}
\eea
\end{changemargin}
In this context, the extra overall minus sign is a result~\cite{BernMorgan} of using three-point amplitudes with spinor strings of the form $\bar{u}(p)\slashed{\pol}^+(k_1) u(p+k_1)$, when really they should have spinor strings of the form $\bar{u}(p)\slashed{\pol}^+(k_1) u(-p-k_1)$. Now that we understand how to deal with a loop of fermions, it is natural to ask what the analogous prescription is for a loop of gluons. Clearly, to start we need to write down three-point gluon amplitudes
\bea
A^{tree}_{\mu^2}\left(-p_{g},k_1,(p-k_1)_g\right) &=& i \sqrt{2}\left(\pol^+(k_1)\cdot p~ g_{\rho \sigma}+k_{1\,\rho}~\pol^+_\sigma(k_1)-k_{1\,\sigma}~ \pol^+_\rho(k_1)\right)\pol^{*\,\rho}(p)\pol^\sigma(p-k_1)\nn
\\
A^{tree}_{\mu^2}\left(-p_{g},k_1^{1234},(p-k_1)_g\right) &=& i \sqrt{2}\left(\pol^-(k_1)\cdot p~ g_{\rho \sigma}+k_{1\,\rho}~\pol^-_\sigma(k_1)-k_{1\,\sigma}~ \pol^-_\rho(k_1)\right)\pol^{*\,\rho}(p)\pol^\sigma(p-k_1)\nn
\eea
without committing to a specific choice of polarization for the gluons with $p$-dependent external momenta. These degrees of freedom will eventually be summed over. Actually, the correct summation procedure is fairly obvious~\cite{PeskinSchroeder}. We can use the na\"{i}ve replacement
\be
\sum_\lambda \pol^\lambda_\rho(k_1)\pol^{*\,\lambda}_\sigma(k_1) \rightarrow -g_{\rho \sigma}
\ee
valid in Abelian gauge theory, provided that we correct for the fact that we are overcounting states by including the quintuple cut of a ghost loop. This is simple since the contribution from a ghost loop is nothing but the contribution from a complex scalar loop with an extra overall minus sign coming the fact that the ghost field obeys Fermi-Dirac statistics:
\be
A_{1;\,{\rm ghost}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5} = -A_{1;\,{\rm scalar}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5}\,{\rm .}
\ee
Returning to the quintuple cut of the gluon loop, we have
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1;\,{\rm gluon}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5} = \left(i \sqrt{2}\right)^5 \pol^{*\,\rho_1}(p_*) \Big(\pol^-(k_1)\cdot p_*~ g_{\rho_1 \sigma_1}+k_{1\,\rho_1}~\pol^-_{\sigma_1}(k_1)
\el-k_{1\,\sigma_1}~ \pol^-_{\rho_1}(k_1)\Big)\pol^{\sigma_1}(p_*-k_1) \pol^{*\,\rho_2}(p_*-k_1) \Big(\pol^-(k_2)\cdot (p_*-k_1)~ g_{\rho_2 \sigma_2}+k_{2\,\rho_2}~\pol^-_{\sigma_2}(k_2)
\el-k_{2\,\sigma_2}~ \pol^-_{\rho_2}(k_2)\Big)\pol^{\sigma_2}(p_*-k_1-k_2)
\pol^{*\,\rho_3}(p_*-k_1-k_2) \Big(\pol^+(k_3)\cdot (p_*-k_1-k_2)~ g_{\rho_3 \sigma_3}+k_{3\,\rho_3}~\pol^+_{\sigma_3}(k_3)
\el-k_{3\,\sigma_3}~ \pol^+_{\rho_3}(k_3)\Big)\pol^{\sigma_3}(p_*+k_4+k_5)
\pol^{*\,\rho_4}(p_*+k_4+k_5) \Big(\pol^+(k_4)\cdot (p_*+k_4+k_5)~ g_{\rho_4 \sigma_4}+k_{4\,\rho_4}~\pol^+_{\sigma_4}(k_4)
\el-k_{4\,\sigma_4}~ \pol^+_{\rho_4}(k_4)\Big)\pol^{\sigma_4}(p_*+k_5)\pol^{*\,\rho_5}(p_*+k_5) \left(\pol^+(k_5)\cdot (p_*+k_5)~ g_{\rho_5 \sigma_5}+k_{5\,\rho_5}~\pol^+_{\sigma_5}(k_5)-k_{5\,\sigma_5}~ \pol^+_{\rho_5}(k_5)\right)\pol^{\sigma_5}(p_*)
\elale \left(i \sqrt{2}\right)^5 \left(\pol^-(k_1)\cdot p_*~ g_{\rho_1 \sigma_1}+k_{1\,\rho_1}~\pol^-_{\sigma_1}(k_1)-k_{1\,\sigma_1}~ \pol^-_{\rho_1}(k_1)\right)\left(-g^{\sigma_1 \rho_2}\right)
\el\left(\pol^-(k_2)\cdot (p_*-k_1)~ g_{\rho_2 \sigma_2}+k_{2\,\rho_2}~\pol^-_{\sigma_2}(k_2)-k_{2\,\sigma_2}~ \pol^-_{\rho_2}(k_2)\right)\left(-g^{\sigma_2 \rho_3}\right)
\el\left(\pol^+(k_3)\cdot (p_*-k_1-k_2)~ g_{\rho_3 \sigma_3}+k_{3\,\rho_3}~\pol^+_{\sigma_3}(k_3)-k_{3\,\sigma_3}~ \pol^+_{\rho_3}(k_3)\right)\left(-g^{\sigma_3 \rho_4}\right)
\el\left(\pol^+(k_4)\cdot (p_*+k_4+k_5)~ g_{\rho_4 \sigma_4}+k_{4\,\rho_4}~\pol^+_{\sigma_4}(k_4)-k_{4\,\sigma_4}~ \pol^+_{\rho_4}(k_4)\right)\left(-g^{\sigma_4 \rho_5}\right)
\el\left(\pol^+(k_5)\cdot (p_*+k_5)~ g_{\rho_5 \sigma_5}+k_{5\,\rho_5}~\pol^+_{\sigma_5}(k_5)-k_{5\,\sigma_5}~ \pol^+_{\rho_5}(k_5)\right)\left(-g^{\sigma_5 \rho_1}\right)\,{\rm .}
\label{g5ptpent}
\eea
\end{changemargin}
Finally, we combine together all of the above results with the appropriate multiplicities:
\cmb{-1.0 in}{0 in}
\bea
A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5} &=& 3 A_{1;\,{\rm scalar}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5}+4 A_{1;\,{\rm fermion}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5}
\el
+\left(A_{1;\,{\rm gluon}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5}-A_{1;\,{\rm scalar}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Big|_{I_5}\right) \,{\rm ,}\nn
\label{5ptpent}
\eea
\cme
where we have dealt with the ghost loop contribution as discussed above. One can straightforwardly check (numerically using {\it e.g.} S@M~\cite{S@M}) that, after projecting (\ref{5ptpent}) onto the dual conformal basis using eq. (\ref{PtoDCB}), the result agrees with that obtained earlier using the leading singularity method (eq. (\ref{B5ptDCB})). Evaluating the numerator algebra becomes slightly more involved for quintuple cuts of one-loop six-gluon amplitudes, but we will still be able to use the above procedure to great effect.
We are finally in a position to outline the strategy that we will use to solve, say, $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6)$ to all orders in $\e$. This amplitude works well as an example because its full analytical form is known~\cite{oneloopselfdual}:
\cmb{-.6 in}{0 in}
\bea
&&A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6) ={A_{6;~\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} \bigg(- s_3 s_4 I_{4}^{(1,2),\,D=4-2\e}- {s_4 s_5} I_{4}^{(2,3),\,D=4-2\e}
\el- {s_5 s_6} I_{4}^{(3,4),\,D=4-2\e} -{s_1 s_6 }I_{4}^{(4,5),\,D=4-2\e} - {s_1 s_2 }I_{4}^{(5,6),\,D=4-2\e}- {s_2 s_3 } I_{4}^{(1,6),\,D=4-2\e}+ (s_3 s_6 - t_2 t_3) I_{4}^{(1,4),\,D=4-2\e}
\el+ (s_1 s_4 - t_1 t_3) I_{4}^{(2,5),\,D=4-2\e} + (s_2 s_5 - t_1 t_2) I_{4}^{(3,6),\,D=4-2\e}
+ \e \sum_{i=1}^6 \pol(i+1, i+2, i+3, i+4) I_5^{(i),\,D=6-2\e}
\el+\e~ \textrm{tr}[\slashed{k_1}\slashed{k_2}\slashed{k_3}\slashed{k_4}\slashed{k_5}\slashed{k_6}] I_6^{D=6-2\e}\bigg) \,{\rm .}
\label{6ptMHV}
\eea
\cme
We will, of course, mostly be interested in evaluating all\footnote{In the next subsection we will go through the exercise of determining how many independent NMHV gluon amplitudes there are (ignoring $\Nsym$ supersymmetry for now).} six-gluon NMHV amplitudes, but the strategy utilized for $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6)$ carries over in a completely straightforward fashion to the other six-gluon amplitudes.
The general idea is that, while the leading singularity method does not fix everything to all orders in $\e$ starting at six points, the method is very powerful and {\it does} fix everything up to terms with trivial soft and collinear limits. To illustrate this point let us discuss to what extent the universal factorization properties of $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6)$ under soft and collinear limits determine the analytic form of the amplitude, given that we already know $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)$ to all orders in $\e$. It turns out that there is only one function in $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6)$ that is not constrained in this approach: One can check that
\be
\textrm{tr}[\slashed{k_1}\slashed{k_2}\slashed{k_3}\slashed{k_4}\slashed{k_5}\slashed{k_6}]
\ee
has no soft or collinear limits in any channel.\footnote{A soft or collinear limit for planar amplitudes is particularly simple because one only has to consider nearest-neighbor pairs of momenta. If unfamiliar, see~\cite{Dixon96rev} for an elementary discussion of planar soft and collinear limits.} Therefore any attempt to deduce the form of the one-loop six-gluon MHV amplitude from that of the one-loop five-gluon amplitude by demanding consistency of the soft and collinear limits will miss terms like that above.
This ambiguity is reflected in the solution of the leading singularity equations for $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6)$. Solving the system of $15\times 2 = 30$ equations in $15 + 6 = 21$ unknowns\footnote{Here we would like to remind the reader that, if desired, the one-loop scalar hexagon integral may be expressed as a linear combination of six one-loop scalar pentagon integrals (see eq. (\ref{hexred})).} determines 20 of the unknown integral coefficients in terms of one of the pentagon coefficients, say that associated to $I_5^{(1),\,4-2\e}$. The point is that if we can evaluate one pentagon coefficient using $D$-dimensional unitarity, then the leading singularity equations, which require only four dimensional inputs, give us everything else. This is a much better strategy than trying to evaluate the quintuple cut of each pentagon independently because it allows one to solve for all the pentagon coefficients with a minimum of effort beyond that required to determine the coefficients of the boxes.
Before going any further, we should clarify a potentially confusing point about the solution to the leading singularity equations for $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6)$. Suppose we let $B_1$ be the coefficient of $I_5^{(1),\,4-2\e}$. Then a generic box coefficient, say that of $I_{4}^{(1,6),\,D=4-2\e}$, will have the form $\alpha_{16}+\beta_{16} B_1$. It may seem strange that the box coefficient associated to $I_{4}^{(1,6),\,D=4-2\e}$ depends on the pentagon coefficient $B_1$. This apparent paradox is resolved by projecting the geometric basis onto the dual conformal basis: the pentagons $I_{5}^{(1),\,D=4-2\e}$ and $I_{5}^{(6),\,D=4-2\e}$ each contribute to the coefficient of $I_{4}^{(1,6),\,D=4-2\e}$ in the dual conformal basis after the formula (\ref{PtoDCB}) is applied to them. Remarkably, these extra contributions conspire to cancel all of the $B_1$ dependence that was present in the coefficient of $I_{4}^{(1,6),\,D=4-2\e}$, considered as an element of the geometric basis.
In solving the leading singularity equations, we were free to choose any pentagon coefficient we wanted as the parameter undetermined by the system. The reason that we chose the coefficient of $I_5^{(1),\,4-2\e}$ is that it is particularly simple to determine this integral coefficient using quintuple cuts. This follows from the fact that $A^{tree}_{\mu^2}\left((p-k_1)_s,k_1^{1234},k_6,(-p-k_6)_{\bar{s}}\right)$, $A^{tree}_{\mu^2}\left((p-k_1)_{\bar{f}},k_1^{1234},k_6,(-p-k_6)_f\right)$, and $A^{tree}_{\mu^2}\left((p-k_1)_{g},k_1^{1234},k_6,(-p-k_6)_g\right)$ can each be represented by a single Feynman diagram:
\cmb{-.6 in}{0 in}
\bea
A^{tree}_{\mu^2}\left((p-k_1)_s,k_1^{1234},k_6,(-p-k_6)_{\bar{s}}\right) &=& -{i \spab1.{p}.6^2 \over s_6 \spab1.p.1}
\label{mu4tree1}\\ A^{tree}_{\mu^2}\left((p-k_1)_{\bar{f}},k_1^{1234},k_6,(-p-k_6)_f\right) &=& {i (p+k_6)\cdot \pol^+(k_6)\over \spab1.p.1}\bar{u}(p+k_6)\slashed{\pol}^-(k_1)u(p-k_1)
\label{mu4tree2}\\ A^{tree}_{\mu^2}\left((p-k_1)_{g},k_1^{1234},k_6,(-p-k_6)_g\right) &=& -{2 i\pol^\rho(p-k_1) \pol^{*\,\sigma}(p+k_6) \over \spab1.p.1}\Big(\pol^-(k_1)\cdot p ~\pol^+(k_6)\cdot p ~g_{\rho \sigma}
\nonumber\\&&+ \pol^+(k_6)\cdot p ~k_{1\,\sigma} \pol^-_\rho(k_1)-\pol^+(k_6)\cdot p ~k_{1\,\rho}~ \pol^-_\sigma(k_1) + \pol^-(k_1)\cdot p ~k_{6\,\sigma}~ \pol^+_{\rho}(k_6)
\nonumber\\&&- \pol^-(k_1)\cdot p ~k_{6\,\rho}~ \pol^+_{\sigma}(k_6) - k_1\cdot k_6 \pol^-_\rho(k_1)\pol^+_\sigma(k_6)\Big)
\label{mu4tree3}\eea
\cme
Using the same logic that was employed for the five-point pentagon coefficient calculated above and the results of eqs. (\ref{mu4tree1})-(\ref{mu4tree3}), it is straightforward to compute $B_1$. One subtlety is that the line $p^2-\mu^2$ is left uncut in the evaluation of this integral coefficient. As a result, the expression for $\mu^2$ is not given simply by $p_*^2$, the way that it was in the five-point example worked out in detail above. Instead, one has the relation $p_*^2-2 p_*\cdot k_1 = \mu^2$. Also, to use the framework of eqs. (\ref{5ptos1}) and (\ref{5ptos2}), we have to make the following adjustments: (\ref{5ptos1}) becomes
\be
p^\nu = L_1 K_1^\nu + L_2 K_2^\nu + L_3 K_3^\nu + L_4 K_4^\nu+k_1
\ee
and the $K_i$ four-vectors all need to be shifted by $-k_1$ ({\it i.e.} instead of $K_1 = k_1 + k_2 + k_3$, we have $K_1 = k_2 + k_3$).
Before presenting the results of our all-orders-in-$\e$ six-gluon NMHV calculations, we make some remarks about how we expect the strategy outlined for six-gluon MHV amplitudes to generalize to higher-multiplicity amplitudes. First of all, we conjecture that the number of unconstrained by the one-loop soft/collinear bootstrap is controlled by kinematics as opposed to dynamics ({\it i.e.} independent of the $k$ in $\rm{N}^k$MHV). We interpret the fact that this is true at the six-point level (in the sense that the leading singularities for both MHV and NMHV amplitudes fix everything up to a single pentagon coefficient) as evidence for this proposal. If this conjecture is indeed correct, it follows that the number of terms in arbitrary $n$-point amplitudes left unconstrained by the soft/collinear bootstrap is equal to the number of unconstrained terms in the $n$-gluon MHV amplitudes. The number of such terms is 6 in the seven-gluon and 21 in the eight-gluon MHV amplitudes and we expect the answer to be the same for their non-MHV counterparts.
Thus, we conclude this subsection by conjecturing that the number of terms left unconstrained by the soft/collinear bootstrap at the $n$-point level (pentagon coefficients undetermined by the leading singularity method) is
\be
\left(\begin{array}{c}n - 1 \\ 5\end{array}\right) = {(n-5)(n-4)(n-3)(n-2)(n-1)\over 120} \,{\rm ,}
\ee
equal to the number of pentagons at the $(n-1)$-point level. Loosely speaking, we can think of this result as the statement that, at the $n$-point level an independent object analogous to $\textrm{tr}[\slashed{k_1}\slashed{k_2}\slashed{k_3}\slashed{k_4}\slashed{k_5}\slashed{k_6}]$ can be constructed for each pentagon integral at the $(n-1)$-point level without spoiling any of the soft/collinear constraints relating $n$-point one-loop planar amplitudes to $(n-1)$-point one-loop planar amplitudes.
\subsection{The All-Orders in $\e$ Planar One-Loop $\Nsym$ NMHV Six-Gluon Amplitudes}
\label{gresults}
In this subsection, we give our formulae for the one-loop planar six-gluon NMHV pentagon coefficients in $\Nsym$ and discuss the structural similarities between our results and certain two-loop planar six-gluon integral coefficients entering into the NMHV amplitudes calculated in~\cite{KRV}. Our first task, of course, is to understand how many independent NMHV gluon amplitudes there are (delaying a discussion of the constraints coming from $\Nsym$ supersymmetry until Subsection \ref{gendisonshell}). Na\"{i}vely, there are a large number of possibilities:
\begin{tabular}{p{0.47\linewidth} p{0.4\linewidth}}
\begin{enumerate}
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5^{1234},k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4^{1234},k_5,k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4,k_5^{1234},k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4,k_5,k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3,k_4^{1234},k_5^{1234},k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3,k_4^{1234},k_5,k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3,k_4,k_5^{1234},k_6^{1234}\right)$
\end{enumerate} &
\begin{enumerate}
\setcounter{enumi}{10}
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2^{1234},k_3^{1234},k_4^{1234},k_5,k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2^{1234},k_3^{1234},k_4,k_5^{1234},k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2^{1234},k_3^{1234},k_4,k_5,k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2^{1234},k_3,k_4^{1234},k_5^{1234},k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2^{1234},k_3,k_4^{1234},k_5,k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2^{1234},k_3,k_4,k_5^{1234},k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2,k_3^{1234},k_4^{1234},k_5^{1234},k_6\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2,k_3^{1234},k_4^{1234},k_5,k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2,k_3^{1234},k_4,k_5^{1234},k_6^{1234}\right)$
\item $A_{1}^{1-{\rm loop}}\left(k_1,k_2,k_3,k_4^{1234},k_5^{1234},k_6^{1234}\right)$
\end{enumerate}\\
\end{tabular}
many of which are obviously related by parity\footnote{Recall that CP is a good symmetry of perturbative scattering amplitudes even in pure $\mathcal{N} = 0$ Yang-Mills.} or cyclic symmetry\footnote{Recall from \ref{planar} that, for example, the amplitudes $A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6)$ and $A_{1}^{1-{\rm loop}}(k_2^{1234},k_3^{1234},k_4,k_5,k_6,k_1^{1234})$ are equal by virtue of the color structure in the planar limit.}. In particular, we will now show that all of the above can be related to $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$, $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)$, and $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4,k_5^{1234},k_6\right)$ (amplitudes 1., 2., and 6. in the above). In fact, we will see in Section \ref{supercomp} that it is possible to derive a beautiful all-orders-in-$\e$ $\Nsym$ supersymmetrization of the six-point NMHV amplitudes using only these three all-orders-in-$\e$ component amplitudes as inputs. To start, we see immediately that amplitudes 11. - 20. are related to 1. - 10. by parity which acts on the amplitudes by reversing the helicities of all gluons. Next, we see that 4. and 10. are related to 1. through a series of cyclic shifts followed by a relabeling of the momenta of the external gluons. Similarly, amplitudes 7. and 8. are related to 2. through cyclic shifts followed by a relabeling. Finally, 3., 5., and 9. are related to 2. through cyclic shifts, a relabeling, {\it and} a parity transformation. Amplitude 6. can't be related to 1. or 2. through some combination of parity and cyclicity, so we need to include it in our basis as well.
Now that we understand why it makes sense to focus on $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$, $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)$, and $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4,k_5^{1234},k_6\right)$, we present our results for these amplitudes. To begin, let us recall the results of the calculations performed in~\cite{BDDKNMHV}. The authors of that work determined the box coefficients for all NMHV gluon amplitudes in the dual conformal basis. The $6 - 2\e$ dimensional pentagon coefficients, however, were undetermined. It was found that
\bea
\nonumber A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)&&=-{1\over2}B_1 \Bigg( s_4 s_5 I_4^{(2,3)}+s_1 s_2 I_4^{(5,6)}+s_6 t_1 I_4^{(3,5)}+s_3 t_1 I_4^{(2,6)}\Bigg) \\ \nonumber
&&-{1\over2}B_2 \Bigg( s_5 s_6 I_4^{(3,4)}+s_2 s_3 I_4^{(6,1)}+s_1 t_2 I_4^{(4,6)}+s_4 t_2 I_4^{(1,3)}\Bigg) \\ \nonumber
&&-{1\over2}B_3 \Bigg( s_6 s_1 I_4^{(4,5)}+s_3 s_4 I_4^{(1,2)}+s_2 t_3 I_4^{(1,5)}+s_5 t_3 I_4^{(2,4)}\Bigg) \\ \nonumber
&& +K_1 \e I_5^{(1),6-2 \e}+K_2 \e I_5^{(2),6-2 \e}+K_3 \e I_5^{(3),6-2 \e}
\\
&& +K_4 \e I_5^{(4),6-2 \e}+K_5 \e I_5^{(5),6-2 \e}+K_6 \e I_5^{(6),6-2 \e} \, ,
\eea
\bea
\nonumber A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4,k_5^{1234},k_6\right)&&=-{1\over2}G_1 \Bigg( s_4 s_5 I_4^{(2,3)}+s_1 s_2 I_4^{(5,6)}+s_6 t_1 I_4^{(3,5)}+s_3 t_1 I_4^{(2,6)}\Bigg) \\ \nonumber
&&-{1\over2}G_2 \Bigg( s_5 s_6 I_4^{(3,4)}+s_2 s_3 I_4^{(6,1)}+s_1 t_2 I_4^{(4,6)}+s_4 t_2 I_4^{(1,3)}\Bigg) \\ \nonumber
&&-{1\over2}G_3 \Bigg( s_6 s_1 I_4^{(4,5)}+s_3 s_4 I_4^{(1,2)}+s_2 t_3 I_4^{(1,5)}+s_5 t_3 I_4^{(2,4)}\Bigg) \\ \nonumber
&& +F_1 \e I_5^{(1),6-2 \e}+F_2 \e I_5^{(2),6-2 \e}+F_3 \e I_5^{(3),6-2 \e}
\\
&& +F_4 \e I_5^{(4),6-2 \e}+F_5 \e I_5^{(5),6-2 \e}+F_6 \e I_5^{(6),6-2 \e} \, ,
\eea
and
\bea
\nonumber A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)&&=-{1\over2}D_1 \Bigg( s_4 s_5 I_4^{(2,3)}+s_1 s_2 I_4^{(5,6)}+s_6 t_1 I_4^{(3,5)}+s_3 t_1 I_4^{(2,6)}\Bigg) \\ \nonumber
&&-{1\over2}D_2 \Bigg( s_5 s_6 I_4^{(3,4)}+s_2 s_3 I_4^{(6,1)}+s_1 t_2 I_4^{(4,6)}+s_4 t_2 I_4^{(1,3)}\Bigg) \\ \nonumber
&&-{1\over2}D_3 \Bigg( s_6 s_1 I_4^{(4,5)}+s_3 s_4 I_4^{(1,2)}+s_2 t_3 I_4^{(1,5)}+s_5 t_3 I_4^{(2,4)}\Bigg) \\ \nonumber
&& +H_1 \e I_5^{(1),6-2 \e}+H_2 \e I_5^{(2),6-2 \e}+H_3 \e I_5^{(3),6-2 \e}
\\
&& +H_4 \e I_5^{(4),6-2 \e}+H_5 \e I_5^{(5),6-2 \e}+H_6 \e I_5^{(6),6-2 \e} \,.
\eea
All of the spin factors which entered into the box coefficients ($B_i$, $G_i$, and $D_i$) were determined. They are given by
\bea
B_1 &=& B_0 \label{B1}\\
B_2 &=& \bigg({\spab1.{2+3}.4\over t_2}\bigg)^4 B_0\Big|_{j\rightarrow j+1}+\bigg({\spa2.3 \spb5.6 \over t_2}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]}\Big|_{j\rightarrow j+1} \\
B_3 &=& \bigg({\spab3.{1+2}.6\over t_3}\bigg)^4 B_0\Big|_{j\rightarrow j-1}+\bigg({\spa1.2 \spb4.5 \over t_3}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]}\Big|_{j\rightarrow j-1} \, ,
\eea
\bea
G_1 &=& \bigg({\spab5.{4+6}.2\over t_1}\bigg)^4 B_0+\bigg({\spa1.3 \spb4.6 \over t_1}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]} \\
G_2 &=& \bigg({\spab3.{2+4}.6\over t_2}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]}\Big|_{j\rightarrow j+1}+\bigg({\spa5.1 \spb2.4 \over t_2}\bigg)^4 B_0\Big|_{j\rightarrow j+1} \\
G_3 &=& \bigg({\spab1.{2+6}.4\over t_3}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]}\Big|_{j\rightarrow j-1}+\bigg({\spa3.5 \spb6.2 \over t_3}\bigg)^4 B_0\Big|_{j\rightarrow j-1} \, ,
\eea
and
\bea
D_1 &=& \bigg({\spab4.{1+2}.3\over t_1}\bigg)^4 B_0+\bigg({\spa1.2 \spb5.6 \over t_1}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]} \\
D_2 &=& \bigg({\spab1.{2+4}.3\over t_2}\bigg)^4 B_0\Big|_{j\rightarrow j+1}+\bigg({\spa2.4 \spb5.6 \over t_2}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]}\Big|_{j\rightarrow j+1} \\
D_3 &=& \bigg({\spab4.{1+2}.6\over t_3}\bigg)^4 B_0\Big|_{j\rightarrow j-1}+\bigg({\spa1.2 \spb3.5 \over t_3}\bigg)^4 B_0^{\langle~\rangle \leftrightarrow [~]}\Big|_{j\rightarrow j-1} \, ,
\eea
where
\be
B_0 = i{\spa1.2 \spa2.3 \spb4.5 \spb5.6 \spab3.{1+2}.6 \spab1.{2+3}.4 t_1^3 \over s_1 s_2 s_4 s_5 (t_1 t_2-s_2 s_5)(t_1 t_3-s_1 s_4)} \, . \label{B0}
\ee
Using the strategy outlined it \ref{effgcomp}, we reproduce the above and, furthermore, find explicit expressions for the $K_i$, $G_i$, and $H_i$.
Although the raw answers obtained using the method described in the last subsection are already compact enough to fit on a single page, it is clearly desirable to find more compact formulae. In their work on the two-loop planar NMHV gluon amplitudes~\cite{KRV}, Kosower, Roiban, and Vergu derived explicit expressions for all possible $\mu$-integral hexabox coefficients (see Figure \ref{hexabox} for an illustration). Motivated by issues of IR consistency that we will elaborate on in Section \ref{WL/MHV}, we evaluated the answers they obtained numerically and were able to find a straightforward mapping between their results and ours. To explain this relationship, it is useful to consider a concrete example.
We consider the coefficient of the $s_1$-channel hexabox integral (see Figure \ref{hexabox}) that appears in the amplitude $A_{1}^{2-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$ calculated to all orders in $\e$.
\FIGURE{
\resizebox{0.6\textwidth}{!}{\includegraphics{hexaboxdiag.eps}}
\vbox{\vskip 1 in}
\caption{The $s_1$-channel hexabox $\mu$-integral topology of $A_{1}^{2-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$. The factor of $\mu^2$ should be thought of as belonging to the hexagon subdiagram.}
\label{hexabox}}
It turns out that this $\mu$-integral hexabox coefficient and $K_2$ (coefficient of the $s_1$-channel $6 - 2\e$ dimensional pentagon coefficient that appears in $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$ are simply related:
\be
K_2 = {C_2 \over 2 s_1} \mathcal{K}_2 \, {\rm ,}
\label{hexpentrel}
\ee
where we have given the hexabox coefficient the convenient label $\mathcal{K}_2$ and $C_2$ is one of the variables that we used to define the reduction of the one-loop scalar hexagon integral to a sum of six scalar pentagons in \ref{muint}:
\be
I_6 = {1\over 2}\sum_{i = 1}^6 C_i I_5^{(i)}\,{\rm .}
\label{hexred}
\ee
This relation makes a certain amount of sense if we think about collapsing the box in the $\mu$-integral hexabox in Figure \ref{hexabox} to a point. This turns the hexabox $\mu$-integral into a pentagon $\mu$-integral. Evidently, $s_1$ appears because we are working with the $s_1$ channel hexabox and, perhaps, $C_2$ appears because we are relating an object with six external legs to one with five. In any case, the above relation will allow us to exploit extremely simple results found for the NMHV hexabox coefficients~\cite{KRV} to write beautiful formulas for the $K_i$, $F_i$, and $H_i$. We find
\bea
K_1 &=& {i \over 2} C_1 {\Big(\spab2.{1+6}.5 \spab1.{2+3}.4 \spab3.{1+2}.6 +\spab5.{1+6}.2 \spa1.2 \spa2.3 \spb4.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
K_2 &=& {i \over 2} C_2 {\spab3.{1+2}.6^2 \spa1.2^2 \spb4.5^2 t_1^2 \over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \\
K_3 &=& {i \over 2} C_3 {\spab1.{2+3}.4^2 \spa2.3^2 \spb5.6^2 t_1^2 \over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \\
K_4 &=& {i \over 2} C_4 {\Big(\spab2.{1+6}.5 \spab1.{2+3}.4 \spab3.{1+2}.6 +\spab5.{1+6}.2 \spa1.2 \spa2.3 \spb4.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
K_5 &=& {i \over 2} C_5 {\spab3.{1+2}.6^2 \spa1.2^2 \spb4.5^2 t_1^2 \over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \\
K_6 &=& {i \over 2} C_6 {\spab1.{2+3}.4^2 \spa2.3^2 \spb5.6^2 t_1^2 \over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \, ,
\eea
\bea
F_1 &=& {i \over 2} C_1 {\Big(\spab5.{6+1}.2 \spab3.{2+4}.6 \spab1.{3+5}.4 +\spab2.{6+1}.5 \spb6.2 \spb2.4 \spa1.5 \spa3.5 \Big)^2\over s_6 s_3 \spab5.{6+1}.2 \spab2.{6+1}.5} \nonumber \\ \\
F_2 &=& {i \over 2} C_2 {\Big(\spab3.{1+2}.6 \spab1.{3+5}.4 \spab5.{4+6}.2 +\spab6.{1+2}.3 \spa1.3 \spa3.5 \spb2.6 \spb4.6 \Big)^2\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \nonumber \\ \\
F_3 &=& {i \over 2} C_3 {\Big(\spab1.{2+3}.4 \spab5.{4+6}.2 \spab3.{5+1}.6 +\spab4.{2+3}.1 \spb2.4 \spb4.6 \spa3.1 \spa5.1 \Big)^2\over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \nonumber \\ \\
F_4 &=& {i \over 2} C_4 {\Big(\spab5.{6+1}.2 \spab3.{2+4}.5 \spab1.{3+5}.4 +\spab2.{6+1}.5 \spb2.6 \spb4.2 \spa5.1 \spa5.3 \Big)^2\over s_6 s_3 \spab5.{6+1}.2 \spab2.{6+1}.5} \nonumber \\ \\
F_5 &=& {i \over 2} C_5 {\Big(\spab3.{1+2}.6 \spab1.{3+5}.4 \spab5.{4+6}.2 +\spab6.{1+2}.3 \spa1.3 \spa3.5 \spb2.6 \spb4.6 \Big)^2\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \nonumber \\ \\
F_6 &=& {i \over 2} C_6 {\Big(\spab1.{2+3}.4 \spab5.{4+6}.2 \spab3.{5+1}.6 +\spab4.{2+3}.1 \spb2.4 \spb4.6 \spa3.1 \spa5.1 \Big)^2\over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \, ,\nonumber \\
\eea
and
\bea
H_1 &=& {i \over 2} C_1 {\Big(\spab2.{6+1}.5 \spab1.{2+4}.3 \spab4.{1+2}.6 +\spab5.{6+1}.2 \spa1.2 \spa2.4 \spb3.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
H_2 &=& {i \over 2} C_2 {\spa1.2^2 \Big(\spab3.{1+2}.6 \spab4.{1+2}.3 \spb5.3 +\spab6.{1+2}.3 \spab4.{1+2}.6 \spb5.6\Big)^2\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \nonumber \\ \\
H_3 &=& {i \over 2} C_3 {\spb5.6^2 \Big(\spab1.{2+3}.4 \spab4.{1+2}.3 \spa2.4 +\spab4.{2+3}.1 \spab1.{2+4}.3 \spa2.1\Big)^2\over s_2 s_5 \spab4.{2+3}.1 \spab1.{2+3}.4} \nonumber \\ \\
H_4 &=& {i \over 2} C_4 {\Big(\spab2.{6+1}.5 \spab1.{2+4}.3 \spab4.{1+2}.6 +\spab5.{6+1}.2 \spa1.2 \spa2.4 \spb3.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
H_5 &=& {i \over 2} C_5 {\spa1.2^2 \Big(\spab3.{(1+2)}.6 \spab4.{1+2}.3 \spb5.3 +\spab6.{1+2}.3 \spab4.{1+2}.6 \spb5.6\Big)^2\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \nonumber \\ \\
H_6 &=& {i \over 2} C_6 {\spb5.6^2 \Big(\spab1.{2+3}.4 \spab4.{1+2}.3 \spa2.4 +\spab4.{2+3}.1 \spab1.{2+4}.3 \spa2.1\Big)^2\over s_2 s_5 \spab4.{2+3}.1 \spab1.{2+3}.4} \, {\rm .}\nonumber \\
\eea
We also checked these results against a Feynman diagram calculation.
These results have a couple of striking features of which we have only a partial understanding. The numerators of all the spin factors (divided by the appropriate $C_i$) are perfect squares. Furthermore, the pentagon coefficients possess a certain $i \rightarrow i+3$ symmetry:
\be
{K_1 \over C_1} = {K_4 \over C_4} \qquad
{K_2 \over C_2} = {K_5 \over C_5} \qquad
{K_3 \over C_3} = {K_6 \over C_6} \, {\rm .}
\label{mystrel}
\ee
with analogous formulas for the $F_i$ and $H_i$. As we will see in Section \ref{WL/MHV}, this $i \rightarrow i+3$ symmetry is related to the action of parity when the amplitude is written in a way that makes a hidden symmetry\footnote{This hidden symmetry is called dual superconformal invariance and some background and motivation for it is provided in the second part of Appendix \ref{ADS/CFT}.} of the planar S-matrix as manifest as possible. In the next section we explore an interesting connection between all-orders-in-$\e$ one-loop $\Nsym$ amplitudes and the first two non-trivial orders in the $\alpha'$ expansion of tree-level superstring amplitudes. The explicit one-loop results presented so far in this paper will provide us with useful explicit cross-checks on the relations we propose.
\section{New Relations Between One-Loop Amplitudes in $\Nsym$ Gauge Theory and Tree-Level Amplitudes in Open Superstring Theory}
\label{gsrel}
Before reviewing the scattering of massless modes in open superstring theory, we motivate what follows. Stieberger and Taylor~\cite{ST1} calculated the lowest-order, $\Ord(\alpha'^2)$, stringy corrections to $\Nsym$ tree-level gluon MHV amplitudes.\footnote{As mentioned in the introduction, tree-level amplitudes of massless particles in open superstring constructions compactified to four dimensions have a universal form~\cite{ST1}.} They found that their result\footnote{In Stieberger and Taylor's notation, say at the six-point level, $[[1]]_1 = s_1$, $\{\,[[1]]_1\,\} = s_1 + s_2 + s_3 + s_4 + s_5 + s_6$, $[[1]]_2 = t_1$, and $\{\,[[1]]_2\,\} = t_1 + t_2 + t_3$.},
\bea
A^{tree}_{str}\left(k_1^{1234},k_2^{1234},k_3,\cdots,k_n\right)\Bigl|_{\Ord(\alpha'^2)} &=& -{\pi^2\over 12} A_{n;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}\Bigg(
\sum_{k=1}^{\left[\frac{n}{2}-1\right]}\{\,[[1]]_k [[2]]_k\,\}-
\sum_{k=3}^{\left[\frac{n}{2}-1\right]}\{\,[[1]]_k [[2]]_{k-2}\,\}\nonumber\\
&& +~C^{(n)}+\sum_{j<k<\ell<m<n} \pol(j,k,\ell,m)\Bigg)\, , \label{mhvn}\nonumber\\
C^{(n)} &=& \Bigg\{ \begin{array}{ll}
-\{\,[[1]]_{\frac{n}{2}-2} [[{n \over 2}+1]]_{\frac{n}{2}-2}\}
& \mbox{$n>4$, even,}\nonumber \\
-\big\{\,[[1]]_{\frac{n-5}{2}} [[\frac{n+1}{2}]]_{\frac{n-3}{2}}\big\} & \mbox{$n>5$, odd},\end{array}
\label{stringMHVoal2}
\eea
was precisely equal to $-6 \zeta(2)$ times the analogous one-loop $\Nsym$ amplitude with a factor of $\mu^4$ inserted into the numerator of each basis integral, $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,\cdots,k_n\right)[\mu^4]\Bigl|_{\e\to0}$. This non-obvious connection was actually made by showing that both
$${A^{tree}_{str}\left(k_1^{1234},k_2^{1234},k_3,\cdots,k_n\right)\Bigl|_{\Ord(\alpha'^2)}\over \spa1.2^4} ~~~~{\rm and}~~~~ {A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,\cdots,k_n\right)[\mu^4]\Bigl|_{\e\to0} \over \spa1.2^4}$$
are, apart from trivial constants, equal to the all-plus one-loop amplitude in pure Yang-Mills theory~\cite{oneloopselfdual}, $A_{1;\,\mathcal{N}=0}^{1-{\rm loop}}\left(k_1,k_2,\cdots,k_n\right)$. The only reason an equivalence between
$${A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,\cdots,k_n\right)[\mu^4]\Bigl|_{\e\to0}\over\spa1.2^4} ~~~~{\rm and}~~~~ {A_{1;\,\mathcal{N}=0}^{1-{\rm loop}}\left(k_1,k_2,k_3,\cdots,k_n\right)}$$ is possible is that both have the same manifest invariance under cyclic shifts $i \rightarrow i+1$. It is hard to imagine that additional relationships between $\mathcal{N}=0$ and $\Nsym$ amplitudes could exist because, in general, there is no reason to expect $\mathcal{N}=0$ and $\Nsym$ amplitudes to have similar symmetry properties (for more general amplitudes there is no trick analogous to dividing the one-loop MHV amplitude by $\spa1.2^4$). Indeed, it is incredibly likely that this relation between pure Yang-Mills and $\Nsym$ is purely accidental. However, additional relations between superstring tree amplitudes and $\Nsym$ one-loop amplitudes are a more realistic possibility. It is this possibility that we discuss in this section. The new results presented are based on unpublished work done in collaboration with Lance J. Dixon~\cite{myfourth}.
\subsection{Organization of the Tree-Level Open Superstring S-matrix}
\label{BornInfeld}
For the simple case of a $U(1)$ gauge group, it has been known since the work of Fradkin and Tseytlin~\cite{FT} that the effective action governing the low-energy dynamics of open superstrings ending on a single Dirchlet 3-brane (though the connection between gauge symmetry and D-branes remained hidden until the work of Dai, Leigh, and Polchinski in~\cite{DLP} and Leigh in~\cite{L}) is nothing but a supersymmetrization of the Born-Infeld action. This action, expressed in terms of the Maxwell field strength tensor,
\bea
\mathcal{L}_{BI} &=& {1 \over (2 \pi g \alpha')^2} \left(1 - \sqrt{{\rm Det}\big(~g_{\mu \nu} + (2 \pi g \alpha') F_{\mu \nu} ~\big)} \right) \nonumber \\
&=& {1 \over (2 \pi g \alpha')^2} \left(1 - \sqrt{1 + {(2 \pi g \alpha')^2\over2}F_{\mu \nu}F_{\mu \nu} - {(2 \pi g \alpha')^4\over16}~ (F_{\mu \nu} \tilde{F}_{\mu \nu})^2} \right) \nonumber \\
&=& {1 \over (2 \pi g \alpha')^2} \left(1 - \sqrt{1 + {(2 \pi g \alpha')^2\over2}F_{\mu \nu}F_{\mu \nu} - {(2 \pi g \alpha')^4\over4}~ \Big(F_{\mu \nu}F_{\nu \rho}F_{\rho \sigma}F_{\sigma \mu} - {1 \over 2}~\big(F_{\mu \nu}F_{\nu \mu}\big)^2\Big)} \right) \nonumber \\
&=& -{1 \over 4} F_{\mu \nu} F_{\mu \nu} + 3~ \zeta(2) g^2 \alpha'^2 \left(F_{\mu \nu}F_{\nu \rho}F_{\rho \sigma}F_{\sigma \mu} - {1 \over 4}~\big(F_{\mu \nu}F_{\nu \mu}\big)^2\right) + \Ord\left(g^4 \alpha'^4 F^6\right)
\label{BI}
\eea
was proposed in~\cite{BI} as an alternative description of electrodynamics. In the context of string scattering, the constant $\alpha'$ is identified with the string tension. A natural generalization to the case of a $U(\Nc)$ gauge group is realized~\cite{T97} when the open superstrings under consideration end on a stack of $\Nc$ coincident $D_3$-branes. This situation is unfortunately much more complicated to describe with an effective action and there is no known analog of \eqn{BI}; the non-Abelian Born-Infeld action, as it is commonly called, is only known up to fourth order~\cite{Wulff,KKNSWrev} in $\alpha'$ (reference~\cite{KKNSWrev} is a review with many more references to the original literature). For us, only the first two non-trivial orders in this expansion play an important role. Due to the fact that there is no $\Nsym$ supersymmetrizable operator of mass dimension six\footnote{In this section we use lower-case $d$ for operator dimensions and upper-case $D$ for spacetime dimensions.} ($d = 6$) that one can write down in terms of non-Abelian field strengths and covariant derivatives, the first two non-trivial orders in the $\alpha'$ expansion are actually $\Ord(\alpha'^2)$ and $\Ord(\alpha'^3)$. In our conventions, the non-Abelian Born-Infeld action is given by
\bea
\mathcal{L}_{NABI} &=& -{1\over4} \Tr \Big\{ F_{\mu \nu} F_{\mu \nu} \Big\}
\nonumber \\ && + \zeta(2) g^2 \alpha'^2 \Tr \Big\{ {1\over2}F_{\mu \nu} F_{\nu \rho}
F_{\rho \sigma} F_{\sigma \mu}
+ F_{\mu \nu} F_{\nu \rho}
F_{\sigma \mu} F_{\rho \sigma}
- {1\over8} F_{\mu \nu} F_{\rho \sigma}
F_{\nu \mu} F_{\sigma \rho}
- {1\over4} F_{\mu \nu} F_{\nu \mu}
F_{\rho \sigma} F_{\sigma \rho } \Big\}
\nonumber\\ && - 8 ~\zeta(3) \alpha'^3 \Tr \Big\{{i g^3\over \sqrt{2}} \Big( F_{\mu \nu} F_{\nu \rho}
F_{\rho \sigma} F_{\tau \mu} F_{\sigma \tau} + F_{\mu \nu} F_{\sigma \tau} F_{\nu \rho} F_{\tau \mu} F_{\rho \sigma} - {1\over2} F_{\mu \nu} F_{\nu \rho} F_{\sigma \tau} F_{\rho \mu} F_{\tau \sigma}\Big)
\nonumber \\ && + g^2 \Big( {1\over2} (D_{\mu}F_{\nu \rho}) (D_{\mu} F_{\rho \sigma}) F_{\tau \nu} F_{\sigma \tau}
+ {1\over2}(D_{\mu} F_{\nu \rho}) F_{\tau \nu} (D_{\mu} F_{\rho \sigma}) F_{\sigma \tau} - F_{\mu \nu} (D_{\mu} F_{\rho \sigma}) (D_{\tau} F_{\nu \rho}) F_{\sigma \tau}
\nonumber\\ && - {1\over8} (D_{\mu} F_{\nu \rho}) F_{\sigma \tau} (D_{\mu} F_{\rho \nu}) F_{\tau \sigma} + (D_{\tau} F_{\mu \nu}) F_{\rho \sigma} (D_{\mu} F_{\nu \rho}) F_{\sigma \tau} \Big)\Big\} + \Ord\left(\alpha'^4 \Tr F^6\right)\,.
\label{nonABI}
\eea
The form reproduced above is very nearly that given in~\cite{BMM}, but their conventions are slightly different\footnote{It also appears that their overall normalization differs from ours by a factor of two.}. We use the conventions of~\cite{Dixon96rev}, in which
\be
F_{\mu \nu} = \partial_\mu A_\nu - \partial_\nu A_\mu - {i g \over \sqrt{2}} [A_\mu, A_\nu]
\ee
and
\be
D_\mu \Phi = \partial_\mu \Phi - {i g \over \sqrt{2}} [A_\mu, \Phi] \, .
\ee
Results essentially identical to those above appeared in~\cite{KS} (see also the later work of~\cite{Drummond03}) and the derivative terms at $\Ord(\alpha'^3)$ were obtained earlier in~\cite{Bilal}. We have normalized our $\Ord(\alpha'^2)$ and $\Ord(\alpha'^3)$ effective Lagrangians so that they reproduce the appropriate terms in the $\alpha'$ expansion of the string scattering results given in~\cite{ST2}, where a representative leading four-point color-ordered partial amplitude is
\bea
A^{tree}_{str}(k_1^{1234},k_2^{1234},k_3,k_4) &=& A_{4;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} {\Gamma(1+\alpha' s)\Gamma(1+\alpha' t)\over \Gamma(1 + \alpha'(s+t))} \nonumber \\
&=& A_{4;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \bigg( 1 - \zeta(2) s t ~\alpha'^2 + \zeta(3) s t (s+t) ~\alpha'^3
\nonumber \\ && - {\zeta(4) \over 4} s t (4 s^2+s t+4 t^2)~\alpha'^4 +\Ord(\alpha'^5)\bigg) \, .
\label{ST4pt}
\eea
\subsection{New Relations}
\label{results}
We now return to the observed correspondence between the results of~\cite{oneloopselfdual} and~\cite{ST1} discussed briefly at the beginning of this section. By comparing the two references it is easy to see that
\be
A^{tree}_{str}\left(k_1^{h_1},\cdots,k_n^{h_n}\right)\Bigl|_{\Ord(\alpha'^2)} = -6 \zeta(2) A_{1}^{1-{\rm loop}}\left(k_1^{h_1},\cdots,k_n^{h_n}\right)[\mu^4]\Bigl|_{\e\to0} \, {\rm ,}
\label{oldconj}
\ee
where the gluon helicity configuration is MHV and should of course match on both sides of eq. (\ref{oldconj}).
Since our notation may not be completely obvious, we consider an illustrative example. Specifically, we check that eq. \ref{oldconj} holds for the five-gluon MHV amplitude $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)$. In terms of unevaluated scalar Feynman integrals~\cite{oneloopselfdual}, we have
\bea
A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right) &=& {- A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} \Bigg( s_{2}s_{3} I_4^{(1),~D=4-2\e}+s_{3}s_{4} I_4^{(2),~D=4-2\e}+ s_{4}s_{5} I_4^{(3),~D=4-2\e}\nonumber
\\ && +s_{5}s_{1} I_4^{(4),~D=4-2\e}+s_{1}s_{2} I_4^{(5),~D=4-2\e}-2 \e ~\pol(k_1,k_2,k_3,k_4) I_5^{D=6-2\e}\Bigg) \, {\rm .}\nn
\eea
Applying the dimension shift operation of~\cite{oneloopselfdual} to the amplitude sends $\e \to \e-2$ and $I_n^D \to I_n^D [\mu^4]$:
\bea
A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)[\mu^4] &&= {- A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2 } \Bigg( s_{2}s_{3} I_4^{(1),~D=4-2\e}[\mu^4] + s_3 s_{4} I_4^{(2),~D=4-2\e}[\mu^4]
\nonumber \\ && + s_{4}s_{5} I_4^{(3),~D=4-2\e}[\mu^4]+s_{5}s_{1} I_4^{(4),~D=4-2\e}[\mu^4] + s_{1}s_{2} I_4^{(5),~D=4-2\e}[\mu^4]
\nonumber \\ && -2 (\e-2) ~\pol(k_1,k_2,k_3,k_4) I_5^{D=6-2\e}[\mu^4] \Bigg) \, .
\eea
Applying eq. (\ref{DSmu}) gives
\bea
A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)[\mu^4] && = { A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \e (1-\e)\over 2} \Bigg( s_{2}s_{3} I_4^{(1),~D=8-2\e} +
s_{3}s_{4} I_4^{(2),~D=8-2\e} +
\nonumber \\ && s_{4}s_{5} I_4^{(3),~D=8-2\e}+s_{5}s_{1} I_4^{(4),~D=8-2\e} +
s_{1}s_{2} I_4^{(5),~D=8-2\e}
\nonumber \\ && -2 (\e-2) ~\pol(k_1,k_2,k_3,k_4) I_5^{D=10-2\e} \Bigg) \, .
\eea
Finally, we take the limit as $\e\to0$. As explained in Subsection \ref{GUD}, the terms which survive are those proportional to the ultraviolet singularities of the dimensionally-shifted basis integrals.
\bea
A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)[\mu^4]\Bigl|_{\e\to0} &&= { A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} \Bigg( s_{2}s_{3} \Big(~{1\over6}~\Big) +
s_{3}s_{4} \Big(~{1\over6}~\Big) + s_{4}s_{5} \Big(~{1\over6}~\Big)
\nonumber \\ && +s_{5}s_{1} \Big(~{1\over6}~\Big) +
s_{1}s_{2} \Big(~{1\over6}~\Big) + 4 ~\pol(1,2,3,4) \Big(~{1\over 24}~\Big) \Bigg)
\nonumber \\ && = { A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 12}
\Big( \big\{ s_{2}s_{3} \big\} + \pol(k_1,k_2,k_3,k_4) \Big) \, ,
\eea
where, following \cite{ST1}, $\big\{ s_{2}s_{3} \big\}$ represents the sum of $s_{2}s_{3}$ and its four cyclic permutations.
Finally, plugging this expression into \eqn{oldconj} gives the following prediction for \\ $A^{tree}_{str}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Bigl|_{\Ord(\alpha'^2)}$:
\be
A^{tree}_{str}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)\Bigl|_{\Ord(\alpha'^2)}= -\zeta(2) {A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2 }
\Big( \big\{ s_{2}s_{3} \big\} + \pol(k_1,k_2,k_3,k_4) \Big) \, .
\ee
By comparing to the all-$n$ result for the $\Ord(\alpha'^2)$ stringy corrections given at the beginning of this section, it is clear that the prediction of the conjecture for the $\Ord(\alpha'^2)$ piece of $A^{tree}_{str}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5\right)$ is correct. It is obvious from the above analysis that we would have been unsuccessful had we performed the dimension shift operation on the expression usually associated with the five-gluon one-loop MHV amplitude,
\bea
{- A_{5;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \over 2} \Bigg( s_{2}s_{3} I_4^{(1),~D=4-2\e}+s_{3}s_{4} I_4^{(2),~D=4-2\e}
+ s_{4}s_{5} I_4^{(3),~D=4-2\e} + \\ +s_{5}s_{1} I_4^{(4),~D=4-2\e}+s_{1}s_{2} I_4^{(5),~D=4-2\e}\Bigg) \, , \nonumber
\label{ordepsans}
\eea
illustrating that eq. (\ref{oldconj}) is only applicable if one works to all orders in $\e$ on the field theory side. We wish to stress that, although we find the language of~\cite{oneloopselfdual} convenient, we could have used the coefficients of the UV poles of $\Nsym$ one-loop MHV amplitudes considered in $D = 8 - 2 \e$ to define the right-hand side of \eqn{oldconj} and nothing would have changed, apart from maybe an unimportant overall minus sign.
Now, suppose we want to generalize the Stieberger-Taylor relation. One obvious question is whether we can relax their requirement that the helicity configuration on both sides of (\ref{oldconj}) be MHV. Indeed, we will see that the relation actually holds for general helicity configurations. Fortunately, Stieberger and Taylor calculated all six-point NMHV open superstring amplitudes in~\cite{ST4} (unfortunately not in a form as elegant as eq. (\ref{stringMHVoal2})). As a first check, we verified that $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$, $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)$, and $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4,k_5^{1234},k_6\right)$ all satisfy (\ref{oldconj}). There exists, in fact, a more general way to argue that relation (\ref{oldconj}) should be helicity-blind. Furthermore, it is possible to show that one can use all-orders-in-$\e$ $\Nsym$ Yang-Mills amplitudes to derive the $\Ord(\alpha'^3)$ stringy corrections as well. It was pointed out in~\cite{DunbarTurner} that the $\Nsym$ theory considered in $D = 8 - 2\e$ has UV divergences and that the requirements that the counterterm Lagrangian respect $\Nsym$ supersymmetry and have $d = 8$ uniquely fix it to be the $\Nsym$ supersymmetrization of ${\rm Tr}\{F^4\}$ (2nd line of eq. (\ref{nonABI})), the {\it same} operator that appears at $\Ord(\alpha'^2)$ in the non-Abelian Born-Infeld action of~\cite{T97}. Now it is clear why we found that, up to a trivial constant, one-loop $\Nsym$ gluon amplitudes dimensionally shifted to $D = 8 - 2\e$ are equal to the $\Ord(\alpha'^2)$ stringy corrections to $\Nsym$ gluon tree amplitudes: The underlying effective Lagrangians are completely determined by dimensional analysis and $\Nsym$ supersymmetry. In other words, there is only one $\Nsym$ supersymmetrizable $d = 8$ operator built out of field-strength tensors and their covariant derivatives.
This is not, however, the end of the story. That the non-Abelian Born-Infeld action is fixed to order $\Ord(\alpha'^2)$ by $\Nsym$ supersymmetry is perhaps more widely appreciated than the fact that it is fixed to order $\Ord(\alpha'^3)$ by $\Nsym$ supersymmetry. It is highly non-trivial to prove the above claim (see~\cite{KS,Collinucci}), but it is true; there is a unique $\Nsym$ supersymmetrizable linear combination of the available $d = 10$ operators (schematically, there are only two such operators, $D^2 F^4$ and $F^5$) built out of field strength tensors and their covariant derivatives. On the field theory side, Dunbar and Turner showed that $D = 10 - 2\e$ counterterm Lagrangians are built out of (an appropriate $\Nsym$ supersymmetrization of) the $d = 10$ operators $F^5$ and $D^2 F^4$. The results of~\cite{KS,Collinucci} clearly imply that this $\Nsym$ supersymmetric linear combination, being unique, coincides with the $\Ord(\alpha'^3)$ terms in the non-Abelian Born-Infeld action (eq. (\ref{nonABI})). As an additional check, we evaluated $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6\right)[\mu^6]\Big|_{\e \to 0}$ and $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6\right)[\mu^6]\Big|_{\e \to 0}$ and observed that, up to an overall factor of $60 \zeta(3)$, the results obtained precisely matched the appropriate stringy corrections (eq. (\ref{stringMHVoal2})). These observations indicate that an analogous relationship,
\be
A^{tree}_{str}\left(k_1^{h_1},\cdots,k_n^{h_n}\right)\Bigl|_{\Ord(\alpha'^3)} = 60 \zeta(3) A_{1}^{1-{\rm loop}}\left(k_1^{h_1},\cdots,k_n^{h_n}\right)[\mu^6]\Bigl|_{\e\to0}
\label{newconj}
\ee
holds in this case (again for arbitrary helicity configurations).
To summarize, we have seen that quite a bit of non-trivial information about the low-energy dynamics of open superstrings is encoded in all-orders-in-$\e$ one-loop $\Nsym$ amplitudes. At this point, one might hope that the trend continues and the stringy corrections are all somehow encoded in the $\Nsym$ theory considered in some higher dimensional spacetime. Unfortunately, there is no analog of \eqn{oldconj} and \eqn{newconj} at $\Ord(\alpha'^4)$. It is not hard to see this explicitly at the level of four-point amplitudes.
Based on our experience so far, one might expect the four-point MHV amplitude dimensionally shifted to $D=12-2\e$ to match the $\Ord(\alpha'^4)$ stringy correction given in eq. (\ref{ST4pt}) up to a multiplicative constant. However, a short calculation shows that
\be
A_{1}^{1-{\rm loop}}(k_1^{1234},k_2^{1234},k_3,k_4)[\mu^8]\Big|_{\e \to 0} = {s t (2 s^2+s t+2 t^2)\over 840} A_{4;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}}
\ee
which does {\it not} have the same $s$ and $t$ dependence as the $\Ord(\alpha'^4)$ stringy correction,
\be
A^{tree}_{str}(k_1^{1234},k_2^{1234},k_3,k_4)\Big|_{\Ord(\alpha'^4)} = - {\zeta(4) \over 4} s t (4 s^2+s t+4 t^2) A_{4;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \, .
\ee
Although it was originally hoped that $\Nsym$ supersymmetry would constrain the non-Abelian Born-Infeld action to all orders in $\alpha'$, it is now clear that this already fails to work at $\Ord(\alpha'^4)$~\cite{Collinucci}. Since it is illuminating, we repeat the argument of~\cite{Collinucci}. One can easily see that there must be more than one independent $\Nsym$ superinvariant at $\Ord(\alpha'^4)$ by comparing the $\Ord(\alpha'^4)$ terms in the Abelian Born-Infeld action to the $\Ord(\alpha'^4)$ terms in the non-Abelian Born-Infeld action responsible for the $\Ord(\alpha'^4)$ piece of the four-point tree open superstring amplitude. It is clear from eq. (\ref{BI}) that the Abelian Born-Infeld action doesn't contain any derivative terms. On the other hand, operators of the form $(D F)^4$ are the only dimension ten operators which can enter into and produce the observed $\Ord(\alpha'^4)$ four-point tree-level superstring amplitude~\cite{Bilal},
\bea
A^{tree}_{str}(k_1^{1234},k_2^{1234},k_3,k_4)\Bigl|_{\Ord(\alpha'^4)} = - {\zeta(4) \over 4} s t (4 s^2+s t+4 t^2) A_{4;\langle 12 \rangle}^{\textrm{\scriptsize{MHV}}} \, .
\label{restateordalp4}
\eea
Since the $\Ord(\alpha'^4)$ terms in the Abelian Born-Infeld action form an $\Nsym$ superinvariant by themselves (since they are present in the Abelian case where no derivative terms are allowed), the linear combination of operators of the form $(D F)^4$ responsible for the above result must be part of an distinct $\Nsym$ superinvariant.
Before leaving this section, we make one further remark about our results at $\Ord(\alpha'^2)$ that is relevant to $n$-gluon scattering. One might expect that the stringy corrections at this order in $\alpha'$ would obey a photon-decoupling relation exactly like the one in pure Yang-Mills at tree level, where replacing a single gluon by a photon produces a vanishing result. This turned out to be too simplistic. The ${\rm Tr}\{F^4\}$ operator that governs the dynamics at this order in $\alpha'$ can, in fact, couple one or even two photons to gluons. However, once you have replaced at least three external photons, the matrix elements do vanish, so long as at least one of the gluons touching the insertion of ${\rm Tr}\{F^4\}$ is off-shell. Figure \ref{ThreePhotonFigure} illustrates how this works.
For example, replacing, for sake of argument, gluons $k_2^{1234}$, $k_3$, and $k_4$ by photons results in the identity
\bea
&& 0 = A_{str}^{tree}(k_1^{1234},k_2^{1234},k_3,k_4,k_5)\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_2^{1234},k_4,k_3,k_5)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_3,k_4,k_2^{1234},k_5)\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_3,k_2^{1234},k_4,k_5)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_4,k_3,k_2^{1234},k_5)\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_4,k_2^{1234},k_3,k_5)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_2^{1234},k_3,k_5,k_4)\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_2^{1234},k_4,k_5,k_3)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_3,k_4,k_5,k_2^{1234})\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_3,k_2^{1234},k_5,k_4)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_4,k_3,k_5,k_2^{1234})\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_4,k_2^{1234},k_5,k_3)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_2^{1234},k_5,k_3,k_4)\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_2^{1234},k_5,k_4,k_3)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_3,k_5,k_4,k_2^{1234})\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_3,k_5,k_2^{1234},k_4)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_4,k_5,k_3,k_2^{1234})\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_4,k_5,k_2^{1234},k_3)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_5,k_2^{1234},k_3,k_4)\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_5,k_2^{1234},k_4,k_3)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_5,k_3,k_4,k_2^{1234})\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_5,k_3,k_2^{1234},k_4)\Big|_{\Ord(\alpha'^2)}
\nonumber \\ && + A_{str}^{tree}(k_1^{1234},k_5,k_4,k_3,k_2^{1234})\Big|_{\Ord(\alpha'^2)}+A_{str}^{tree}(k_1^{1234},k_5,k_4,k_2^{1234},k_3)\Big|_{\Ord(\alpha'^2)}\, {\rm .}
\eea
An immediate out-growth of our three-photon decoupling relation for ${\rm Tr}\{F^4\}$ matrix elements is a plausible explanation of the observation~\cite{allplus} that, for the all-plus helicity configuration at one loop in pure Yang-Mills, replacing three gluons by photons always gives zero for the five- and higher-point amplitudes. Stieberger and Taylor showed that MHV ${\rm Tr}\{F^4\}$ matrix elements are closely related to the all-plus one-loop pure Yang-Mills amplitudes and, therefore, it is reasonable to expect the photon-decoupling identity discussed above for ${\rm Tr}\{F^4\}$ matrix elements to carry over to the all-plus one-loop pure Yang-Mills amplitudes as well.
\FIGURE{
\resizebox{0.70\textwidth}{!}{\includegraphics{threephotonvanish.eps}}
\vbox{\vskip 1 in}
\caption{Matrix elements of ${\rm Tr}\{F^4\}$ for a number of gluons and
(a) a single photon, or (b) two photons, can be non-vanishing, as explained
above. On the other hand, matrix elements with (c) three or more
photons have to vanish (except for the case $n=4$ with four photons).}
\label{ThreePhotonFigure}}
\section{On-Shell Superspace and All-Orders One-Loop $\Nsym$ Superamplitudes}
\label{supercomp}
So far in this paper, $\Nsym$ supersymmetry has played a somewhat peripheral role in that all results have been presented in component form and we have rather arbitrarily focused on the computation of $n$-gluon scattering amplitudes. In this section we discuss a powerful formalism that unifies all amplitudes with a given $k$-charge\footnote{Recall that, so far, we have defined the $k$-charge of an amplitude operationally as how many complete copies of the set $\{1,2,3,4\}$ appear in the helicity labels of the amplitude's external lines ({\it e.g.} $\a(k_1,k_2^{123},k_3^4,k_4,k_5^{12},k_6^{34})$ has $k$-charge two).}. We will see that this is naturally accomplished by introducing auxiliary Grassmann variables. Essentially, the point of introducing these variables is that it facilitates a better understanding of the symmetries of the S-matrix and allows one to make more of them manifest. The goal is to replace the scattering amplitudes studied so far (amplitudes with a particular $SU(4)_R$ index structure) with $\Nsym$ {\it superamplitudes}, $SU(4)_R$ invariant objects that contain as components all $\Nsym$ amplitudes of a given multiplicity, $n$. Writing results in terms of superamplitudes will also allow us to give a less heuristic definition of $k$-charge. Indeed, this section should make it much more clear why we use $SU(4)_R$ indices to label component $\Nsym$ scattering amplitudes.
\subsection{General Discussion of $\Nsym$ On-Shell Superspace}
\label{gendisonshell}
The usual construction of the massless $\Nsym$ supermultiplet begins with the anticommutator of the supercharges
\bea
\{Q^a_{~\alpha}, \bar{Q}_{b~\dot{\alpha}}\} &=& \D_b^{\,~a}~ P_{\alpha \dot{\alpha}}\elale
\D_b^{\,~a}~ \lambda_{\alpha} \tilde{\lambda}_{\dot{\alpha}}\,,
\label{fundsusyalg}
\eea
in which one chooses $p^\mu = ({\bf p},0,0,{\bf p})$ to define a preferred reference frame. It then follows that some of the supercharges anticommute with themselves and everything else in this frame and others act as creation ($\bar{Q}_{b~\dot{1}}$) or annihilation ($Q^a_{~1}$) operators on the space of states. This approach is useful for some purposes ({\it e.g.} determining the particle content of the massless supermultiplet) but to describe scattering it is better to try and build a formalism where the supercharges act as creation and annihilation operators on the space of states in a manifestly Lorentz covariant way.\footnote{This alternative approach is not new~\cite{Nair,Sokatchev}, but its power was not properly appreciated until very recently~\cite{DHKSdualconf,EFK,SUSYBCFW}.} Our goal is readily accomplished if we introduce a set of four Grassmann variables, $\{\eta_\ell^1,\eta_\ell^2,\eta_\ell^3,\eta_\ell^4\}$, for each external four-vector, $p_\ell$, in the problem. Then one can easily see that (suppressing the $\ell$ label for now)
\be
Q^a_{~\alpha}= \lambda_{\alpha} \eta^a \qquad {\rm and} \qquad \bar{Q}_{b\,\dot{\alpha}} = \tilde{\lambda}_{\dot{\alpha}} {\partial \over \partial \eta^b}
\label{susycharges}
\ee
together satisfy (\ref{fundsusyalg}). Furthermore, the introduction of the $\eta^a$ allows one to build a super wavefunction (Grassmann coherent state) for each external line
\begin{eqnarray}
\Phi(p,\eta) &=& G^{+}(p) + \G_a(p) \eta^a + \frac{1}{2!2!}\e_{a b c d}S^{a c}(p)\eta^b \eta^d
+ \frac{1}{3!} \e_{a b c d}\bar{\G}^{a}(p)\eta^b\eta^c\eta^d \el
+ \frac{1}{4!}\e_{a b c d} G^{-}(p) \eta^a\eta^b\eta^c \eta^d\,.
\label{supwvfun}
\end{eqnarray}
which makes it possible to consider $\Nsym$ scattering with half of the supersymmetries (the $Q^a_{~\alpha}$ supercharges which we have chosen to implement multiplicatively) made manifest. To convince the reader that (\ref{susycharges}) and (\ref{supwvfun}) make sense, we must construct the covariant analogs of $Q^i_{~1}$ and $\bar{Q}_{j~\dot{1}}$ in the traditional, non-covariant approach ({\it i.e.} we need to identify the relevant creation operators). In fact, given that $Q^a_{~\alpha} \lambda^\alpha = 0$ and $\bar{Q}_{b~\dot{\alpha}}\tilde{\lambda}^{\dot{\alpha}} = 0$ (the supercharges only have components parallel to $\lambda^\alpha$ and $\tilde{\lambda}^{\dot{\alpha}}$), we can read off the analogs of $Q^i_{~1}$ and $\bar{Q}_{j~\dot{1}}$: the annihilation and creation operators are simply the components of the supercharges along the directions of the spinors, $\hat{a}^c \equiv Q^c = \eta^c$ and $\hat{a}_d^{\dagger} \equiv \bar{Q}_{d} = \partial/\partial \eta^d$, and they satisfy the algebra
\be
\{Q^c, \bar{Q}_{d}\} = \D^{~\,c}_{d} \,{\rm .}
\ee
Now that we know what the creation operators are we can act on the super wavefunction (\ref{supwvfun}) in various combinations. All that we have to do to show that (\ref{supwvfun}) is complete and correct is find some combination of creation operators ($\eta$ derivatives) that isolate each term in the super wavefunction. Following~\cite{DHKSdualconf,origElvang}, we have
\bea
&& \Phi(p,\eta)\Big|_{\eta^n = 0} = G^{+}(p) \qquad \bar{Q}_{a} \Phi(p,\eta)\Big|_{\eta^{n} = 0} = \G_{a}(p) \qquad {1\over 2!}\bar{Q}_{a} \bar{Q}_{b} \e^{a b c d} \Phi(p,\eta)\Big|_{\eta^n = 0} = S^{c d}(p)
\el {1\over 3!} \bar{Q}_{a} \bar{Q}_{b} \bar{Q}_{c} \e^{a b c d} \Phi(p,\eta)\Big|_{\eta^n = 0} = \bar{\G}^{d}(p) \qquad
{1\over 4!} \bar{Q}_{a} \bar{Q}_{b} \bar{Q}_{c} \bar{Q}_{d} \e^{a b c d} \Phi(p,\eta)\Big|_{\eta^n = 0} = G^{-}(p) \,{\rm .}\nn
\eea
Evidently, our on-shell superspace construction is well-defined and it therefore makes sense to speak about $\Nsym$ on-shell superamplitudes, $\a(\Phi_1,\cdots, \Phi_n)$, that take into account all elements of the planar\footnote{Clearly, at the moment, this is a choice we are making since supersymmetry commutes with color.} S-matrix with $n$ external states simultaneously. The $n$-point superamplitude is naturally expanded into $k$-charge sectors as\footnote{As discussed in Appendix \ref{SWI}, the $k = 0$ and $k = 1$ sectors (and by parity the $k = n - 1$ and $k = n$ sectors) are identically zero for non-degenerate kinematical configurations.}
\begin{equation}
\a(\Phi_1,\cdots, \Phi_n) = \a_{n;2} + \a_{n;3} + \cdots + \a_{n;\, n-2} \,{\rm .}
\label{kexpans}
\end{equation}
So far, we have defined $k$-charge at the level of component amplitudes. For example, $A(p_1^{1234},p_2^{1234},p_3,p_4)$ and $A(p_1^{1234},p_2^{123},p_3^{4},p_4)$ both have $k$-charge two because one needs two copies of $\{1,2,3,4\}$ to label their external states. At the level of the superamplitude, the $k$-charge of a given term on the right-hand side of (\ref{kexpans}) is determined by the number of Grassmann parameters that appear in it divided by four\footnote{The $SU(4)_R$ rotates the Grassmann parameters into each other and the superamplitude must be a singlet under R-symmetry transformations. This is impossible unless, for a given term, each $SU(4)$ index, $\{1,2,3,4\}$, appears the same number of times.}. We will refer to $\a_{n;k}$ (a $k$-charge sector of the superamplitude) as a superamplitude since there is usually no possibility of confusion.
We now turn to the MHV tree-level superamplitude, $\a_{n;2}^{tree}$, which has the simplest superspace structure and can be completely determined by matching onto the Parke-Taylor formula (or any other component amplitude for that matter). Clearly, the simplest possible superspace structure is given by the eight-fold Grassmann delta function itself and corresponds to the first term on the right-hand side of (\ref{kexpansD}),
\be
\a_{n;2}^{tree} = {1\over 16}\prod_{a = 1}^4 \sum_{i,j = 1}^n \spa{i}.j \eta^a_i \eta^a_j~~ \hat{\a}_{n;2}^{tree} \,{\rm ,}
\label{undetMHVsup}
\ee
where we have used the explicit formula for the Grassmann delta function $\D^{(8)}\left(Q^{a\,\A}\right)$ derived in Appendix \ref{SWI}. Suppose we are interested in computing $A^{tree}\left(p_1^{1234},p_2^{1234},p_3,\cdots,p_n\right)$ using $\a_{n;2}^{tree}$. To compute this amplitude one expands (\ref{undetMHVsup}) and extracts the coefficient of $\eta_1^1\eta_1^2\eta_1^3\eta_1^4\eta_2^1\eta_2^2\eta_2^3\eta_2^4$. We will denote this combination as $\eta_1^{1234} \eta_2^{1234}$. The result of this calculation is the numerator of the familiar Parke-Taylor amplitude times $\hat{\a}_{n;2}^{tree}$:
\be
A^{tree}\left(p_1^{1234},p_2^{1234},p_3,\cdots,p_n\right) = \spa1.2^4~~ \hat{\a}_{n;2}^{tree} \,{\rm .}
\ee
It follows that
\be
\hat{\a}_{n;2}^{tree} = {i \over \spa1.2 \spa2.3 \cdots \spa{n}.1}
\ee
and
\be
\a_{n;2}^{tree} = i{{1\over 16} \prod_{a = 1}^4 \sum_{i,j = 1}^n \spa{i}.j \eta^a_i \eta^a_j \over \spa1.2 \spa2.3 \cdots \spa{n}.1} \,{\rm .}
\label{MHVsup}
\ee
We have successfully given a unified description of all MHV tree amplitudes in $\Nsym$. In fact, for appropriate supersymmetry-preserving variants of dimensional regularization\footnote{We refer the interested reader to Subappendix \ref{4DHS}, where we describe the four dimensional helicity scheme, the particular variant used in most multi-loop studies of $\Nsym$ scattering amplitudes.}, it turns out that the superspace structure in the MHV case is independent of the loop expansion~\cite{oneloopselfdual,dualSmat} and we can write
\be
\a_{n;2} = i{{1\over 16} \prod_{a = 1}^4 \sum_{i,j = 1}^n \spa{i}.j \eta^a_i \eta^a_j \over \spa1.2 \spa2.3 \cdots \spa{n}.1}\Bigg(1+\left({g^2 \Nc \mu^{2\e} e^{-\gamma_E \e} \over (4 \pi)^{2-\e}}\right) M_{1-{\rm loop}}+ \left({g^2 \Nc \mu^{2\e} e^{-\gamma_E \e}\over (4 \pi)^{2-\e}}\right)^2 M_{2-{\rm loop}}+\cdots\Bigg)
\label{MHVsupL}
\ee
as well, although the determination of $M_{L-{\rm loop}}$ may be quite non-trivial\footnote{It is important to point out here that there is a natural seperation of the $M_{L-{\rm loop}}$ functions into even and odd components.}. In the above we still suppress the tree-level gauge coupling and color structure, worrying only about relative factors between different loop orders.
Before moving on to more non-trivial examples, we point out an important subtlety related to the special case of $n = 3$ in eq. (\ref{MHVsup}). Our experience with scattering amplitudes suggests that we should also be able to define the MHV and anti-MHV three-point superamplitudes, $\a_{3;2}$ and $\bar{\a}_{3;2}$, even though, na\"{i}vely, it would appear that any superamplitude with four Grassmann variables in each term must vanish due to supercharge conservation (Appendix \ref{SWI}). After all, the Grassmann polynomial that expresses supercharge conservation already has eight Grassmann variables in each term. There is a way out, however, if one allows for degenerate kinematics. For three-point kinematics we have $p_1 = - p_2 - p_3$ which implies that $0 = p_1^2 = \spa2.3 \spb3.2$. Making the choice $\spb2.3 = 0$ and $\spa3.2 \neq 0$ leads to the three-point MHV superamplitude (setting $n = 3$ in eq. (\ref{MHVsup})) and making the choice $\spa2.3 = 0$ and $\spb3.2 \neq 0$ will lead us to the anti-MHV three-point superamplitude. Setting all of the holomorphic spinor products to zero in the three-point amplitude forces all of the holomorphic spinors to be proportional to one another. Quantitatively, this means that $\lambda_{\alpha\,\ell} = c_{\ell} \chi_\alpha$ for some spinor $\chi_\alpha$ and coefficients $c_{\ell}$. Consequently, the total supercharge can be written as $Q^a_{~\,\alpha} = \chi_\alpha \sum_{\ell = 1}^3 c_\ell \eta^a_\ell$. The point is that now the $\alpha$ dependence just sits in the spinor $\chi_\alpha$ which pulls out of the sum over $\ell$. Overall factors inside delta functions can't lead to non-trivial constraints and it follows that the overall supercharge conserving delta function is only four-fold in this special case. It is very instructive to realize this discussion explicitly and determine the superspace structure of $\bar{\a}_{3;2}$.
We start with momentum conservation
\be
\lambda_{\alpha\,1}\, \tilde{\lambda}_{\dot{\alpha}\,1}+\lambda_{\alpha\,2}\, \tilde{\lambda}_{\dot{\alpha}\,2}+\lambda_{\alpha\,3}\, \tilde{\lambda}_{\dot{\alpha}\,3} = 0
\ee
and project by taking the spinor product of this equation with, say, $\tilde{\lambda}_{\dot{\alpha}\,1}$:
\be
\lambda_{\alpha\,2}\spb2.1 + \lambda_{\alpha\,3}\spb3.1 = 0 \,{\rm .}
\ee
Permuting labels, we can also write
\be
\lambda_{\alpha\,1}\spb1.3 + \lambda_{\alpha\,2}\spb2.3 = 0 \,{\rm .}
\ee
Solving for $\lambda_{\alpha\,1}$ and $\lambda_{\alpha\,3}$ in terms of $\lambda_{\alpha\,2}$, we find
\be
Q^a_{~\,\alpha} = \sum_{\ell = 1}^3 \lambda_{\alpha\,\ell}\,\eta^a_\ell = -\lambda_{\alpha\,2}{\spb2.3 \eta_1^a \over \spb1.3} + \lambda_{\alpha\,2} \eta_2^a -\lambda_{\alpha\,2}{\spb2.1 \eta_3^a \over \spb3.1} = \lambda_{\alpha\,2}\left({\spb2.3 \eta^a_1 + \spb3.1\, \eta^a_2 + \spb1.2 \eta^a_3 \over \spb3.1}\right)
\ee
and the arguments of the last paragraph imply that the superspace structure of $\bar{\a}_{3;2}$ is simply
\bea
\D^{(4)}\left(\spb2.3 \eta^a_1 + \spb3.1\, \eta^a_2 + \spb1.2\, \eta^a_3\right) &=& \left(\spb2.3 \eta^1_1 + \spb3.1\, \eta^1_2 + \spb1.2\, \eta^1_3\right)\left(\spb2.3 \eta^2_1 + \spb3.1\, \eta^2_2 + \spb1.2\, \eta^2_3\right) \times
\el \times \left(\spb2.3 \eta^3_1 + \spb3.1\, \eta^3_2 + \spb1.2 \,\eta^3_3\right)\left(\spb2.3 \eta^4_1 + \spb3.1\, \eta^4_2 + \spb1.2 \,\eta^4_3\right)
\elale \prod_{a = 1}^4 \left(\spb2.3 \,\eta^a_1 + \spb3.1\, \eta^a_2 + \spb1.2\, \eta^a_3\right)\,{\rm .}\nn
\eea
Now that the superspace structure of $\bar{\a}_{3;2}$ is fixed, it is a simple matter to match onto, say, the anti-MHV three-gluon amplitude and determine the entire superamplitude. As it stands, we have
\be
\bar{\a}_{3;2} = \bar{\a}_{3;2}^\prime \prod_{a = 1}^4 \left(\spb2.3 \,\eta^a_1 + \spb3.1\, \eta^a_2 + \spb1.2\, \eta^a_3\right)
\label{partanti3}
\ee
Expanding out both sides of eq. (\ref{partanti3}) for, say, $A^{tree}\left(k_1^{1234}, k_2, k_3\right)$ and extracting the coefficient of $\eta_1^{1234}$, we find
\bea
A^{tree}\left(k_1^{1234}, k_2, k_3\right) &=& \bar{\a}_{3;2}^\prime \prod_{a = 1}^4 \left(\spb2.3\, \eta^a_1 + \spb3.1\, \eta^a_2 + \spb1.2\, \eta^a_3\right)\nonumber \\
{i \spb2.3^4 \over \spb2.3 \spb3.1 \spb1.2} &=& \bar{\a}_{3;2}^\prime \spb2.3^4 \nonumber \\
{i \over \spb2.3 \spb3.1 \spb1.2} &=& \bar{\a}_{3;2}^\prime \,{\rm .}
\eea
Thus, we finally have
\be
\bar{\a}_{3;2} = i {\prod_{a = 1}^4 \left(\spb2.3 \,\eta^a_1 + \spb3.1\, \eta^a_2 + \spb1.2\, \eta^a_3\right) \over \spb1.2 \spb2.3 \spb3.1}\,{\rm .}
\ee
As we will see, the superspace structure of the six-point NMHV superamplitude is in some sense built out of pieces of $\bar{\a}_{3;2}$. The superamplitude $\a_{6;3}$ is of particular interest for us because it represents the desired supersymmetrization of the results derived in Section \ref{gluoncomp}. To proceed, we need to know the superspace structure of $\a_{6;3}$ and which component amplitudes are required to nail it down. Happily, this difficult problem was solved in a recent paper by Elvang, Freedman, and Kiermaier~\cite{EFK}. Here we simply present their formula for $\a_{6;3}$ and refer the reader interested in the derivation to \ref{superNMHV}. The result is
\bea
&&\a_{6;\,3}^{1-{\rm loop}} = {\D^{(8)}(Q^{a\,\A}) \over \spb3.4^4 \spa5.6^4}\Big(A_1^{1-{\rm loop}}\left(p_1^{1234}, p_2, p_3, p_4, p_5^{1234}, p_6^{1234}\right) \prod_{a=1}^4 \left(\spb3.4~\eta_1^a+\spb4.1 ~\eta_3^a +\spb{1}.3~ \eta_4^a\right)
\el+ A_1^{1-{\rm loop}}\left(p_1^{123},p_2^4,p_3,p_4,p_5^{1234},p_6^{1234}\right) \prod_{a=1}^3 \left(\spb3.4~\eta_1^a+\spb4.1 ~\eta_3^a +\spb{1}.3~ \eta_4^a\right)\left(\spb3.4~\eta_2^4+\spb4.2 ~\eta_3^4 +\spb{2}.3~ \eta_4^4\right)
\el+ A_1^{1-{\rm loop}}\left(p_1^{12},p_2^{34},p_3,p_4,p_5^{1234},p_6^{1234}\right)\prod_{a=1}^2 \left(\spb3.4~\eta_1^a+\spb4.1 ~\eta_3^a +\spb{1}.3~ \eta_4^a\right)\prod_{a=3}^4\left(\spb3.4~\eta_2^a+\spb4.2 ~\eta_3^a +\spb{2}.3~ \eta_4^a\right)
\el + A_1^{1-{\rm loop}}\left(p_1^1,p_2^{234},p_3,p_4,p_5^{1234},p_6^{1234}\right) \left(\spb3.4~\eta_1^1+\spb4.1 ~\eta_3^1 +\spb{1}.3~ \eta_4^1\right)\prod_{a=2}^4\left(\spb3.4~\eta_2^a+\spb4.2 ~\eta_3^a +\spb{2}.3~ \eta_4^a\right)
\el + A_1^{1-{\rm loop}}\left(p_1, p_2^{1234}, p_3, p_4, p_5^{1234}, p_6^{1234}\right) \prod_{a=1}^4\left(\spb3.4~\eta_2^a+\spb4.2 ~\eta_3^a +\spb{2}.3~ \eta_4^a\right)\Big)\,{\rm .}\nn
\label{A63}
\eea
It is also well-known that, when used in conjunction with the four dimensional helicity scheme~\cite{4dimHS}, results derived from supersymmetry such as the above hold order-by-order in the dimensional regularization parameter. If we want to use the above formula to give a supersymmetrized version of the all-orders-in-$\e$ six-point NMHV superamplitude, we have to somehow determine the pentagon coefficients of $A_1^{1-{\rm loop}}\left(p_1^{123},p_2^4,p_3,p_4,p_5^{1234},p_6^{1234}\right)$, $A_1^{1-{\rm loop}}\left(p_1^{12},p_2^{34},p_3,p_4,p_5^{1234},p_6^{1234}\right)$, and $A_1^{1-{\rm loop}}\left(p_1^1,p_2^{234},p_3,p_4,p_5^{1234},p_6^{1234}\right)$ (the box coefficients for NMHV amplitudes with pairs of scalars or fermions were already computed in~\cite{Risager} or can be deduced from~\cite{Huang}). In this paper we are focused on computing the unknown pentagon coefficients and we will therefore suppress the box contributions to $\Nsym$ amplitudes throughout the rest of this work except when it is desirable to include them for pedagogical purposes.
Now, there are two ways that we could try and go forward. One would be to try and match the superamplitude onto three different pure-gluon amplitudes that we have already computed and solve a linear system to determine the eighteen unknown pentagon coefficients (one system of three equations in three unknowns suffices to determine all eighteen unknowns by virtue of the leading singularity method). Alternatively, we could try and determine the coefficients directly by combining the leading singularity method with $D$ dimensional unitarity, as was done in Section \ref{gluoncomp} for six-gluon amplitudes. Either way we will have to simplify the rather complicated results obtained using BCFW shifts. In this section we try both approaches although it would probably be easier to use the first to determine everything. In Subsection \ref{ffbarcalc}, we compute amplitudes with fermion/anti-fermion pairs directly because, if we'd like to claim that our approach to $D$ dimensional integrand reconstruction is applicable in principle to one-loop QCD calculations, it is important to see that the calculational method discussed in Section \ref{gluoncomp} can handle amplitudes where some of the external gluons have been replaced by fermions. For the final amplitude with a scalar/anti-scalar pair we obtain the result in Subsection \ref{ssbarcalc} by matching onto the result we derived in Section \ref{gluoncomp} for $A_1^{1-{\rm loop}}\left(p_1^{1234},p_2,p_3^{1234},p_4,p_5^{1234},p_6\right)$.
\subsection{NMHV Amplitudes With a Fermion/Anti-Fermion Pair}
\label{ffbarcalc}
The approach that we use in this subsection to calculate the higher-order in $\e$ pentagon coefficients for $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^{4},k_5,k_6\right)$ is completely analogous to the approach outlined in Subsection \ref{effgcomp} for six-gluon amplitudes. There are, of course, a few minor differences. The diagrammatics are a bit less obvious (see Figure \ref{ffbardiags}) and some of the three-point tree amplitudes used to reconstruct the one-loop integrands are different. We will need the three-point vertex for a fermion, anti-fermion, and a complex scalar in $D$ dimensions. This is just a Yukawa coupling:
\be
A^{tree}_{\mu^2}(-p_{\bar{f}},k_1^{12},(p-k_1)_f) = i \bar{u}(p)u(p-k_1) \,{\rm .}
\ee
As the attentive reader may have suspected, we are actually going to compute \\$A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)$ instead of $A_1^{1-{\rm loop}}\left(k_1^{1},k_2^{234},k_3,k_4,k_5^{1234},k_6^{1234}\right)$ to make life as easy as possible. If we compute $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)$, it will turn out that we can recycle the four-vertices (eqs. (\ref{mu4tree1}) - (\ref{mu4tree3})) that we used in our six-gluon calculations. All we have to do is rewrite eq. (\ref{A63}) with shifted component amplitudes. Going through the proof in \ref{superNMHV}, we see that this is an entirely straightforward exercise and we arrive at
\FIGURE{
\resizebox{.95\textwidth}{!}{\includegraphics{ffbardiags.eps}}
\caption{The generalized unitarity diagrams contributing to the $I_5^{(1)}$ pentagon coefficient of $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^{4},k_5,k_6\right)$.}
\label{ffbardiags}}
\bea
&&\a_{6;\,3}^{1-{\rm loop}} = {\D^{(8)}(Q^{a\,\A}) \over \spb5.6^4 \spa1.2^4}\Big(A_1^{1-{\rm loop}}\left(k_1^{1234}, k_2^{1234}, k_3^{1234}, k_4, k_5, k_6\right) \prod_{a=1}^4 \left(\spb5.6~\eta_3^a+\spb6.3 ~\eta_5^a +\spb{3}.5~ \eta_6^a\right)
\el+ A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right) \prod_{a=1}^3 \left(\spb5.6~\eta_3^a+\spb6.3 ~\eta_5^a +\spb{3}.5~ \eta_6^a\right)\left(\spb5.6~\eta_4^4+\spb6.4 ~\eta_5^4 +\spb{4}.5~ \eta_6^4\right)
\el+ A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{12},k_4^{34},k_5,k_6\right)\prod_{a=1}^2 \left(\spb5.6~\eta_3^a+\spb6.3 ~\eta_5^a +\spb{3}.5~ \eta_6^a\right)\prod_{a=3}^4\left(\spb5.6~\eta_4^a+\spb6.4 ~\eta_5^a +\spb{4}.5~ \eta_6^a\right)
\el + A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right) \left(\spb5.6~\eta_3^1+\spb6.3 ~\eta_5^1 +\spb{3}.5~ \eta_6^1\right)\prod_{a=2}^4\left(\spb5.6~\eta_4^a+\spb6.4 ~\eta_5^a +\spb{4}.5~ \eta_6^a\right)
\el + A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right) \prod_{a=1}^4\left(\spb5.6~\eta_4^a+\spb6.4 ~\eta_5^a +\spb{4}.5~ \eta_6^a\right)\Big)
\label{A63s}
\eea
with very little effort.
Also, when we say compute the amplitude, what we really mean is compute the pentagon coefficient of $I_5^{(1);~D = 6 - 2\e}$. Once this function is determined, the leading singularity equations give the rest of the pentagon coefficients for free. We denote the first diagram in Figure \ref{ffbardiags} $\mathcal{M}_I$, the second $\mathcal{M}_{II}$ and so forth. We present each of these contributions in turn:
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1;\,\mathcal{M}_I}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}} = -\left({-i \over \sqrt{2}}\right)^2
\el\pol^\alpha(p_*+k_5+k_6)[4|\gamma_\alpha u(p_*-k_1-k_2-k_3) \bar{u}(p_*-k_1-k_2-k_3)\gamma_\beta |3\rangle\pol^{*\,\beta}(p_*-k_1-k_2)
\el i \sqrt{2}\pol^{*\,\rho_5}(p_*+k_5+k_6) \left(\pol^+(k_5)\cdot (p_*+k_5+k_6)~ g_{\rho_5 \sigma_5}+k_{5\,\rho_5}~\pol^+_{\sigma_5}(k_5)-k_{5\,\sigma_5}~ \pol^+_{\rho_5}(k_5)\right)\pol^{\sigma_5}(p_*+k_6)
\el \pol^{*\,\sigma}(p_*+k_6) {-2 i \over \spab1.{p_*}.1}\Big(\pol^-(k_1)\cdot p_* ~\pol^+(k_6)\cdot p_* ~g_{\rho \sigma}+ \pol^+(k_6)\cdot p_* ~k_{1\,\sigma} \pol^-_\rho(k_1)
\nonumber\\&&-\pol^+(k_6)\cdot p_* ~k_{1\,\rho}~ \pol^-_\sigma(k_1) + \pol^-(k_1)\cdot p_* ~k_{6\,\sigma}~ \pol^+_{\rho}(k_6)
\nonumber\\&&- \pol^-(k_1)\cdot p_* ~k_{6\,\rho}~ \pol^+_{\sigma}(k_6) - k_1\cdot k_6 \pol^-_\rho(k_1)\pol^+_\sigma(k_6)\Big)\pol^\rho(p_*-k_1)
\el i \sqrt{2}\pol^{*\,\rho_2}(p_*-k_1) \left(\pol^-(k_2)\cdot (p_*-k_1)~ g_{\rho_2 \sigma_2}+k_{2\,\rho_2}~\pol^-_{\sigma_2}(k_2)-k_{2\,\sigma_2}~ \pol^-_{\rho_2}(k_2)\right)\pol^{\sigma_2}(p_*-k_1-k_2)
\elale {2 i \over \spab1.{p_*}.1} g^{\rho_5 \alpha}[4|\gamma_\alpha (\slashed{p}_*-\slashed{k}_1-\slashed{k}_2-\slashed{k}_3+\mu)\gamma_\beta |3\rangle g^{\sigma_2 \beta}
\el\left(\pol^+(k_5)\cdot (p_*+k_5+k_6)~ g_{\rho_5 \sigma_5}+k_{5\,\rho_5}~\pol^+_{\sigma_5}(k_5)-k_{5\,\sigma_5}~ \pol^+_{\rho_5}(k_5)\right)g^{\sigma_5 \sigma}
\el\Big(\pol^-(k_1)\cdot p_* ~\pol^+(k_6)\cdot p_* ~g_{\rho \sigma}+ \pol^+(k_6)\cdot p_* ~k_{1\,\sigma} \pol^-_\rho(k_1)
\nonumber\\&&-\pol^+(k_6)\cdot p_* ~k_{1\,\rho}~ \pol^-_\sigma(k_1) + \pol^-(k_1)\cdot p_* ~k_{6\,\sigma}~ \pol^+_{\rho}(k_6)
\nonumber\\&&- \pol^-(k_1)\cdot p_* ~k_{6\,\rho}~ \pol^+_{\sigma}(k_6) - k_1\cdot k_6 \pol^-_\rho(k_1)\pol^+_\sigma(k_6)\Big)
\el g^{\rho \rho_2} \left(\pol^-(k_2)\cdot (p_*-k_1)~ g_{\rho_2 \sigma_2}+k_{2\,\rho_2}~\pol^-_{\sigma_2}(k_2)-k_{2\,\sigma_2}~ \pol^-_{\rho_2}(k_2)\right)\,{\rm ,}\nn
\label{MI}
\eea
\end{changemargin}
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1;\,\mathcal{M}_{II}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}} = \left({-i \over \sqrt{2}}\right)^4 [4|\gamma_\alpha \pol^{*\,\alpha}(p_*-k_1-k_2-k_3)u(p_*+k_5+k_6)\times
\el \times \bar{u}(p_*+k_5+k_6)\slashed{\pol}^+(k_5)u(p_*+k_6)\left({i (p_*+k_6)\cdot \pol^+(k_6) \over \spab1.{p_*}.1} \bar{u}(p_*+k_6) \slashed{\pol}^-(k_1)u(p_*-k_1)\right)\times
\el \times\bar{u}(p_*-k_1)\slashed{\pol}^-(k_2)u(p_*-k_1-k_2)\bar{u}(p_*-k_1-k_2)\gamma_\beta \pol^{\beta}(p_*-k_1-k_2-k_3) |3\rangle
\elale -{i (p_*+k_6)\cdot \pol^+(k_6) \over 4 \spab1.{p_*}.1} [4|\gamma^\beta(\slashed{p}_*+\slashed{k}_5+\slashed{k}_6+\mu)\slashed{\pol}^+(k_5)(\slashed{p}_*+\slashed{k}_6+\mu)\slashed{\pol}^-(k_1)(\slashed{p}_*-\slashed{k}_1+\mu)\times
\el\times \slashed{\pol}^-(k_2)(\slashed{p}_*-\slashed{k}_1-\slashed{k}_2+\mu)\gamma_\beta |3\rangle \,{\rm ,}\nn
\label{MII}
\eea
\end{changemargin}
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1;\,\mathcal{M}_{III}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}} = -(i)^2 [4|u(p_*-k_1-k_2-k_3)\bar{u}(p_*-k_1-k_2-k_3)|3\rangle \times
\el
\left(-\sqrt{2} i (p_*+k_5+k_6)\cdot \pol^+(k_5)\right) {-i \spab1.{p_*}.6^2 \over s_6 \spab1.{p_*}.1} \left(-\sqrt{2} i (p_*-k_1)\cdot \pol^-(k_2)\right)
\elale 2 i [4|\slashed{p}_*-\slashed{k}_1-\slashed{k}_2-\slashed{k}_3+\mu|3\rangle
(p_*+k_5+k_6)\cdot \pol^+(k_5){\spab1.{p_*}.6^2 \over s_6 \spab1.{p_*}.1}(p_*-k_1)\cdot \pol^-(k_2)\,{\rm ,}
\label{MIII}
\eea
\end{changemargin}
\begin{changemargin}{-.6 in}{0 in}
\bea
&&A_{1;\,\mathcal{M}_{IV}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}} = (i)^2 [4| u(p_*+k_5+k_6) \left( -{i\over \sqrt{2}}\bar{u}(p_*+k_5+k_6) \slashed{\pol}^+(k_5) u(p_*+k_6)\right) \times
\el {i (p_*+k_6)\cdot \pol^+(k_6)\over \spab1.{p_*}.1}\bar{u}(p_*+k_6)\slashed{\pol}^-(k_1) u(p_*-k_1)\left(-{i\over \sqrt{2}}\bar{u}(p_*-k_1)\slashed{\pol}^-(k_2)u(p_* - k_1 - k_2)\right) \bar{u}(p_*-k_1-k_2)|3\rangle
\el = {i (p_*+k_6)\cdot \pol^+(k_6)\over 2 \spab1.{p_*}.1}[4| (\slashed{p}_*+\slashed{k}_5+\slashed{k}_6)\slashed{\pol}^+(k_5) (\slashed{p}_*+\slashed{k}_6)\slashed{\pol}^-(k_1) (\slashed{p}_*-\slashed{k}_1)\slashed{\pol}^-(k_2)\slashed{p}_*-\slashed{k}_1-\slashed{k}_2)|3\rangle\,{\rm .}
\label{MIV}
\eea
\end{changemargin}
Combining eqs. (\ref{MI}) - (\ref{MIV}) with the appropriate multiplicities, we finally find
\cmb{-.8 in}{0 in}
\bea
&&A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}} = A_{1;\,\mathcal{M}_{I}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}}
\el+ A_{1;\,\mathcal{M}_{II}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}}
+ 3 A_{1;\,\mathcal{M}_{III}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}}
\el+ 3 A_{1;\,\mathcal{M}_{IV}}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)\Big|_{I_5^{(1)}} \,{\rm .}
\label{unsimpfermans}
\eea
\cme
As before, we should project the pentagon integrals onto the dual conformal basis using (\ref{PtoDCB}) before attempting to simplify the result. After trying all BCFW shifts, we were able to find a simple formula for the above pentagon coefficient. Suppressing the overall factor of $\e$ from (\ref{PtoDCB}), we find:
\cmb{-.8 in}{0 in}
\bea
&&K_1^{f \bar{f}} = {i \over 2}C_1 {\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6}\times
\el \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab3.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.3\spb4.5\spb5.6 \Big)\,{\rm ,}
\label{fermcoef(1)}
\eea
\cme
where the $C_i$ appear in the reduction of a massless scalar hexagon integral to a sum of one mass pentagons and they entered into our six gluon results in Subsection \ref{gresults}.
The $I_5^{(1);~D = 6 - 2\e}$ pentagon coefficient of $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)\Big|_{I_5^{(1)}}$ can be derived in a completely analogous fashion. We will not describe the calculation in detail because it is extremely similar to that above but the final result is, of course, important and we present it using the similar notation:
\cmb{-.8 in}{0 in}
\bea
&&H_1^{\bar{f} f} = -{i \over 2}C_1 {\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6}\times
\el \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb3.5\spb5.6 \Big)\,{\rm .}
\label{antifermcoef(1)}
\eea
\cme
In $H_1^{\bar{f} f}$ above, the origin of the overall minus sign is a reflection of the fact that our $D$ dimensional generalized unitarity technique naturally computes $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{4},k_4^{123},k_5,k_6\right)\Big|_{I_5^{(1)}}$, which is off by a minus sign from $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)\Big|_{I_5^{(1)}} \equiv H_1^{\bar{f} f}$.
Before leaving this subsection, let us say a few more words about how we derived eqs. (\ref{fermcoef(1)}) and (\ref{antifermcoef(1)}). We treat (\ref{fermcoef(1)}) but (\ref{antifermcoef(1)}) is no more difficult. In fact, (\ref{unsimpfermans}) is particularly easy to simplify down to (\ref{fermcoef(1)}) due to its similarity to $A_{1}^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)\Big|_{I_5^{(1)}}$. Comparing eq. (\ref{fermcoef(1)}) above to
\be
K_1 = {i \over 2} C_1 {\Big(\spab2.{1+6}.5 \spab1.{2+3}.4 \spab3.{1+2}.6 +\spab5.{1+6}.2 \spa1.2 \spa2.3 \spb4.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2}
\ee
from Subsection \ref{gresults}, we see that one BCFW shift is particularly helpful in determining the analytic structure of $K_1^{f \bar{f}}$. Suppose we make the shift
\bea
\lambda_3 &&\rightarrow \lambda_3(z) = \lambda_3 + z \lambda_4\nn
\tilde{\lambda}_4 &&\rightarrow \tilde{\lambda}_4(z) = \tilde{\lambda}_4 - z \tilde{\lambda}_3 \,{\rm ,}
\eea
on the unsimplified formula for $K_1^{f\bar{f}}$ given by eq. (\ref{unsimpfermans}). What we will find is that this shift evaluated at a random phase-space point looks like
$$K_1^{f\bar{f}} + K_1 z$$
evaluated numerically at the random point. This immediately tells us to just take one of the factors of
$$\Big(\spab2.{(1+6)}.5 \spab1.{(2+3)}.4 \spab3.{(1+2)}.6 +\spab5.{(1+6)}.2 \spa1.2 \spa2.3 \spb4.5 \spb5.6 \Big)$$
in $K_1$ and replace $\lambda_3$ with $\lambda_4$ in that factor to get $K_1^{f \bar{f}}$. Of course, there is no guarantee that something like this will work in general, but we will see that we are also able to guess a simple result for the first pentagon coefficient of the scalar/anti-scalar amplitude considered in the next subsection.
\subsection{NMHV Amplitudes With a Scalar/Anti-Scalar Pair}
\label{ssbarcalc}
In this subsection, we use the results that we have so far for $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$, $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)$, $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2,k_3^{1234},k_4,k_5^{1234},k_6\right)$,\\ $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)$, and $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)$ and eq. (\ref{A63s}) for $\a_{6;3}$ to deduce an expression for the $I_5^{(1);~D = 6 - 2\e}$ pentagon coefficient of\\ $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{12},k_4^{34},k_5,k_6\right)$. Clearly, the first step is to expand eq. (\ref{A63s}) and extract the coefficient of $\eta_1^{1234}\eta_3^{1234}\eta_5^{1234}$. Doing this results in the relation
\bea
&&A_1^{1-{\rm loop}}\left(k_1^{1234}, k_2, k_3^{1234}, k_4, k_5^{1234}, k_6\right) = {\spab1.{3+5}.6^4 \over \spa1.2^4 \spb5.6^4} A_1^{1-{\rm loop}}\left(k_1^{1234}, k_2^{1234}, k_3^{1234}, k_4, k_5, k_6\right)
\el+4{\spab1.{3+5}.6^3 \spa3.1 \spb6.4\over \spa1.2^4 \spb5.6^4} A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^4,k_5,k_6\right)
\el+ 6{\spab1.{3+5}.6^2 \spa3.1^2 \spb6.4^2\over \spa1.2^4 \spb5.6^4} A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{12},k_4^{34},k_5,k_6\right)
\el + 4{\spab1.{3+5}.6 \spa3.1^3 \spb6.4^3 \over \spa1.2^4 \spb5.6^4} A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)
\el +{\spa3.1^4 \spb6.4^4\over \spa1.2^4 \spb5.6^4}A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)
\label{scaleqn}
\eea
which can be used to trivially solve for the $I_5^{(1);~D = 6 - 2\e}$ pentagon coefficient of \\$A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{12},k_4^{34},k_5,k_6\right)$.
We can simplify the complicated looking expression that results by using BCFW shifts. By examining the results that we have so far
\cmb{-1.2 in}{0 in}
\bea
K_1 &=& {i \over 2} C_1 {\Big(\spab2.{1+6}.5 \spab1.{2+3}.4 \spab3.{1+2}.6 +\spab5.{1+6}.2 \spa1.2 \spa2.3 \spb4.5 \spb5.6 \Big)^2\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6} \\
K_1^{f \bar{f}} &=& {i \over 2}C_1 {\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6}\times \nonumber \\
&& \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab3.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.3\spb4.5\spb5.6 \Big)\\
H_1^{\bar{f} f} &=& -{i \over 2}C_1 {\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6}\times \nonumber \\
&& \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb3.5\spb5.6 \Big)\\
H_1 &=& {i \over 2} C_1 {\Big(\spab2.{6+1}.5 \spab1.{2+4}.3 \spab4.{1+2}.6 +\spab5.{6+1}.2 \spa1.2 \spa2.4 \spb3.5 \spb5.6 \Big)^2\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6}
\eea
\cme
we see that a natural guess for the $I_5^{(1);~D = 6 - 2\e}$ pentagon coefficient of \\$A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{12},k_4^{34},k_5,k_6\right)$ is
\be
M_1^{s s^*} = {i \over 2}C_1 {\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)^2\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6}
\label{scalguess}
\ee
because it has a factor in common with both $K_1^{f \bar{f}}$ and $H_1^{\bar{f} f}$. In fact, this is {\it almost} right. If we try subtracting eq. (\ref{scalguess}) from the expression obtained by solving eq. (\ref{scaleqn}), we find that what's left over is easily determined using BCFW analysis to be
$$ {i\over 6}C_1 \spa1.2^2 \spb5.6^2$$
which implies that the entire pentagon coefficient is given by
\be
M_1^{s s^*} = {i \over 2}C_1 \left({\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)^2\over \spab2.{1+6}.5 \spab5.{1+6}.2 s_3 s_6} + {1\over 3}\spa1.2^2 \spb5.6^2\right)
\ee
All of the necessary pieces are now in place and we can use them to determine all of the higher-order terms in the $\Nsym$ superamplitude $\a_{6;3}$. We do this in the next subsection.
\subsection{The $\Nsym$ Supersymmetrization of the Six-Point NMHV Amplitudes}
\label{supersymresults}
In this subsection we use the results derived in the last two subsections and some of those derived in Section \ref{gluoncomp} to write down the full form of the higher-order in $\e$ contributions (in the dual conformal basis) to the one-loop planar six-point NMHV superamplitude. We present all of the higher-order pieces of the component scattering amplitudes that appear in eq. (\ref{A63s}) for $\a_{6;3}$. To determine the other pentagon coefficients we first solved the requisite leading singularity equations numerically using our $I_5^{(1);~D = 6 - 2\e}$ pentagon coefficients as inputs. Then, based on the analytical formulas obtained in this section and the last, we found it straightforward to guess appropriate compact formulas for the remaining undetermined coefficients, checking everything numerically against our numbers from the leading singularity equations. If we first define
\bea
K_1 &=& {i \over 2} C_1 {\Big(\spab2.{1+6}.5 \spab1.{2+3}.4 \spab3.{1+2}.6 +\spab5.{1+6}.2 \spa1.2 \spa2.3 \spb4.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
K_2 &=& {i \over 2} C_2 {\spab3.{1+2}.6^2 \spa1.2^2 \spb4.5^2 t_1^2 \over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \\
K_3 &=& {i \over 2} C_3 {\spab1.{2+3}.4^2 \spa2.3^2 \spb5.6^2 t_1^2 \over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \\
K_4 &=& {i \over 2} C_4 {\Big(\spab2.{1+6}.5 \spab1.{2+3}.4 \spab3.{1+2}.6 +\spab5.{1+6}.2 \spa1.2 \spa2.3 \spb4.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
K_5 &=& {i \over 2} C_5 {\spab3.{1+2}.6^2 \spa1.2^2 \spb4.5^2 t_1^2 \over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \\
K_6 &=& {i \over 2} C_6 {\spab1.{2+3}.4^2 \spa2.3^2 \spb5.6^2 t_1^2 \over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \, {\rm ,}
\eea
\bea
K_1^{f \bar{f}} &=& {i \over 2}C_1 {\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab3.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.3\spb4.5\spb5.6 \Big)\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2}\times \nonumber \\
&& \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big) \\
K_2^{f \bar{f}} &=& {i \over 2} C_2 {\Big(\spab3.{1+2}.6 \spa1.2 \spb4.5 t_1\Big)\Big(\spb4.5\spa1.2\spab4.{5+6}.3\spab3.{4+5}.6\Big)\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \\
K_3^{f \bar{f}} &=& {i \over 2} C_3 {\Big(\spab1.{2+3}.4 \spa2.3 \spb5.6 t_1\Big)\Big(\spb5.6\spa2.3\spab1.{2+3}.4\spab4.{1+2}.3\Big)\over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \\
K_4^{f \bar{f}} &=& {i \over 2}C_4 {\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab3.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.3\spb4.5\spb5.6 \Big)\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2}\times \nonumber \\
&& \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big) \\
K_5^{f \bar{f}} &=& {i \over 2} C_5 {\Big(\spab3.{1+2}.6 \spa1.2 \spb4.5 t_1\Big)\Big(\spb4.5\spa1.2\spab4.{5+6}.3\spab3.{4+5}.6\Big)\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \\
K_6^{f \bar{f}} &=& {i \over 2} C_6 {\Big(\spab1.{2+3}.4 \spa2.3 \spb5.6 t_1\Big)\Big(\spb5.6\spa2.3\spab1.{2+3}.4\spab4.{1+2}.3\Big)\over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1} \, {\rm ,}
\eea
\bea
M_1^{s s^*} &=& {i \over 2}C_1 \left({\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} + {1\over 3}\spa1.2^2 \spb5.6^2\right) \nn\\
M_2^{s s^*} &=& {i \over 2} C_2 \left({\spb4.5^2 \spa1.2^2\spab4.{5+6}.3^2\spab3.{4+5}.6^2 \over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3}+ {1\over 3}\spa1.2^2 \spb5.6^2\right) \\
M_3^{s s^*} &=& {i \over 2} C_3 \left({\spb5.6^2 \spa2.3^2\spab1.{2+3}.4^2\spab4.{1+2}.3^2 \over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1}+ {1\over 3}\spa1.2^2 \spb5.6^2\right) \\
M_4^{s s^*} &=& {i \over 2}C_4 \left({\Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} + {1\over 3}\spa1.2^2 \spb5.6^2\right) \nn\\
M_5^{s s^*} &=& {i \over 2} C_5 \left({\spb4.5^2 \spa1.2^2\spab4.{5+6}.3^2\spab3.{4+5}.6^2 \over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3}+ {1\over 3}\spa1.2^2 \spb5.6^2\right) \\
M_6^{s s^*} &=& {i \over 2} C_6 \left({\spb5.6^2 \spa2.3^2\spab1.{2+3}.4^2\spab4.{1+2}.3^2 \over s_2 s_5 \spab1.{2+3}.4 \spab4.{2+3}.1}+ {1\over 3}\spa1.2^2 \spb5.6^2\right) \, {\rm ,}
\eea
\bea
H_1^{\bar{f} f} &=& -{i \over 2}C_1 {\Big(\spab2.{6+1}.5 \spab1.{2+4}.3 \spab4.{1+2}.6 +\spab5.{6+1}.2 \spa1.2 \spa2.4 \spb3.5 \spb5.6 \Big)\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2}\times \nonumber \\
&& \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)\\
H_2^{\bar{f} f} &=& {i \over 2}C_2 {\spa1.2 \Big(\spab3.{1+2}.6 \spab4.{1+2}.3 \spb5.3 +\spab6.{1+2}.3 \spab4.{1+2}.6 \spb5.6\Big)\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3}\times \nonumber \\
&& \times \Big(\spb4.5 \spa1.2\spab4.{5+6}.3\spab3.{4+5}.6\Big) \\
H_3^{\bar{f} f} &=& {i \over 2}C_3 {\spb5.6 \Big(\spab1.{2+3}.4 \spab4.{1+2}.3 \spa2.4 +\spab4.{2+3}.1 \spab1.{2+4}.3 \spa2.1\Big)\over s_2 s_5 \spab4.{2+3}.1 \spab1.{2+3}.4}\times \nonumber \\
&& \times \Big(\spb5.6 \spa2.3\spab1.{2+3}.4\spab4.{1+2}.3\Big) \\
H_4^{\bar{f} f} &=& -{i \over 2}C_4 {\Big(\spab2.{6+1}.5 \spab1.{2+4}.3 \spab4.{1+2}.6 +\spab5.{6+1}.2 \spa1.2 \spa2.4 \spb3.5 \spb5.6 \Big)\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2}\times \nonumber \\
&& \times \Big(\spab2.{1+6}.5\spab1.{2+3}.4\spab4.{1+2}.6+\spab5.{1+6}.2\spa1.2\spa2.4\spb4.5\spb5.6\Big)\\
H_5^{\bar{f} f} &=& {i \over 2}C_5 {\spa1.2 \Big(\spab3.{1+2}.6 \spab4.{1+2}.3 \spb5.3 +\spab6.{1+2}.3 \spab4.{1+2}.6 \spb5.6\Big)\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3}\times \nonumber \\
&& \times \Big(\spb4.5 \spa1.2\spab4.{5+6}.3\spab3.{4+5}.6\Big) \\
H_6^{\bar{f} f} &=& {i \over 2}C_6 {\spb5.6 \Big(\spab1.{2+3}.4 \spab4.{1+2}.3 \spa2.4 +\spab4.{2+3}.1 \spab1.{2+4}.3 \spa2.1\Big)\over s_2 s_5 \spab4.{2+3}.1 \spab1.{2+3}.4}\times \nonumber \\
&& \times \Big(\spb5.6 \spa2.3\spab1.{2+3}.4\spab4.{1+2}.3\Big) \,{\rm ,}
\eea
and
\bea
H_1 &=& {i \over 2} C_1 {\Big(\spab2.{6+1}.5 \spab1.{2+4}.3 \spab4.{1+2}.6 +\spab5.{6+1}.2 \spa1.2 \spa2.4 \spb3.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
H_2 &=& {i \over 2} C_2 {\spa1.2^2 \Big(\spab3.{1+2}.6 \spab4.{1+2}.3 \spb5.3 +\spab6.{1+2}.3 \spab4.{1+2}.6 \spb5.6\Big)^2\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \nonumber \\ \\
H_3 &=& {i \over 2} C_3 {\spb5.6^2 \Big(\spab1.{2+3}.4 \spab4.{1+2}.3 \spa2.4 +\spab4.{2+3}.1 \spab1.{2+4}.3 \spa2.1\Big)^2\over s_2 s_5 \spab4.{2+3}.1 \spab1.{2+3}.4} \nonumber \\ \\
H_4 &=& {i \over 2} C_4 {\Big(\spab2.{6+1}.5 \spab1.{2+4}.3 \spab4.{1+2}.6 +\spab5.{6+1}.2 \spa1.2 \spa2.4 \spb3.5 \spb5.6 \Big)^2\over s_6 s_3 \spab2.{6+1}.5 \spab5.{6+1}.2} \nonumber \\ \\
H_5 &=& {i \over 2} C_5 {\spa1.2^2 \Big(\spab3.{1+2}.6 \spab4.{1+2}.3 \spb5.3 +\spab6.{1+2}.3 \spab4.{1+2}.6 \spb5.6\Big)^2\over s_1 s_4 \spab3.{1+2}.6 \spab6.{1+2}.3} \nonumber \\ \\
H_6 &=& {i \over 2} C_6 {\spb5.6^2 \Big(\spab1.{2+3}.4 \spab4.{1+2}.3 \spa2.4 +\spab4.{2+3}.1 \spab1.{2+4}.3 \spa2.1\Big)^2\over s_2 s_5 \spab4.{2+3}.1 \spab1.{2+3}.4} \nonumber \\
\eea
in what will hopefully be obvious notation given what has been discussed so far, then the final form of our answer reads
\bea
\a_{6;3}^{1-{\rm loop}} =&& \cdots + \e \,{\D^{(8)}(Q^{a\,\A}) \over \spb5.6^4 \spa1.2^4}\sum_{\ell = 1}^6 \Big(K_\ell \prod_{a=1}^4 \left(\spb5.6~\eta_3^a+\spb6.3 ~\eta_5^a +\spb{3}.5~ \eta_6^a\right)
\el+ K_\ell^{f \bar{f}} \prod_{a=1}^3 \left(\spb5.6~\eta_3^a+\spb6.3 ~\eta_5^a +\spb{3}.5~ \eta_6^a\right)\left(\spb5.6~\eta_4^4+\spb6.4 ~\eta_5^4 +\spb{4}.5~ \eta_6^4\right)
\el+ M_\ell^{s s^*}\prod_{a=1}^2 \left(\spb5.6~\eta_3^a+\spb6.3 ~\eta_5^a +\spb{3}.5~ \eta_6^a\right)\prod_{a=3}^4\left(\spb5.6~\eta_4^a+\spb6.4 ~\eta_5^a +\spb{4}.5~ \eta_6^a\right)
\el + H_\ell^{\bar{f} f} \left(\spb5.6~\eta_3^1+\spb6.3 ~\eta_5^1 +\spb{3}.5~ \eta_6^1\right)\prod_{a=2}^4\left(\spb5.6~\eta_4^a+\spb6.4 ~\eta_5^a +\spb{4}.5~ \eta_6^a\right)
\el + H_\ell \prod_{a=1}^4\left(\spb5.6~\eta_4^a+\spb6.4 ~\eta_5^a +\spb{4}.5~ \eta_6^a\right)\Big)\,I_5^{(\ell);~D = 6 - 2 \e}
\label{A63fin}
\eea
where, as usual, we have suppressed the well-known box integral contributions. In Section \ref{WL/MHV} we will see that there is another form for the higher-order in $\e$ contributions to $\a^{1-{\rm loop}}_{6;3}$ that is significantly simpler than eq. (\ref{A63fin}).
\subsection{The Structure of $\a_{n;2}$ At One, Two, and Higher Loops}
\label{BDSsect}
In this subsection we review the ongoing program of research dedicated to understanding the multi-loop structure of the planar MHV superamplitude in $\Nsym$. This program was begun with the seminal paper of Bern, Dixon, Dunbar, and Kosower (BDDK)~\cite{BDDKMHV} which computed all one-loop MHV superamplitudes in $\Nsym$ (as usual, we will only be interested in the planar contributions). Recall the notation used in eq. (\ref{MHVsupL}):
\be
\a_{n;2} = i{{1\over 16} \prod_{a = 1}^4 \sum_{i,j = 1}^n \spa{i}.j \eta^a_i \eta^a_j \over \spa1.2 \spa2.3 \cdots \spa{n}.1}\Bigg(1+\left({g^2 \Nc \mu^{2\e} e^{-\gamma_E \e} \over (4 \pi)^{2-\e}}\right) M_{1-{\rm loop}}+ \left({g^2 \Nc \mu^{2\e} e^{-\gamma_E \e}\over (4 \pi)^{2-\e}}\right)^2 M_{2-{\rm loop}}+\cdots\Bigg) \,{\rm .}
\ee
After $\a_{n;2}^{tree}$ is factored out, the analytic structure at $L$ loops, $M_{L-{\rm loop}}$, can be determined by comparing to, say, $A^{L-{\rm loop}}_1(p_1^{1234},p_2^{1234},p_3,\cdots,p_n)$ modulo the Parke-Taylor amplitude. Although most of the multi-loop $\Nsym$ literature prior to the development of $\Nsym$ on-shell superspace focused on purely gluonic amplitudes, the discussion of \ref{gendisonshell} makes it clear that, in the MHV sector, one can make this choice and still determine the full superamplitude (Of course it is probably more natural to perform all calculations in a way that preserves as many of the supersymmetries as possible~\cite{DHKSgenunit}). In all of the applications that follow, it will be useful to redefine the contribution from the $L$-th loop as follows:
\be
\left({g^2 \Nc \mu^{2\e} e^{-\gamma_E \e} \over (4 \pi)^{2-\e}}\right)^L M_{L-{\rm loop}} = \left({2 g^2 \Nc e^{-\gamma_E \e} \over (4 \pi)^{2-\e}}\right)^L\M^{(L)}(n,t_i^{[r]},\e) = a^L\M^{(L)}(n,t_i^{[r]},\e) \, {\rm ,}
\ee
where we have made the useful definitions
\be
a \equiv {\lambda e^{-\gamma_E \e} \over (4 \pi)^{2-\e}}~~~~{\rm and}~~~~t_i^{[r]} \equiv (p_i + \cdots + p_{i+r-1})^2 \,{\rm .}
\ee
Using this notation, eq. (\ref{MHVsupL}) reads
\be
\a_{n;2} = i{{1\over 16} \prod_{a = 1}^4 \sum_{i,j = 1}^n \spa{i}.j \eta^a_i \eta^a_j \over \spa1.2 \spa2.3 \cdots \spa{n}.1}\Bigg(1+\sum_{L = 1}^\infty a^L\M^{(L)}(n,t_i^{[r]},\e)\Bigg) \,{\rm .}
\ee
Let us now describe the results of BDDK in~\cite{BDDKMHV} where the structure of $\M^{(1)}(n,t_i^{[r]},\e)$ for all $n$ was determined through $\Ord(\e^0)$. It was found that:
\cmb{0. in}{0 in}
\bea
&&\M^{(1)}(n,t_i^{[r]},\e) =
\el C_\Gamma \sum_{i=1}^{n} \left( -{ 1 \over \e^2 } \bigg(
{ \mu^2 \over -t_i^{[2]} } \bigg)^{\e}
-\sum_{r=2}^{\left[{n\over 2}\right] -1}
\sum_{i=1}^n
\ln \bigg({ -t_i^{[r]}\over -t_i^{[r+1]} }\bigg)
\ln \bigg({ -t_{i+1}^{[r]}\over -t_i^{[r+1]} }\bigg) +
D_n\left(t_i^{[r]}\right) + L_n\left(t_i^{[r]}\right) +{ n \pi^2 \over 6 }\right)
\el+ \Ord(\e)\, {\rm ,}
\label{UnivFunc}
\eea
\cme
where $C_{\G}$ is given by
$$C_\G = {\G(1+\e)\G(1-\e)^2 \over 2 \G(1-2\e)}\,{\rm .}$$
The form of $D_n\left(t_i^{[r]}\right)$ and $L_n\left(t_i^{[r]}\right)$ depends upon whether $n$ is odd or even.
For $n=2m+1$,
$$
D_{2m+1}= -\sum_{r=2}^{m-1} \Bigg( \sum_{i=1}^{n}
\li2 \bigg[ 1- { t_{i}^{[r]} t_{i-1}^{[r+2]}
\over t_{i}^{[r+1]} t_{i-1}^{[r+1]} } \biggr] \Bigg)\, {\rm ,}
$$
$$
L_{2m+1}= -{ 1\over 2} \sum_{i=1}^n
\ln \bigg({ -t_{i}^{[m]}\over -t_{i+m+1}^{[m]} } \bigg)
\ln \bigg({ -t_{i+1}^{[m]}\over -t_{i+m}^{[m]} } \bigg)\, {\rm ,}
$$
whereas for $n=2m$,
$$
D_{2m}= -\sum_{r=2}^{m-2} \Bigg( \sum_{i=1}^{n}
\li2 \bigg[ 1- { t_{i}^{[r]} t_{i-1}^{[r+2]}
\over t_{i}^{[r+1]} t_{i-1}^{[r+1]} } \bigg] \Bigg)
-\sum_{i=1}^{n/2} \li2 \bigg[ 1- { t_{i}^{[m-1]}t_{i-1}^{[m+1]}
\over t_{i}^{[m]}t_{i-1}^{[m]}} \bigg]\, {\rm ,}
$$
$$
L_{2m}=-{1\over 4} \sum_{i=1}^n
\ln \bigg({ -t_{i}^{[m]}\over -t_{i+m+1}^{[m]} } \bigg)
\ln \bigg({ -t_{i+1}^{[m]}\over -t_{i+m}^{[m]} } \bigg)\, {\rm .}
$$
The above only holds for $n \geq 5$. For $n = 4$ we have
\be
\M^{(1)}(4,t_i^{[r]},\e) = C_\G \bigg\{
-{2 \over \e^2} \Big[ \left( {-s\over\mu^2}\right)^{-\e}+ \left({-t\over\mu^2}\right)^{-\e} \Big]
+ \ln^2\left( {-s \over - t} \right) + \pi^2 \bigg\} \,{\rm .}
\label{4ptanal}
\ee
Subsequently, the functions $\M^{(2)}(4,t_i^{[r]},\e)$ and $\M^{(2)}(5,t_i^{[r]},\e)$ were determined through terms of $\Ord(\e^0)$ in~\cite{ABDK} and~\cite{TwoLoopFive} respectively. Remarkably, the following relationships were found:
\bea
\M^{(2)}(4,t_i^{[r]},\e)\Big|_{\Ord(\e^0)} - {1\over 2} \M^{(1)}(4,t_i^{[r]},\e)^2\Big|_{\Ord(\e^0)} &=& \alpha~ \M^{(1)}(4,t_i^{[r]},2\e)\Big|_{\Ord(\e^0)} + \beta \label{4pt2LBDS} \\
\M^{(2)}(5,t_i^{[r]},\e)\Big|_{\Ord(\e^0)} - {1\over 2} \M^{(1)}(5,t_i^{[r]},\e)^2\Big|_{\Ord(\e^0)} &=& \alpha~ \M^{(1)}(5,t_i^{[r]},2\e)\Big|_{\Ord(\e^0)} + \beta \,{\rm ,}
\label{5pt2LBDS}
\eea
where both sides of the above are only considered through $\Ord(\e^0)$. $\alpha$ and $\beta$ are transcendentality two and four numbers respectively.\footnote{For example, $\zeta(2)$ is transcendentality two and $\zeta(4)$ is transcendentality four. One also speaks of functions carrying transcendentality ({\it e.g.} $\li2[1-{-s \over -t}]$ has transcendentality two).} Generically, $L$-loop planar amplitudes in $\Nsym$ are built out of transcendentality $2 L$ numbers and functions~\cite{KotikovLipatov}. In the above, $\alpha$ and $\beta$ have the transcendentality that they do because both sides of eqs. (\ref{4pt2LBDS}) and (\ref{5pt2LBDS}) are expected to have uniform transcendentality four.
Given these striking results, Bern, Dixon, and Smirnov proposed~\cite{BDS} the following ansatz for the analytical structure of all planar MHV superamplitudes,
\be
\ln\left(1+\sum_{L = 1}^\infty a^L \M^{(L)}(n,t_i^{[r]},\e)\right) = \sum_{L = 1}^\infty a^L\left(f^{(L)} \M^{(1)}(n,t_i^{[r]},L \e) + C^{(L)}+E^{(L)}(n,\e)\right) \,{\rm ,}
\label{BDSansatz}
\ee
which they checked for $n = 4$ through three loops. In eq. (\ref{BDSansatz}) above, $f^{(L)}$ and $C^{(L)}$ are numbers of the appropriate transcendentality ($2(L-1)$ and $2 L$ respectively) and $E^{(L)}(n,\e)$ contains higher-order in $\e$ contributions that are unimportant because, usually, both sides of (\ref{BDSansatz}) are expanded to some order in $a$ and then higher-order in $\e$ terms are dropped to put all of the $n$ dependence on the right-hand side into the function $\M^{(1)}(n,t_i^{[r]},L \e)$. Actually, a structure like this is expected in gauge theory on general grounds for the $\e$ pole terms; the IR divergences of planar non-Abelian gauge theory amplitudes are well-understood and known to exponentiate~\cite{Catani,StermanYeomans}. What is really novel about eq. (\ref{BDSansatz}) is that it holds also for the finite terms.
In fact, the so-called BDS ansatz (eq. (\ref{BDSansatz})) is known to be valid to all loop orders if $n = 4 ~{\rm or}~5$~\cite{DHKSward}. We will explain this in Section \ref{WL/MHV} after introducing dual superconformal symmetry. For higher multiplicity, however, life is not so simple. It was proven in~\cite{AMlargen,BLSV1,BDKRSVV} that the BDS ansatz is incomplete at two loops and six points. For this case, which will be the one of primary interest to us, eq. (\ref{BDSansatz}) must be modified:
\be
\M^{(2)}(6,t_i^{[r]},\e)\Big|_{\Ord(\e^0)} - {1\over 2}\M^{(1)}(6,t_i^{[r]},\e)^2\Big|_{\Ord(\e^0)}= \alpha ~\M^{(1)}(6,t_i^{[r]},2\e)\Big|_{\Ord(\e^0)} + \beta + R_{6}^{(2)}\left(t_i^{[r]}\right)\,{\rm ,}
\label{6ptstruct}
\ee
The new term on the right-hand side is called the two-loop, six-point remainder function. It is IR finite and highly constrained. For example, in order to be consistent with the known results for four and five point scattering at two loops, $R_6^{(2)}$ must have vanishing soft and collinear limits in all channels. Also, we know from the discussion above that $R_6^{(2)}$ should be a function of uniform transcendentality four. Furthermore, as we will see in the next section, the remainder function is not an arbitrary function of the $t_i^{[r]}$. In fact, for generic kinematics it is a function of only three independent variables.
\section{Dual Superconformal Symmetry and the Ratio of the Six-Point NMHV and MHV Superamplitudes at Two-Loops}
\label{WL/MHV}
In this section we review developments related to a recently discovered hidden symmetry of the planar $\Nsym$ S-matrix, {\it dual superconformal symmetry}, and we present alternative formulae for the higher-order contributions to $\a_{6;3}^{1-{\rm loop}}$ that manifest the new symmetry as much as possible. The final formula obtained is very simple and is the form of our results used in a recent study of the $\Nsym$ planar NMHV superamplitude at two loops~\cite{KRV}. As will be explained more below, one of the ideas tested in~\cite{KRV} is whether the NMHV superamplitude divided by the MHV superamplitude is dual superconformally invariant, as was proposed earlier in~\cite{DHKSdualconf} by Drummond, Henn, Korchemsky, and Sokatchev. The new results described in this section for $\a_{6;3}^{1-{\rm loop}}$ were obtained in collaboration with one of the authors of~\cite{KRV}, Cristian Vergu. This work has been primarily been about perturbation theory at weak coupling and, therefore, we will describe all developments in perturbation theory even though most of them were first seen non-perturbatively in the strong coupling regime of $\Nsym$ (via the AdS/CFT correspondence). Of course, it would be a shame to completely ignore the strong coupling regime, so we offer a brief account in Appendix \ref{ADS/CFT} for the reader interested in a historical introduction to the ideas discussed in this section.
\subsection{Light-Like Wilson-Loop/MHV Amplitude Correspondence}
\label{WL/MHV2}
Inspired by the AdS/CFT correspondence, a very surprising connection was suggested~\cite{origAldayMald} between two {\it a priori} completely unrelated observables. One of the observables,
\be
{\a_{n;2} \over \a^{tree}_{n;2}} = 1 + \sum_{L = 1}^\infty a^L \M^{(L)}(n,t_i^{[r]},\e)
\label{analstruc}
\ee
was discussed at length in Subsections \ref{gendisonshell} and \ref{BDSsect}. The other, the expectation value of an $n$-gon (denoted $C_n$) light-like Wilson loop
\be
W[C_n] \equiv {1\over \Nc} \langle0|{\rm Tr}\bigg[P\bigg\{ {\rm exp}\left(i g \oint_{C_n} dx^\nu A^a_\nu(x) t^a\right)\bigg\}\bigg]|0\rangle
\label{WLdef}
\ee
has not been introduced so far, so we will analyze its definition in some detail. Of course, we will also have to understand, at least in principle, how to calculate the set of objects introduced above perturbatively if our goal is to establish a connection between eqs. (\ref{analstruc}) and (\ref{WLdef}).
Since it may well be the case that the reader is less familiar with Wilson loop expectation values than with scattering amplitudes, we first take a step back and discuss Wilson loops in general. Wilson loops were introduced by Wilson in~\cite{origWilson} in an attempt to better understand the phenomenon of quark confinement in non-Abelian Yang-Mills theory (he had the gauge group $SU(3)_{\rm color}$ in mind). The asymptotic behavior of Wilson loop expectation values as the circumference of the loop goes to infinity tells you whether your gauge theory (in Euclidean space) is confining in the infrared. If we define $A(C)$ to be the area of the surface of minimal area bounded by $C$ and $L(C)$ to be the circumference of $C$, it can be shown~\cite{Polyakovtext} that if we have
\be
W[C] ~~~~\stackrel{L(C)\rightarrow \infty}{\longrightarrow}~~~~ W_0 e^{-k A(C)} \,{\rm ,}
\ee
then the gauge theory is in the confining phase, provided that the contour $C$ is smooth (without cusps) and is space-like back in Minkowski space (see Figure \ref{splike}).
\FIGURE{
\resizebox{.75\textwidth}{!}{\includegraphics{splikeWL.eps}}
\caption{A smooth space-like Wilson loop. The arrows indicate the direction of traversal; if the loop was unoriented it wouldn't be clear how to make sense of the path-ordering in (\ref{WLdef}).}
\label{splike}}
Although the above application is probably what would come to the minds of most researchers if asked about Wilson loops, we have something completely different in mind. The Wilson loop appearing in eq. (\ref{WLdef}) is of a special type: the contour $C_n$ defining it has cusps connected by light-like segments. The expectation value of unions of light-like Wilson lines enter into the calculation of certain universal soft functions in QCD. These soft functions are important because they control the resummation of large logarithms that often appear at the edges of phase-space when one tries to na\"{i}vely compute next-to-leading (or higher) corrections to cross-sections for processes in QCD. As we shall see, $n$-cusp light-like Wilson loop expectation values also play an important in $\Nsym$, but in a rather different way.
It is now time to return to eq. (\ref{WLdef}),
\be
W[C_n] = {1\over \Nc} \langle0|{\rm Tr}\bigg[P\bigg\{ {\rm exp}\left(i g \oint_{C_n} dx^\nu A^a_\nu(x) t^a\right)\bigg\}\bigg]|0\rangle \,{\rm ,}
\ee
and scrutinize everything that enters into the expression on the right-hand side. In the above, the gauge connection, $A_\nu^a$ is contracted with the $SU(\Nc)$ fundamental representation gauge group generators, $t^a$. The quantity $A_\nu^a t^a$ is integrated around the closed contour $C_n$ depicted in Figure \ref{hexWL} (for $n = 6$). Each cusp of $C_n$ is labeled $x_i^\nu$ and the lines between adjacent cusps have lengths $(x_i - x_{i+1})^2 = 0$. The distances between non-adjacent cusps are in general non-zero. If we introduce the notation $x_{i j}^2 = (x_i - x_j)^2$, we have nine distinct non-zero distances for $n = 6$: $\{x_{1 3}^2,\,x_{2 4}^2,\,x_{35}^2,\,x_{46}^2,\,x_{51}^2,\,x_{62}^2,\,x_{14}^2,\,x_{25}^2,\,x_{36}^2\}$. In field theory, when one is faced with evaluating the expectation value of an exponential of field operators, one simply expands the exponential and applies Wick's theorem in the standard way, usually using Feynman diagrams as a book-keeping device. This case is no different, but the appearance of the path-ordering operator, $P$, tells us to order the integrals that we get out of the exponential's Taylor expansion according to how we are traversing the Wilson loop. In fact the only non-commutative structure in the problem are the $t^a$ generator matrices, so the path-ordering in this case is just an ordering on the $SU(\Nc)$ generators that appear in the argument of the exponential. Finally, we have to trace over gauge theory indices to obtain a gauge invariant functional of $C_n$. Suppose we tried to make sense of $W[C_n]$ without the trace:
\be
W^\prime[C_n] = {1\over \Nc} \langle0|P\bigg\{ {\rm exp}\left(i g \oint_{C_n} dx^\nu A^a_\nu(x) t^a\right)\bigg\}|0\rangle \,{\rm .}
\ee
Under a gauge transformation $\Omega$, $A_\nu^a t^a$ becomes $\Omega^{-1} A_\nu^a t^a \Omega+ {i\over g} \Omega^{-1} \partial_\nu \Omega$. This induces a change in $W^\prime[C_n]$,
\be
\Omega^{-1} {1\over \Nc} \langle0|P\bigg\{ {\rm exp}\left(i g \oint_{C_n} dx^\nu A^a_\nu(x) t^a\right)\bigg\}|0\rangle \Omega \,{\rm ,}
\ee
and we see that $W^\prime[C_n]$ is not gauge invariant. This problem is easily fixed by taking the trace over generator matrices and this brings us back to $W[C_n]$.
\FIGURE{
\resizebox{.75\textwidth}{!}{\includegraphics{Llike6bet.eps}}
\caption{The Feynman diagram for one contribution to $W[C_6]$.}
\label{hexWL}}
Now that we understand how to interpret $W[C_n]$, following~\cite{DKS4pt}, we calculate it to order $\Ord(g^2)$ (lowest non-trivial order). Expanding the path-ordered exponential
gives
\be
P\bigg\{ {\rm exp}\left(i g \oint_{C_n} dx^\nu A^a_\nu(x) t^a\right)\bigg\} = 1 + i g \oint_{C_n} dx^\nu A^a_\nu(x) t^a +{1\over 2!}(i g)^2 \oint_{C_n}\oint_{C_n} dx^\rho dy^\sigma A^a_\rho(x) A^b_\sigma(y) t^a_{i j}t^b_{j k} + \cdots ~{\rm .}
\ee
Truncating the above at $\Ord(g^2)$ and taking its vacuum expectation value gives
\be
1 + {1\over 2!}(i g)^2 \oint_{C_n}\oint_{C_n} dx^\rho dy^\sigma \langle0|A^a_\rho(x) A^b_\sigma(y)|0\rangle t^a_{i j}t^b_{j k}
\ee
since $\langle0|A^a_\nu(x) t^a|0\rangle = 0$ by virtue of Lorentz invariance. Finally, we take the trace over generator matrices, tack on the overall factor of $1/\Nc$, and obtain $W[C_n]$ through $\Ord(g^2)$:
\be
W[C_n]\Big|_{\Ord(g^2)} = 1 - {g^2\over 2! \Nc} \oint_{C_n}\oint_{C_n} dx^\rho dy^\sigma \langle0|A^a_\rho(x) A^b_\sigma(y)|0\rangle t^a_{i j}t^b_{j i} \,{\rm .}
\label{WLsetup}
\ee
Since $\langle0|A^a_\rho(x) A^b_\sigma(y)|0\rangle$ is just the well known two-point correlation function for the Yang-Mills field in position space,
\be
\langle0|A^a_\rho(x) A^b_\sigma(y)|0\rangle = {-g_{\rho \sigma} \D^{a b} \mu^{2\e} \pi^\e e^{\gamma_E \e}\over 4 \pi^2(-(x-y)^2)^{1-\e}}
\ee
it is clear that Wilson loop expectation values are conveniently described by Feynman diagrams. For example, if we parametrize our $n$-gon loop, for $1 \leq i \leq n$, as
\be
\{x^\nu(\tau_i) = x_i^\nu - \tau_i x_{i\,i+1}^\nu,\, y^\nu(\tau_i) = x_i^\nu - \tau_i x_{i\,i+1}^\nu | 0\leq \tau \leq 1\} \,{\rm ,}
\ee
the $\Ord(g^2)$ contribution to $W[C_6]$ shown in Figure \ref{hexWL} can be calculated by integrating over the positions on lines $x_1 - x_2$ and $x_4 - x_5$ where the gluon stretched between them can be absorbed/emitted\footnote{Due to the fact that we have two integrals over the entire closed contour, we pick up a factor of $2!$ (from interchanging the roles of $x^\rho$ and $y^\sigma$) that cancels against the factor of $2!$ in the denominator of eq. (\ref{WLsetup}).}:
\bea
&&-{g^2 \over \Nc} \int_{x_1^\rho}^{x_2^\rho} dx^\rho \int_{x_4^\sigma}^{x_5^\sigma} dy^\sigma {-g_{\rho \sigma} \D^{a b} \mu^{2\e} \pi^\e e^{\gamma_E \e}\over 4 \pi^2(-(x-y)^2)^{1-\e}} t^a_{i j}t^b_{j i}
\elale -{g^2 \over \Nc} \int_{0}^{1} (-d\tau_1 x_{12}^\rho) \int_{0}^{1} (-d\tau_4 x_{45}^\sigma) {-g_{\rho \sigma} \mu^{2\e} \pi^\e e^{\gamma_E \e} \over 4 \pi^2(-(x_1-x_4-\tau_1 x_{12} + \tau_4 x_{45})^2)^{1-\e}} t^a_{i j}t^a_{j i}
\elale {\mu^{2\e}g^2 \pi^\e e^{\gamma_E \e}\over 4 \pi^2 \Nc}\int_{0}^{1} d\tau_1 \int_{0}^{1} d\tau_4 {x_{12}\cdot x_{45} \over (-(x_1-x_4-\tau_1 x_{12} + \tau_4 x_{45})^2)^{1-\e}} C_F \Nc
\elale {\mu^{2\e}g^2 \pi^\e e^{\gamma_E \e} C_F \over 4 \pi^2}\int_{0}^{1} d\tau_1 \int_{0}^{1} d\tau_4 {x_{12}\cdot x_{45} \over (-(x_1-x_4-\tau_1 x_{12} + \tau_4 x_{45})^2)^{1-\e}}
\eea
On general grounds, we expect such a contribution to be a real number for $(x_1-x_4-\tau_1 x_{12} + \tau_4 x_{45})^2 < 0$ and $\e$ sufficiently small, real, and positive. In this paper we will never have to leave the region where these conditions are satisfied.
\FIGURE{
\resizebox{.9\textwidth}{!}{\includegraphics{WL4bet.eps}}
\caption{The complete set of Feynman diagrams required to calculate $W[C_4]$ to $\Ord(g^2)$.}
\label{WL4f}}
Of course, for $n = 6$, we will have to add a very large number of topologically distinct contributions in order to obtain a gauge invariant result. It will be simpler and get the point across just as effectively if we follow~\cite{DKS4pt} and calculate $W[C_4]$ in detail to $\Ord{(g^2)}$. The complete set of diagrams for the $\Ord{(g^2)}$ correction to $W[C_4]$ are shown in Figure \ref{WL4f}. In this case the only non-zero invariants are $x_{13}^2$ and $x_{24}^2$. Clearly, the first line of diagrams in Figure \ref{WL4f} vanish once the light-like character of the Wilson loop is taken into account. The second line of diagrams are divergent due to presence of the cusps. These divergences come from the regions of parameter-space where positions of absorption/emission approach a cusp. Such divergences are short distance and therefore ultraviolet in nature. Finally, we will see that the last line of diagrams are finite in four dimensions. If we denote the diagram in class $(\ell)$ that has a gluon stretched between lines $x_i - x_{i+1}$ and $x_j - x_{j+1}$ as $\mathcal{W}_{ij}^{(\ell)}$,
we have
\bea
\mathcal{W}_{i i}^{(1)} &=& 0 ~{\rm for~all}~1 \leq i \leq 4 \\
\mathcal{W}_{1 2}^{(2)} &=& \mathcal{W}_{3 4}^{(2)} = -{g^2 C_F e^{\gamma_E \e}(-x_{13}^2 \pi \mu^2)^\e \over 8 \pi^2 \e^2}\\
\mathcal{W}_{2 3}^{(2)} &=& \mathcal{W}_{1 4}^{(2)} = -{g^2 C_F e^{\gamma_E \e}(-x_{24}^2 \pi \mu^2)^\e \over 8 \pi^2 \e^2}\\
\mathcal{W}_{1 3}^{(3)} &=& \mathcal{W}_{2 4}^{(3)} = {g^2 C_F e^{\gamma_E \e}\left(\ln^2\left({x_{13}^2 \over x_{24}^2}\right)+\zeta(2)\right) \over 16 \pi^2} \,{\rm .}
\eea
Taking into account the fact that, in the large $\Nc$ limit,
$$C_F = {\Nc^2-1\over 2\Nc} \longrightarrow {\Nc \over 2} \, {\rm ,}$$
we make the replacement
\be
{g^2 C_F e^{\gamma_E \e} \pi^\e \over 8 \pi^2} \longrightarrow a
\ee
and find that the $\Ord(a)$ analytic structure of $W[C_4]$ is given by~\cite{DKS4pt}
\be
W[C_4]\Big|_{\Ord(a)} = -{1 \over \e^2}\bigg(\left(-x_{13}^2 \mu^2\right)^\e+\left(-x_{24}^2 \mu^2\right)^\e\bigg)+{1\over 2}\left(\ln^2\left({x_{13}^2 \over x_{24}^2}\right)+\pi^2\right) + \Ord(\e)\,{\rm .}
\label{4ptWLfin}
\ee
This is a remarkable result. Recall eq. (\ref{4ptanal}), where we wrote down the one-loop analytic structure of the four-point MHV superamplitude:
\be
\M^{(1)}(4,t_i^{[r]},\e') = C_\G \bigg\{
-{2 \over \e'^2} \Big[ \left( {-s\over\mu'^2}\right)^{-\e'}+ \left({-t\over\mu'^2}\right)^{-\e'} \Big]
+ \ln^2\left( {-s \over - t} \right) + \pi^2 \bigg\} \,{\rm ,}
\label{4ptanal2}
\ee
where $C_\G$ is given by
$$C_\G = {\G(1+\e')\G(1-\e')^2 \over 2 \G(1-2\e')}\,{\rm .}$$
Up to some redefinition of $a$, $\e$, and $\mu$, the above expression for $W[C_4]$ at lowest non-trivial order matches the above formula for $\M^{(1)}(4,t_i^{[r]},\e')$ exactly if we make the identifications
\be
s \leftrightarrow x_{13}^2 ~~~~{\rm and}~~~~ t \leftrightarrow x_{24}^2\,{\rm .}
\ee
This surprising connection captures the essence of the light-like Wilson loop/MHV amplitude correspondence in planar $\Nsym$. Even more remarkably, the work of~\cite{DKS4pt} generalizes. There is now a large body of evidence for the following relation
\be
\ln \Bigg({\a_{n;2}\over \a_{n;2}^{tree}}\Bigg)\Bigg|_{{\rm finite};\,\Ord(a^L)} = \ln \bigg(W[C_n]\bigg)\Bigg|_{{\rm finite};\,\Ord(a^L)} + D_n^{(L)}\, {\rm ,}
\label{equiv}
\ee
valid for all multiplicity and for all-loop orders (see Appendix \ref{ADS/CFT} for a bit more discussion). In the above, $D_n^{(L)}$ is transcendentality $2 L$ number. As one might guess from eqs. (\ref{4ptWLfin}) and (\ref{4ptanal2}) there is also a relation between the IR poles on the amplitude side and UV poles on the Wilson loop side. As hinted at above, one must make some non-trivial redefinitions of parameters in order to make this precise. See~\cite{DixonStermanMagnea} for a discussion of the IR poles.
The key observation is that there is a superconformal symmetry (see Appendix \ref{sconf} if unfamiliar with superconformal symmetry) acting on the Wilson loop in a natural way because it is defined in a configuration space (where the Lagrangian density that possesses this symmetry is constructed). Ultraviolet divergences in the Wilson loop due to the presence of cusps breaks the subgroup of conformal transformations in a controlled fashion. The action of the conformal symmetry is anomalous and one can derive non-perturbatively valid anomalous conformal Ward identities that fix the finite part of $W[C_n]$ that comes from the breaking of the conformal symmetry up to an additive constant at all loop orders~\cite{DHKSward}. What remains must be a function of the conformal cross-ratios. For example, at the six-point level, there are three such cross-ratios
\be
u_1 = {x_{1 3}^2 x_{4 6}^2\over x_{1 4}^2 x_{3 6}^2} \qquad u_2 = {x_{2 4}^2 x_{5 1}^2\over x_{2 5}^2 x_{1 4}^2} \qquad u_3 = {x_{3 5}^2 x_{6 2}^2\over x_{3 6}^2 x_{2 5}^2}\,{\rm ,}
\label{crsrts}
\ee
each of which is invariant under conformal transformations. Now recall that the special conformal transformations can be obtained by conjugating the spatial translations by the conformal inversion operator, $I$ (Appendix \ref{sconf}). Furthermore, it is straightforward to see that $(x_{i j})^{\alpha \dot{\alpha}} = (x_i^\mu - x_j^\mu)(\sigma_\mu)^{\alpha \dot{\alpha}}$ transforms under inversion as:
\be
I[x_{i j}] = x_i^{-1} - x_j^{-1} = -x_j^{-1} (x_i - x_j)x_i^{-1} = -x_j^{-1}x_{i j}x_i^{-1}\,{\rm .}
\ee
Due to the fact that $u_2$ and $u_3$ are obtained by cyclicly permuting $u_1$, we can rest assured that they are conformally invariant if $u_1$ is. We see that
\be
I[u_1] = {I[x_{13}^2] I[x_{46}^2] \over I[x_{1 4}^2] I[x_{3 6}^2]} = {{x_{13}^2 \over x_1^2 x_3^2}{x_{46}^2\over x_4^2 x_6^2} \over {x_{1 4}^2 \over x_1^2 x_4^2} {x_{3 6}^2 \over x_3^2 x_6^2}} = {x_{1 3}^2 x_{4 6}^2\over x_{1 4}^2 x_{3 6}^2}
\ee
and $u_1$ is indeed invariant under inversion. This actually implies the invariance of $u_1$ under the full conformal group, since it is obviously invariant under Poincar\'{e} transformations and dilatations.
We are now in a position to make some remarks about the analytic structure of the $n$-point MHV superamplitudes in $\Nsym$. As we will discuss more in the next subsection, the fact the Wilson loop/MHV amplitude correspondence of eq. (\ref{equiv}) exists implies the existence of a novel hidden symmetry of the planar $\Nsym$ MHV amplitudes through the identification $x_{i}^\mu - x_{i+1}^\mu = p_i^\mu$. This symmetry is ``hidden'' because it cannot have its origin in the Lagrangian (it acts naturally in momentum space). This hidden symmetry is called dual superconformal invariance for reasons that should now be clear. In fact, the dual conformal subgroup already tells us quite a bit of useful information about the analytic structure of the MHV superamplitude. For instance, the reason that the BDS ansatz gives the exact finite part for $n = 4~{\rm or} ~5$ external states is obvious once one understands that the ansatz is just the contribution of the conformal anomaly and that the conformal anomaly is exact for four or five points; due to the light-like nature of the Wilson loops under consideration, no conformally invariant cross-ratios can even be written down for four or five particles in $\Nsym$. In fact, if dual conformal symmetry was not broken by IR divergences, we would expect the full non-perturbative answer to be just a constant times the appropriate tree amplitude.
We can also make precise the arguments of the two-loop six-point remainder function mentioned in Subsection \ref{BDSsect}. Recall eq. (\ref{6ptstruct}) for the analytic structure of $\ln(1+\sum_{L = 1} a^L \M^{(L)}(6,t_i^{[r]},\e))$ expanded up to second order in perturbation theory:
\be
\M^{(2)}(6,t_i^{[r]},\e)\Big|_{\rm finite} - {1\over 2}\M^{(1)}(6,t_i^{[r]},\e)^2\Big|_{\rm finite}= \alpha ~\M^{(1)}(6,t_i^{[r]},2\e)\Big|_{\rm finite} + \beta + R_{6}^{(2)}\left(t_i^{[r]}\right)\,{\rm .}
\ee
Given everything that we have discussed so far, it is clear that the six-point two-loop remainder function
$R_{6}^{(2)}\left(t_i^{[r]}\right)$ must actually be a function of three dual conformally invariant cross-ratios. If we use the dictionary
\cmb{-.2 in}{0 in}
\bea
&&x_{13}^2 \leftrightarrow s_{1} \qquad x_{24}^2 \leftrightarrow s_{2} \qquad x_{35}^2 \leftrightarrow s_{3} \qquad x_{46}^2 \leftrightarrow s_{4}\nonumber \\
x_{15}^2 & \leftrightarrow & s_{5} \qquad x_{26}^2 \leftrightarrow s_{6} \qquad x_{14}^2 \leftrightarrow t_{1}\qquad
x_{25}^2 \leftrightarrow t_{2} \qquad
x_{36}^2 \leftrightarrow t_{3}
\eea
\cme
we see that, from the point of view of dual conformal symmetry, eq. (\ref{crsrts}) becomes
\be
u_1 = {s_1 s_4\over t_1 t_3} \qquad u_2 = {s_2 s_5\over t_2 t_1} \qquad u_3 = {s_3 s_6\over t_3 t_2}
\ee
and we have
\be
R_{6}^{(2)}\left(t_i^{[r]}\right) = R_{6}^{(2)}(u_1,u_2,u_3) \, {\rm .}
\ee
So far, we have really only used the dual conformal subgroup of the dual superconformal symmetry. In the next section we will describe the full dual symmetry group~\cite{DHKSdualconf}.
\subsection{Dual Superconformal Invariance and the Pentagon Coefficients of the Planar $\Nsym$ One-Loop Six-Point NMHV Superamplitude}
\label{DSI}
To realize the dual superconformal generators on their dual superspace~\cite{DHKSdualconf} we introduce variables $\theta_{i\,\alpha}^{a}$ to solve the $\delta^{(8)}$($Q^a_{~\alpha}$) supercharge conservation constraint in much the same way that the $x_{i\,\alpha\dot{\alpha}}$ of the last subsection solve the $\delta^{(4)}$($P_{\alpha \dot{\alpha}}$) momentum conservation constraint. In other words,
\be
\theta_{i\,\alpha}^{a}-\theta_{i+1\,\alpha}^{a} = \lambda_{i\,\alpha}\eta_i^a
\ee
is the supersymmetric complement of the relation
\be
x_{i\,\alpha\dot{\alpha}} - x_{i+1\,\alpha\dot{\alpha}} = \lambda_{i\,\alpha}\tilde{\lambda}_{i\,\dot{\alpha}} \,{\rm .}
\ee
Intuitively, since (dual) superconformal symmetry naturally acts in (momentum) position space and position and momentum are not mutually compatible observables, we expect the algebra of the ordinary superconformal group (see Appendix \ref{sconf}) and the dual superconformal group to be somehow entangled. This intuition is correct; the sketch below shows that there is indeed some overlap between the non-trivial generators of the superconformal (left-hand side) and the dual superconformal (right-hand side) groups:
\bea
~~~~~P_{\alpha \dot{\alpha}}~~~~~&&~~~~~\mc{K}_{\alpha \dot{\alpha}} \nonumber \\
Q^a_{~\alpha}~~~~~~~~~~\bar{Q}_{b\,\dot{\alpha}} &=& \bar{\mc{S}}_{b\,\dot{\alpha}}~~~~~~~~~~\mc{S}^a_{~\alpha} \nonumber \\
S_{a\,\alpha}~~~~~~~~~~\bar{S}^b_{~\dot{\alpha}} &=& \bar{\mc{Q}}^b_{~\dot{\alpha}}~~~~~~~~~~\mc{Q}_{a\,\alpha} \nonumber \\
~~~~~K_{\alpha \dot{\alpha}}~~~~~&&~~~~~\mathcal{P}_{\alpha \dot{\alpha}}
\label{dsalgstruc}
\eea
In the above, the generators $Q^a_{~\alpha}$ and $P_{\alpha \dot{\alpha}}$ on the superconformal side and $\mc{Q}_{a\,\alpha}$ and $\mathcal{P}_{\alpha \dot{\alpha}}$ on the dual superconformal side are actually realized in a pretty trivial fashion and were just included to make the table look more symmetrical:
\bea
Q^a_{~\alpha} &=& \sum_{i = 1}^n \lambda_{i\,\alpha}\eta^a_i\qquad {\rm and} \qquad P_{\alpha \dot{\alpha}} = \sum_{i = 1}^n \lambda_{i\,\alpha}\tilde{\lambda}_{i\,\dot{\alpha}}\nonumber \\
\mc{Q}_{a\,\alpha} &=& \sum_{i = 1}^n {\partial\over \partial \T_{i}^{a\,\alpha}}\qquad {\rm and} \qquad \mathcal{P}_{\alpha \dot{\alpha}} = \sum_{i = 1}^n {\partial\over \partial x_{i}^{~\,\alpha \dot{\alpha}}}\,{\rm .}
\eea
The generators $S_{a\,\alpha}$ and $K_{\alpha \dot{\alpha}}$ on the superconformal side and $\mc{S}^a_{~\alpha}$ and $\mc{K}_{\alpha \dot{\alpha}}$ on the dual superconformal side are a lot more complicated:
\bea
S_{a\,\alpha} &=& \sum_{i = 1}^n {\partial\over \partial \lambda_{i}^{\,~\alpha}}{\partial\over \partial \eta^a_i}\qquad {\rm and} \qquad K_{\alpha \dot{\alpha}} = \sum_{i = 1}^n {\partial\over \partial \lambda_{i}^{\,~\alpha}}{\partial\over \partial \tilde{\lambda}_{i}^{~\,\dot{\alpha}}}\nonumber \\
\mc{S}^a_{~\alpha} &=& \sum_{i = 1}^n \left(-\theta^b_{i\,\alpha}\theta^{a\,\beta}_{i}{\partial\over \partial \T_{i}^{b\,\B}}+x_{i\,\alpha}^{~~\,\dot{\beta}}\theta^{a\,\beta}_{i}{\partial\over \partial x_{i}^{~\,\B \dot{\B}}}+\lambda_{i\,\alpha}\theta^{a\,\gamma}_{i}{\partial\over \partial \lambda_{i}^{\,~\g}}+x_{i+1\,\alpha}^{\,~~~~~\dot{\beta}}\eta^a_i{\partial\over \partial \tilde{\lambda}_{i}^{~\,\dot{\B}}}-\theta^b_{i+1\,\alpha}\eta^a_i{\partial\over \partial \eta^b_i}\right) \qquad {\rm and} \nonumber \\
\mc{K}_{\alpha \dot{\alpha}} &=& \sum_{i = 1}^n \left(x_{i\,\alpha}^{~~\,\dot{\beta}}x_{i\,\dot{\alpha}}^{~~\,\beta}{\partial\over \partial x_{i}^{~\,\B \dot{\B}}}+x_{i\,\dot{\alpha}}^{~~\,\beta}\theta_{i\,\alpha}^b{\partial\over \partial \T_{i}^{b\,\B}}+x_{i\,\dot{\alpha}}^{\,~~\beta}\lambda_{i\,\alpha}{\partial\over \partial \lambda_{i}^{\,~\B}}+x_{i+1\,\alpha}^{~~~~~\,\dot{\beta}}\tilde{\lambda}_{i\,\dot{\alpha}}{\partial\over \partial \tilde{\lambda}_{i}^{~\,\dot{\B}}}+\tilde{\lambda}_{i\,\dot{\alpha}}\theta^b_{i+1\,\alpha}{\partial\over \partial \eta_{i}^b}\right)\,{\rm .}\nn
\eea
Finally, the generators $\bar{Q}_{b\,\dot{\alpha}}$ and $\bar{S}^b_{~\dot{\alpha}}$ on the superconformal side and $\bar{\mc{S}}_{b\,\dot{\alpha}}$ and $\bar{\mc{Q}}^b_{~\dot{\alpha}}$ on the dual superconformal side:
\bea
\bar{Q}_{b\,\dot{\alpha}} &=& \sum_{i = 1}^n \tilde{\lambda}_{i\,\dot{\alpha}}{\partial\over \partial \eta^b_{i}}\qquad {\rm and} \qquad \bar{S}^b_{~\dot{\alpha}} = \sum_{i = 1}^n \eta_{i}^b{\partial\over \partial \tilde{\lambda}_{i}^{~\,\dot{\alpha}}} \label{QbarSbar} \\
\bar{\mc{S}}_{b\,\dot{\alpha}} &=& \sum_{i = 1}^n \left(x_{i\,\dot{\alpha}}^{\,~~\beta}{\partial\over \partial \T_{i}^{b\,\B}}+\tilde{\lambda}_{i\,\dot{\alpha}}{\partial\over \partial \eta^b_{i}}\right)\qquad {\rm and} \qquad \bar{\mc{Q}}^b_{~\dot{\alpha}} = \sum_{i = 1}^n \left(\theta^{b\,\alpha}_{i}{\partial\over \partial x_{i}^{~\,\alpha \dot{\alpha}}}+\eta_{i}^{b}{\partial\over \partial \tilde{\lambda}_{i}^{~\,\dot{\alpha}}}\right)\,{\rm .}\nonumber
\eea
actually match up if we restrict to the on-shell superspace introduced in Section \ref{supercomp} (by ignoring all $\theta_{i\,\alpha}^{a}$ terms).
Of course, if one wants to check explicitly that all the (anti)commutation relations (see Appendix \ref{sconf}) are satisfied, one needs the rest of the representation. The rest of the ordinary and dual superconformal generators are given in Appendix \ref{sconf}. The above discussion was just intended to give the reader some sense of how the ordinary and dual superconformal algebras fit together. Although we will not use it here, it is worth emphasizing that the superconformal and dual superconformal algebras do not commute (this is clear from the form of eq. (\ref{dsalgstruc})). It turns out that their closure is a Yangian~\cite{DHPYangian1,BHMPYangian2}. It is also worth pointing out that, since the dual superconformal generators are first order differential operators, one may expect them to be better behaved at the quantum level than the usual superconformal generators. It is therefore reasonable to suppose that formulae for one-loop superamplitudes which make the dual superconformal symmetry as manifest as possible will be simpler than those of Section \ref{supercomp} (there the dual superconformal symmetry was hidden).
In~\cite{DHKSdualconf}, Drummond, Henn, Korchemsky, and Sokatchev constructed a set of six dual superconformally invariant functions,
\bea
R_{1 4 6} &=& {\D^{(4)}\left(\spb4.5 \eta_6^a+\spb5.6\eta_4^a+
\spb6.4\eta_5^a\right) \prod_{i = 1}^6 \spa{i}.{i+1}\over t_1 \spa1.2 \spa2.3 \spab1.{5+6}.4 \spab3.{4+5}.6\spb4.5\spb5.6}~~~R_{2 5 1} = {\D^{(4)}\left(\spb5.6 \eta_1^a+\spb6.1\eta_5^a+
\spb1.5\eta_6^a\right) \prod_{i = 1}^6 \spa{i}.{i+1}\over t_2 \spa2.3 \spa3.4 \spab2.{6+1}.5 \spab4.{5+6}.1\spb5.6\spb6.1} \nonumber \\
R_{3 6 2} &=& {\D^{(4)}\left(\spb6.1 \eta_2^a+\spb1.2\eta_6^a+
\spb2.6\eta_1^a\right) \prod_{i = 1}^6 \spa{i}.{i+1}\over t_3 \spa3.4 \spa4.5 \spab3.{1+2}.6 \spab5.{6+1}.2\spb6.1\spb1.2}~~~R_{4 1 3} = {\D^{(4)}\left(\spb1.2 \eta_3^a+\spb2.3\eta_1^a+
\spb3.1\eta_2^a\right) \prod_{i = 1}^6 \spa{i}.{i+1} \over t_1 \spa4.5 \spa5.6 \spab4.{2+3}.1 \spab6.{1+2}.3\spb1.2\spb2.3} \nonumber \\
R_{5 2 4} &=& {\D^{(4)}\left(\spb2.3 \eta_4^a+\spb3.4\eta_2^a+
\spb4.2\eta_3^a\right)\prod_{i = 1}^6 \spa{i}.{i+1} \over t_2 \spa5.6 \spa6.1 \spab5.{3+4}.2 \spab1.{2+3}.4\spb2.3\spb3.4}~~~R_{6 3 5} = {\D^{(4)}\left(\spb3.4 \eta_5^a+\spb4.5\eta_3^a+
\spb5.3\eta_4^a\right) \prod_{i = 1}^6 \spa{i}.{i+1}\over t_3 \spa6.1 \spa1.2 \spab6.{4+5}.3 \spab2.{3+4}.5\spb3.4\spb4.5} \nn
\eea
which they then used to write all of the one-loop box coefficients of $\a^{1-{\rm loop}}_{6;3}$ in a way that meshes well with dual superconformal symmetry. More precisely, they found that they could express all the leading singularities in the computation of the NMHV superamplitude in a manifestly dual superconformally invariant way using $R_{1 4 6}$ and its cyclic permutations. In $\Nsym$ there is a choice of basis (the dual conformal basis introduced in Subsection \ref{GUD}) where the dual superconformal properties of the theory at loop level are as manifest as possible. At the one-loop level, this basis consists of $D = 4 - 2\e$ box integrals and $D = 6 - 2\e$ pentagon integrals.
The simplicity of the results for boxes suggests that we should try to play the same game for the (now known) pentagon coefficients of the NMHV superamplitude. This is actually not as straightforward as it sounds, due to the fact that the $R_{pqr}$ above do not form a linearly independent set~\cite{DHKSdualconf,DHsuperBCFW}. In fact, for each pentagon topology, it is possible to fit an ansatz of the form
\bea
&&C_i {i \D^{(8)}(Q^{a\,\alpha})\over \spa1.2\spa2.3\spa3.4\spa4.5\spa5.6\spa6.1} \left(z_1^{(i)} R_{413} + z_2^{(i)} R_{524} + z_3^{(i)} R_{635}+ z_4^{(i)} R_{146}+z_5^{(i)} R_{251}+z_6^{(i)} R_{362}\right) \nonumber \\
&=& C_i \a^{tree}_{6;2} \left(z_1^{(i)} R_{413} + z_2^{(i)} R_{524} + z_3^{(i)} R_{635}+ z_4^{(i)} R_{146}+z_5^{(i)} R_{251}+z_6^{(i)} R_{362}\right)
\eea
using just five component amplitudes (for example those used to fix the form of $\a^{1-{\rm loop}}_{6;3}$ in Section \ref{supercomp}). Fortunately, there is an obvious, preferred, maximally symmetric solution: $z_i^{(i)} = z_{i+3}^{(i)}$. For example, for the pentagon coefficient of $I_5^{(5)}$, we set $z_5^{(5)} = z_2^{(5)}$. This choice then forces
\be
\left(z_1^{(5)}\right)^{\langle~\rangle \leftrightarrow [~]} = z_4^{(5)} \qquad \left(z_3^{(5)}\right)^{\langle~\rangle \leftrightarrow [~]} = z_6^{(5)}
\ee
as well. The other topologies behave in a completely analogous fashion. To simplify the result, it is convenient to work numerically with complex spinors. It is then possible to recognize the origin of the imaginary parts of the $z_i$ as coming from the natural odd six-point invariant
$$\spb1.2\spa2.3\spb3.4\spa4.5\spb5.6\spa6.1-\spa1.2\spb2.3\spa3.4\spb4.5\spa5.6\spb6.1\,{\rm .}$$
We can now determine the rest of the structure by experimenting with real-valued candidate expressions that respect all the constraints of the problem and have the right BCFW shifts in all channels. In the end, we find
\cmb{0 in}{0 in}
\bea
&&\a^{1-{\rm loop}}_{6;3} =
\el \cdots + \frac{i}{6} \e ~\a^{tree}_{6;2}\sum_{i = 1}^6 C_i\bigg(
{1\over 2} \left(2 s_{i+1} s_{i-2} - t_{i} t_{i+1}\right) t_{i-1} \left(R_{i+2\,i-1\,i+1} + R_{i-1\,i+2\,i-2}\right) \nonumber \\
&& - \Big([i \,i+1] \langle i+1\, i+2\rangle [i+2 \,i+3] \langle i+3\, i+4\rangle [i+4 \,i+5] \langle i+5\,i\rangle \nonumber \\
&&- \langle i\, i+1\rangle [i+1 \,i+2] \langle i+2\, i+3\rangle [i+3\, i+4] \langle i+4\, i+5\rangle [i+5\, i]\Big) \left(R_{i+2 \, i-1\,i+1} - R_{i-1\,i+2\,i-2}\right) \nonumber \\
&& + {1\over 2} \left(2 s_{i-1} s_{i+2} - t_{i-1} t_{i+1}\right) t_{i} \left(R_{i+3\,i\,i+2} + R_{i\,i+3\,i-1}\right) + {1\over 2} \left(2 s_{i+3} s_{i} - t_{i} t_{i-1}\right) t_{i+1} \left(R_{i+1\,i-2\,i} + R_{i-2\,i+1\,i-3}\right) \nonumber \\
&& -\Big([i \,i+1] \langle i+1\, i+2\rangle [i+2 \,i+3] \langle i+3\, i+4\rangle [i+4 \,i+5] \langle i+5\,i\rangle \nonumber \\
&&- \langle i\, i+1\rangle [i+1 \,i+2] \langle i+2\, i+3\rangle [i+3\, i+4] \langle i+4\, i+5\rangle [i+5\, i]\Big) \left(R_{i+1\,i-2\,i} - R_{i-2\,i+1\,i-3}\right) \bigg) I^{(i),\,D = 6 - 2\e}_5 \,{\rm .} \nn
\label{sd}
\eea
\cme
Remarkably, when written in this form, the pentagon contributions to the one-loop six-point NMHV superamplitude are related by cyclic symmetry. We can use this fact to explain relation (\ref{mystrel}), reproduced below for convenience:
\be
{K_1 \over C_1} = {K_4 \over C_4} \qquad
{K_2 \over C_2} = {K_5 \over C_5} \qquad
{K_3 \over C_3} = {K_6 \over C_6} \, {\rm .}
\ee
Examining eq. (\ref{sd}), it is trivial to see that the only piece of a given pentagon that does not return to itself under $i \rightarrow i+3$, is the $C_i$ coefficient out front. Evidently, relation (\ref{mystrel}) is a property of the full superamplitude because the symmetric choice
$$z_2^{(i)} = z_5^{(i)}$$
in our ansatz was necessary to manifest the cyclic symmetry of the superamplitude; writing the superamplitude in the form given by eq. (\ref{sd}) shows that there are not enough independent R-invariant structures, to support a full $i \rightarrow i+6$ symmetry for the coefficients divided by their $C_i$. That there are only three independent R-invariant structures can be understood as a consequence of parity invariance in the superamplitude; parity acts on R-invariants by shifting their indices from $i$ to $i+3$.
Now that we have in hand a pretty formula for the pentagon coefficients of $\a^{1-{\rm loop}}_{6;3}$ built out of dual superconformal invariants, it would be nice if there was some application of our result. It is to this that we turn in the next subsection.
\subsection{Ratio of the Six-Point NMHV and MHV Superamplitudes at Two-Loops}
\label{rationfunc}
In~\cite{DHKSdualconf}, Drummond, Henn, Korchemsky, and Sokatchev made an interesting all-loop prediction based on a remarkable one-loop calculation in their paper. They calculated the parity even part of the NMHV ratio function, $R_{{\rm NMHV}} \equiv \a_{6;3}/\a_{6;2}\,$, to $\Ord(a)$ and found a dual superconformally invariant function. This is a non-trivial result because both $\a_{6;3}$ and $\a_{6;2}$ have IR divergences. The universal, helicity-blind structure of the IR divergences guarantees that the NMHV ratio function is finite to all loop orders. However, at loop level the dual superconformal symmetry is anomalous. One way to circumvent this problem might be to write
\be
\a_{6;3} = \a_{6;2} \Big(R_{\rm NMHV} + \Ord(\e)\Big)
\label{ratiofunc}
\ee
to all loop orders and hope that all of the messiness associated with dual superconformal anomalies resides in the $\a_{6;2}$ prefactor.\footnote{Recently, Beisert, Henn, McLoughlin, and Plefka developed a technique to address these anomalies directly by deforming the dual superconformal generators~\cite{BHMPYangian2}.} It is not {\it a priori} clear that the ratio function, $R_{\rm NMHV}$, should have any special properties.. For example, as discussed in~\cite{DHKSdualconf}, it is not obvious that the dual superconformal generator $\bar{\mc{Q}}^a_{~\dot{\alpha}}$ annihilates the ratio function, because this generator (eq. (\ref{QbarSbar})) is sensitive to the dependence of $R_{\rm NMHV}$ on the dual variables, $x_{i\,\alpha \dot{\alpha}}$, and the dependence of the finite parts of $\a_{6;3}$ and $\a_{6;2}$ on the dual variables is fairly complicated (even at $\Ord(a)$). Therefore it is interesting to check by explicit calculation that $R_{\rm NMHV}$ is given by a dual superconformally invariant function. We have already seen that pulling a factor of $\a_{6;2}^{tree}$ out of $\a_{6;3}^{1-{\rm loop}}$ is a natural operation and simplifies the formula for the one-loop NMHV superamplitude. The question is whether $\a_{6;3}^{1-{\rm loop}}$ simplifies when one factors out the entire one-loop MHV superamplitude.
DHKS carried out this analysis and they found that $R_{\rm NMHV}^{1-{\rm loop}}$ could be expressed in terms of R-invariants and linear combinations of two mass hard, two mass easy, and one mass boxes (see eqs. (\ref{6ptgMHV}) and (\ref{6ptgNMHV})). When evaluated through $\Ord(\e^0)$ (see eqs. (\ref{box1})-(\ref{box2h})), these box integrals give rise to logarithms and dilogarithms. After simplifying all logarithms and dilogarithms, non-trivial cancellations occur and DHKS found the simple dual superconformally invariant result:\footnote{In eq. (\ref{1Lratfunc}) the $\pi^2/3$ factors are inessential and depend on precisely how one defines the analytic structure of the MHV amplitude. We follow the conventions of DHKS in~\cite{DHKSgenunit}.}
\bea
R_{\rm NMHV}^{1-{\rm loop}} &=& {1 \over 4} \sum_{i=1}^6 R_{i\,i+3\,i+5} \bigg(-\ln\left(u_i\right)\ln\left(u_{i+1}\right)+\ln\left(u_{i+1}\right)\ln\left(u_{i+2}\right)+\ln\left(u_{i+2}\right)\ln\left(u_i\right) \nonumber \\
&+&\li2\left(1-u_i\right)+\li2\left(1-u_{i+1}\right)+\li2\left(1-u_{i+2}\right)-{\pi^2\over 3}\bigg) \,{\rm .}
\label{1Lratfunc}
\eea
It is important to note that, in eq. (\ref{1Lratfunc}) above, the index $i$ is understood to be mod 3 for the $u_i$ and mod 6 for the $R_{i\,i+3\,i+5}$. Given the validity of eq. (\ref{ratiofunc}) at one loop and six points, it is reasonable to suspect that something similar will happen at higher loops as well. However, a na\"{i}ve extrapolation from one to higher loops is often dangerous. For example, the BDS ansatz is exact at the one-loop $n$-point level, but is incomplete at two loops and six points, as discussed in \ref{BDSsect}. NMHV configurations first appear at the six-point level and, consequently, the first really non-trivial check of (\ref{ratiofunc}) is at two loops and six points. To this end, Kosower, Roiban, and Vergu recently computed the two-loop six-point NMHV superamplitude and they verified (\ref{ratiofunc}) for the parity even part of the ratio function~\cite{KRV}. Before they could check (\ref{ratiofunc}) at $\Ord(a^2)$, they had to resolve a technical problem related to $\e$ poles induced by $\mu$-integrals at the two-loop level.
In order to understand the problem we need to recall the discussion of Subsection \ref{gresults} where we introduced $\mu$-integral hexabox integrals. We did not properly define this integral in \ref{gresults} because it was not necessary at the time. The $\mu$-integral hexabox integral depicted in Figure \ref{hexabox} is given by
\cmb{-.8 in}{0 in}
\bea
&&I_{(4,6)}^{(2);\,D = 4 - 2\e}[\mu^2] = \int {dp^{4-2\e}\over (2\pi)^{4-2\e}}
{1\over p^2 (p-k_2)^2(p-k_1-k_2)^2}\int {dq^4\over (2 \pi)^4}\int {d^{-2\e} \mu \over (2\pi)^{-2\e} }\times
\el
\times{\mu^2 \over ((q+p)^2+2 \vec{\mu}\cdot p-\mu^2)(q^2-\mu^2)((q-k_3)^2-\mu^2)((q-k_3-k_4)^2-\mu^2)((q-k_3-k_4-k_5)^2-\mu^2)((q+k_1+k_2)^2-\mu^2)}\,{\rm .} \nn
\eea
\cme
In this case it turns out that, to leading order, the above integral factorizes~\cite{BDKRSVV} and we can write
\cmb{0 in}{0 in}
\bea
&I_{(4,6)}^{(2);\,D = 4 - 2\e}[\mu^2]& = I_3^{(2);\,D = 4 - 2\e} I_{6}^{D = 4 - 2\e}[\mu^2] = \left(-{1 \over \e^2}(-s_{1})^{-1-\e}\right)\left(-\e I_6^{D = 6 - 2\e}\right)
\elale \left(-{1 \over \e^2}(-s_{1})^{-1-\e}\right)\left(-{\e \over 2} \sum_{i = 1}^6 C_i I_5^{(i);\,D = 6 - 2\e}\right) \,{\rm ,}
\label{hexaboxrel}
\eea
\cme
where the last equality follows from eq. (\ref{hexred}). One can check numerically that (\ref{hexaboxrel}) is valid through $\Ord(\e^0)$; the hexabox $\mu$-integral can be evaluated through $\Ord(\e^0)$, apart from trivial factors, is a $1/\e$ pole times a certain linear combination of the finite one-loop functions $I_5^{(i);\,D = 6}$. In our discussion of planar gluon NMHV amplitudes in Section \ref{gluoncomp}, we noted a close connection between the one-loop pentagon coefficients we calculated and appropriate $\mu$-integral hexabox coefficients. For the sake of concreteness, we go back to the particular example discussed in \ref{gresults}, where we wrote down the relationship between the coefficients of $\e \,I_5^{(2);\,D = 6 - 2\e}$ and $I_{(4,6)}^{(2);\,D = 4 - 2\e}[\mu^2]$:
\be
K_2 = {C_2 \over 2 s_1} \mathcal{K}_2 \, {\rm .}
\ee
If we use the above relation to express the $\mathcal{K}_2$ in terms of $K_2$, we find that the contribution from this NMHV $\mu$-integral hexabox to the ratio function at $\Ord(a^2)$ looks like
\be
\mc{K}_2 I_{(4,6)}^{(2);\,D = 4 - 2\e}[\mu^2] = {(-s_1)^{-\e} K_2 \over \e \,C_2} \sum_{i=1}^6 C_i I_5^{(i);\,D = 6} + \Ord(\e^0)
\label{hexaboxepole}
\ee
To see how this is all related to our one-loop NMHV pentagon coefficients, let us take a step back and remember what we're trying to calculate. Since we want $R_{\rm NMHV}$ to two loops\footnote{In eq. (\ref{ratexp}), $\hat{\a}_{6;3}^{L-{\rm loop}}$ denotes the superamplitude with a factor of $\a_{6;2}^{tree}$ stripped off.}
\bea
R_{\rm NMHV} &=& {\hat{\a}_{6;3}^{tree} + a \,\hat{\a}_{6;3}^{1-{\rm loop}} + a^2 \hat{\a}_{6;3}^{2-{\rm loop}} +\cdots\over 1 + a\, \M^{(1)}(n,t_i^{[r]},\e)+a^2 \M^{(2)}(n,t_i^{[r]},\e)+\cdots}
\elale \hat{\a}_{6;3}^{tree} + a \left(\hat{\a}_{6;3}^{1-{\rm loop}}-\hat{\a}_{6;3}^{tree} \M^{(1)}(n,t_i^{[r]},\e)\right)
\el+ a^2 \left(\hat{\a}_{6;3}^{2-{\rm loop}} - \hat{\a}_{6;3}^{1-{\rm loop}} \M^{(1)}(n,t_i^{[r]},\e) + \hat{\a}_{6;3}^{tree} \M^{(1)}(n,t_i^{[r]},\e)^2 - \hat{\a}_{6;3}^{tree} \M^{(2)}(n,t_i^{[r]},\e)\right)
\el + \Ord(a^3)\,{\rm ,}\nn
\label{ratexp}
\eea
we see that there are other places for us to look for $1/\e$ poles at $\Ord(a^2)$ besides the actual two loop contributions. It is possible for one-loop contributions of $\Ord(\e)$ to hit the universal soft singular terms (see eq. (\ref{1LIR})) in another one-loop contribution and interfere to produce $1/\e$ singularities. For instance, there will be a contribution of the form
\be
-\left(\e\, K_2 I_5^{(2);\,D = 6 - 2 \e}\right)\left(-{1 \over \e^2}\sum^6_{i=1}\left(-s_{i\,i+1}\right)^{-\e}\right) = {K_2 \sum^6_{i=1}\left(-s_{i\,i+1}\right)^{-\e} \over \e} I_5^{(2);\,D = 6 - 2 \e} + \Ord(\e^0)\nn
\label{pentagonepole}
\ee
coming from the cross-term $-\hat{\a}_{6;3}^{1-{\rm loop}} \M^{(1)}(n,t_i^{[r]},\e)$. This shows that, to obtain all IR divergent contributions to the parity even part of $R_{\rm NMHV}$, the even terms in the one-loop NMHV pentagon coefficients of eq. (\ref{sd}) must be included. Indeed, the authors of~\cite{KRV} have checked at the level of superamplitudes using our results that the even part of $R_{\rm NMHV}$ is dual superconformally invariant. It is now possible to explain why the hexabox coefficients derived by Kosower, Roiban, and Vergu in~\cite{KRV} are so similar to our one-loop pentagon coefficients. A close connection between them is necessary for all of the exotic IR divergent contributions (those that have their origin in $\mu$-integrals) to cancel out in the calculation of the ratio function.
Actually, with a modest amount of additional effort, we can simplify our NMHV pentagon coefficients further and explicitly make contact with the form used by Kosower, Roiban, and Vergu in carrying out their analysis of the two-loop NMHV ratio function. Kosower, Roiban, and Vergu make use of a particular rearrangement of eq. (\ref{sd}). This rearrangement is very nice because, with it, the usual MHV level notions\footnote{At the MHV level, the ``even components'' are simply those terms in the amplitude with no explicit factors of $\pol(i,j,k,\ell)$ and the ``odd components'' are those terms with such factors. We remind the reader that $\pol(i,j,k,\ell)$ was defined in eq. (\ref{epstensor}).} of ``even components'' and ``odd components'' actually make sense in the context of the one-loop NMHV amplitude as well.
Recall the form of (\ref{sd}) and collect all terms in the above proportional to each R-invariant structure\footnote{There are six such structures: $R_{3 6 2} + R_{6 3 5}$, $R_{4 1 3} + R_{1 4 6}$, $R_{5 2 4} + R_{2 5 1}$, $R_{3 6 2} - R_{6 3 5}$, $R_{4 1 3} - R_{1 4 6}$, and $R_{5 2 4} - R_{2 5 1}$.}:
\bea
&&\a^{1-{\rm loop}}_{6;3} = \cdots + \frac{i}{6} \e ~\a^{tree}_{6;2}\left\{{\frac{1}{2}}\sum_{i = 1}^6 C_i I^{(i),\,D = 6 - 2\e}_5\right\}\sum_{i = 1}^3 \left(2 s_{i+1} s_{i-2} - t_{i} t_{i+1}\right) t_{i-1} \left(R_{i+2\,i-1\,i+1} + R_{i-1\,i+2\,i-2}\right) \nonumber \\
&& + \frac{i}{6} \e ~\a^{tree}_{6;2}\sum_{i = 1}^3 (-1)^{i} \left( C_i I^{(i),\,D = 6 - 2\e}_5 - C_{i+1} I^{(i+1),\,D = 6 - 2\e}_5 + C_{i-3} I^{(i-3),\,D = 6 - 2\e}_5 - C_{i-2} I^{(i-2),\,D = 6 - 2\e}_5\right)\times
\el\times \Big(\spb1.2 \spa2.3 \spb3.4 \spa4.5 \spb5.6 \spa6.1 - \spa1.2 \spb2.3 \spa3.4 \spb4.5 \spa5.6 \spb6.1\Big) \left(R_{i+2 \, i-1\,i+1} - R_{i-1\,i+2\,i-2}\right) \,{\rm .}
\label{sd2}
\eea
Using eq. (\ref{hexred}), reproduced below for the convenience of reader,
\be
I^{D = 6 - 2\e}_6 = {1 \over 2}\sum_{i = 1}^6 C_i I^{(i),\,D = 6 - 2\e}_5
\ee
the first line of eq. (\ref{sd2}) can be put into a form that bears a close resemblance to the even components of the higher order pieces of the one-loop MHV superamplitude; it is proportional to the one-loop scalar hexagon integral (see eq. (\ref{6ptMHV})), $I^{D = 6 - 2\e}_6$.
In fact, a similar simplification is possible for the terms proportional to $R_{i+2 \, i-1\,i+1} - R_{i-1\,i+2\,i-2}$ as well, although it is not at all obvious. We have numerically checked that
\be
C_i = {2(-1)^i \pol(i+1,i+2,i+3,i+4)\over \spb1.2 \spa2.3 \spb3.4 \spa4.5 \spb5.6 \spa6.1 - \spa1.2 \spb2.3 \spa3.4 \spb4.5 \spa5.6 \spb6.1}\,.
\ee
Using this relation we see that the terms in eq. (\ref{sd2}) not directly proportional to $I^{D = 6 - 2\e}_6$ bear a striking resemblance to the odd components of the higher order pieces of the one-loop MHV superamplitude (again, see eq. (\ref{6ptMHV})). Putting everything together, we find
\bea
&&\a^{1-{\rm loop}}_{6;3} = \cdots + \frac{i}{6} \e ~\a^{tree}_{6;2} I^{D = 6 - 2\e}_6 \sum_{i = 1}^3 \left(2 s_{i+1} s_{i-2} - t_{i} t_{i+1}\right) t_{i-1} \left(R_{i+2\,i-1\,i+1} + R_{i-1\,i+2\,i-2}\right) \nonumber \\
&& + \frac{i}{3} \e ~\a^{tree}_{6;2}\sum_{i = 1}^3 \bigg( \pol(i+1,i+2,i+3,i+4) I^{(i),\,D = 6 - 2\e}_5 + \pol(i+2,i+3,i+4,i+5) I^{(i+1),\,D = 6 - 2\e}_5 \nonumber \\
&&- \pol(i-2,i-1,i,i+1) I^{(i-3),\,D = 6 - 2\e}_5 - \pol(i-1,i,i+1,i+2) I^{(i-2),\,D = 6 - 2\e}_5\bigg)\left(R_{i+2 \, i-1\,i+1} - R_{i-1\,i+2\,i-2}\right)\nn
\label{sdfinal}
\eea
for the higher-order components of the planar one-loop NMHV superamplitude. Eq. (\ref{sdfinal}) is particularly important because it was the form utilized by Kosower, Roiban, and Vergu for their analysis in~\cite{KRV}. It is now clear that, indeed, the notions of even and odd that were used in the context of the planar one-loop MHV superamplitude make sense at the NMHV level as well.
\section{Summary}
\label{sum}
In this work we have discussed several recent developments in the theory of the $\Nsym$ S-matrix. After reviewing some of the most important computational techniques in \ref{revcomp}, we discussed a simple refinement of the $D$ dimensional unitarity technique of Bern and Morgan in \ref{effgcomp}. One notable feature of our approach is that all integrands are reconstructed in $D$ dimensions directly from tree amplitudes without any need for supersymmetric decompositions. While our approach to $D$ dimensional unitarity is probably already familiar to experts in the field, to the best of our knowledge no detailed exposition of the ideas have appeared in print so far.\footnote{A particularly interesting recent study~\cite{BCDHI} further develops $D$ dimensional generalized unitarity along different lines. In future work it would be very nice to make contact with the formalism developed in~\cite{BCDHI}.} We also discuss how our approach to $D$ dimensional unitarity meshes well with the leading singularity method in the context of all-orders-in-$\e$ one-loop $\Nsym$ calculations. In \ref{gresults} we presented simple formulae for the higher-order in $\e$ pentagon coefficients of the planar one-loop six-gluon NMHV amplitudes $A^{1-{\rm loop}}_{1}(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6)$, $A^{1-{\rm loop}}_{1}(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6)$, and $A^{1-{\rm loop}}_{1}(k_1^{1234},k_2,k_3^{1234},k_4,k_5^{1234},k_6)$. Na\"{i}vely, these results may seem rather useless because, if one only cares about the massless $\Nsym$ S-matrix, one never needs the pentagon coefficients.
However, we argue in \ref{gsrel} that, actually, the higher-order in $\e$ pentagon coefficients are useful because they contain non-trivial information about tree-level scattering of massless modes in open superstring theory. After reviewing the non-Abelian Born-Infeld action in \ref{BornInfeld}, we argued in \ref{results} that matrix elements of the non-Abelian Born-Infeld action at $\Ord(\alpha'^2)$ and $\Ord(\alpha'^3)$ can be predicted from all-orders-in-$\e$ $\Nsym$ amplitudes dimensionally shifted to either $D = 8-2\e$ or $D = 10-2\e$. As an amusing by-product of our analysis, we were able to use another close connection between the one-loop all-plus amplitudes in pure Yang-Mills and our stringy corrections at $\Ord(\alpha'^2)$ to understand the vanishing of the all-plus amplitudes when three or more gluons are replaced by photons for $n > 4$.
At this point, in Section \ref{supercomp}, we explained how to supersymmetrize the results of \ref{gresults}. To this end, we introduced the $\Nsym$ on-shell superspace in \ref{gendisonshell} and discussed some important examples of $\Nsym$ superamplitudes. In \ref{ffbarcalc}, we first explain that, following Elvang, Freedman, and Kiermaier, one can choose the five component amplitudes $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1234},k_4,k_5,k_6\right)$, $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^{4},k_5,k_6\right)$,\\ $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{12},k_4^{34},k_5,k_6\right)$,$A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)$, \\and $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3,k_4^{1234},k_5,k_6\right)$ and determine the full $\Nsym$ superamplitude in terms of them. We then showed how to extend the methods of \ref{effgcomp} to deal with\\ $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{123},k_4^{4},k_5,k_6\right)$ and $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{1},k_4^{234},k_5,k_6\right)$. It is crucial that our techniques be applicable to amplitudes with external fermions if we want them to be useful for theories with less supersymmetry such as QCD. Finally, in \ref{ssbarcalc} we determine $A_1^{1-{\rm loop}}\left(k_1^{1234},k_2^{1234},k_3^{12},k_4^{34},k_5,k_6\right)$ indirectly and in \ref{supersymresults} we complete the process by collecting our results. We write down for the first time the higher-order pentagon contributions to the six-point NMHV $\Nsym$ superamplitude.
We then show in \ref{DSI} (after reviewing some of the developments that led to the discovery of dual superconformal invariance in \ref{BDSsect} and \ref{WL/MHV2}) that the superamplitude takes on a significantly simpler form if expressed in terms of the R-invariants of Drummond, Henn, Korchemsky, and Sokatchev. Remarkably, in this form, the pentagon coefficients are related by cyclic symmetry. We can understand the greater simplicity of this formula by comparing the explicit operator realization of the ordinary and dual superconformal symmetries. Some of the ordinary superconformal generators are expressed in terms of 2nd-order partial differential operators, whereas {\it all} of the dual superconformal generators are expressed in terms of 1st-order differential operators. This is to be expected since differences of dual variables are just momenta; the dual superconformal symmetry acts naturally in momentum space. As a result, it is not too surprising that, when expressed in terms of R-invariants, the pentagon coefficients look even simpler than those presented in \ref{supersymresults}, where dual superconformal symmetry was obscured.
Finally, in \ref{rationfunc} we explain the relevance of our results to the study of the dual superconformal properties of (the parity even part of) the two-loop NMHV ratio function in dimensional regularization. Our higher-order-in-$\e$ pentagon coefficients can interfere with $1/\e^2$ poles in the one-loop MHV superamplitude to produce contributions of order $1/\e$. Thus, the results written down in \ref{DSI} in terms of dual superconformal R-invariants are necessary to produce a finite result for the two-loop NMHV ratio function in on-shell superspace if one is working in dimensional regularization. To this end, we further improve the results presented in \ref{DSI} by seperating them into even and odd components. This decomposition is clearly natural because the results of \ref{DSI} simplify still more. In particular, the even components of the higher order pieces of the one-loop NMHV superamplitude can be rearranged and put into a form where they are actually proportional to the one-loop hexagon integral in $D = 6 - 2 \e$. This feature of the even components was exploited recently in a study of the two-loop NMHV ratio function by Kosower, Roiban, and Vergu~\cite{KRV}.
There has been a tremendous amount of recent progress on the planar $\Nsym$ S-matrix\footnote{Most, if not all, of the works cited here were significantly influenced by the seminal work of Witten~\cite{witten}.}~\cite{newpap1,newpap2,newpap3,newpap4,newpap5,newpap6,newpap7,newpap8,newpap9,newpap10,newpap11,newpap12,newpap13,newpap14,newpap15,newpap16,newpap17,newpap18,newpap19,newpap20,newpap21,newpap22,newpap23,newpap24,newpap25,newpap26,newpap27,newpap28,newpap29,DelDuca1,DelDuca2,DelDuca3,newpap30,newpap32,newpap33,newpap34,newpap35,newpap36}, which, unfortunately, we don't have time to say much about. A recurring theme in recent papers on the subject is the idea that one should be able to learn everything there is to know about the planar $\Nsym$ S-matrix using only four dimensional information. In Subsection \ref{effgcomp} we showed that, generically, four dimensional generalized unitarity cuts are not sufficient to determine one-loop planar $\Nsym$ scattering amplitudes to all orders in the dimensional regularization parameter. Clearly, the analysis of Subsection \ref{results} suggests that the one-loop pentagon coefficents missed by the leading singularity method are important and should be determined independently ({\it e.g.} by using maximal generalized unitarity in $D$ dimensions). However, in the spirit of the recent developments, we should first check whether, perhaps, our predictions for the $\Ord(\alpha'^2)$ and $\Ord(\alpha'^3)$ stringy corrections to $\Nsym$ amplitudes don't really rely on {\it all} pentagon coefficients but only some linear combination thereof.
Recall from the discussion of Subsection \ref{effgcomp} that, at the one-loop $n$-point level, amplitudes computed via the leading singularity method are not uniquely determined to all orders in $\e$ but have
\be
{(n-5)(n-4)(n-3)(n-2)(n-1)\over 120}
\ee
pentagon coefficients that must be determined by some other method. It is conceivable that, after performing the dimension shift operation and summing over all contributions, all the undetermined coefficients actually drop out. In fact, there is evidence that this happens for $\Nsym$ amplitudes dimensionally shifted to $D = 8 - 2\e$; we checked that we could derive the appropriate tree-level stringy corrections at $\Ord(\alpha'^2)$ for both MHV and NMHV $n = 6$ amplitudes, $n = 7$ MHV amplitudes, and $n = 8$ MHV amplitudes using the leading singularity method (without $D$ dimensional unitarity). However, for the $\Ord(\alpha'^3)$ stringy corrections, this no longer works. For example, one can check that the one-loop $\Nsym$ six-point MHV amplitude cannot be used to compute $A^{tree}_{str}\left(k_1^{1234},k_2^{1234},k_3,k_4,k_5,k_6\right)$ at $\Ord(\alpha'^3)$ unless {\it all} pentagon coefficients in the amplitude are determined. Our conclusion is that there are still some questions that can be answered by calculating $\Nsym$ amplitudes to all orders in $\e$ that cannot (at least not obviously) be answered by calculating amplitudes in a framework that requires only four dimensional inputs.
In this review we have seen that our approach to all-orders one-loop $\Nsym$ scattering amplitudes opens up several interesting avenues of exploration. Besides the connection that we found between stringy corrections and dimensionally shifted one-loop amplitudes, it seems plausible, for example, that a variant of our approach will be the right way to think about computing scattering amplitudes at a generic point in the $\Nsym$ moduli space. It will also be quite interesting to see whether our approach to $D$ dimensional integrand construction is useful for theories with less or no supersymmetry. Our expectation is that the approach advocated here will continue to be relevant and useful for future higher-loop studies of $\Nsym$ and more general amplitudes in dimensional regularization.
\section*{Acknowledgments}
I am grateful to Matt Strassler and Lance Dixon for their guidance, support, and collaboration over the last few years. I would like to thank the theory group at SLAC for its hospitality the last few summers and Carola Berger, Zvi Bern, Darren Forde, Tanju Gleisberg, and Daniel Ma\^{i}tre for enlightening conversations and correspondence. I am also very grateful to Cristian Vergu for enlightening correspondence and collaboration. I am especially grateful to Fernando Alday, Nima Arkani-Hamed, Jacob Bourjaily, Freddy Cachazo, James Drummond, Henriette Elvang, Johannes Henn, Juan Maldacena, David McGady, and Jaroslav Trnka for helping me to better understand exciting recent developments in planar $\Nsym$ gauge theory and related topics. A special acknowledgment goes to Johannes Henn, Karl Landsteiner, Andy O'Bannon, Radu Roiban, Matt Strassler, and Cristian Vergu for providing me with some very useful comments and advice on earlier drafts of this work. In this work, figures were drawn primarily using the Jaxodraw~\cite{JaxD} package. This research was supported in part by the US Department of Energy under contract DE-FG02-96ER40949.
|
\section{Introduction}
Application of category theory to rational conformal field theory has a long history. Representations of chiral symmetries of a rational conformal field theory form a certain kind of braided tensor category known as a {\em modular} category \cite{ms,Hu} and many features of conformal field theory such as boundary extensions, bulk field space, symmetries of conformal field theory have neat categorical interpretations. In particular category theory is well suited for studying extensions of chiral algebras. According to A. Kirillov, Y.-Z. Huang and J. Lepowsky (see \cite{KiO} and references therein) a chiral extension of a chiral algebra ${\cal{V}}$ corresponds to a certain commutative algebra in the modular category of representations of ${\cal{V}}$. More precisely the commutative algebras in question should be simple, separable and ribbon (concepts explained in section \ref{prel}). Moreover the category of representations of the extended chiral algebra can be read off from the corresponding commutative algebra as the category of its local modules. Relatively simple categorical arguments show that there are only a finite number of simple, separable, ribbon, commutative algebras in a given modular category and that all maximal algebras have equivalent categories of local modules (see \cite{dmno} and references therein). This immediately implies that a rational chiral algebra has only a finite number of extensions. Moreover maximal extensions of a rational chiral algebra all have the same representation type.
All this indicates the special role played by maximal (rational) chiral algebras and their categories of representations. The property that characterises categories of representations of maximal chiral algebras is the absence of non-trivial separable, ribbon, commutative algebras in the category. Such categories were called {\em completely anisotropic} in \cite{dmno}.
This simplest class of maximal chiral algebras is formed by holomorphic chiral algebras; that is, chiral algebras with no non-trivial representations. Clearly tensor products of holomorphic algebras are again holomorphic.
Here we look at another class of maximal chiral algebras closed under tensor products. This class (as with the class of holomorphic chiral algebras) is defined by their categories of representations.
In this paper we deal with a specific type of modular category with just two simple objects: $I$ and $X$; and the tensor product decompositions:
$$I{\otimes} I = I,\quad X{\otimes} I = I{\otimes} X = X,\quad X{\otimes} X = I\oplus X.$$
Such categories are called {\em Fibonacci} categories. There are four non-equivalent Fibonacci modular categories $\mathcal{F}ib_u$, labelled by primitive roots of unity $u$ of order 10 (section \ref{secfib}).
We prove that tensor powers $\mathcal{F}ib_u^{\boxtimes l}$ of a Fibonacci modular category are completely anisotropic; that is, they do not have non-trivial separable, ribbon, commutative algebras. We do it by describing nonnegative integer matrix (or NIM-) representations of the fusion rules of $\mathcal{F}ib_u^{\boxtimes l}$. They give us an insight into possible module categories over $\mathcal{F}ib_u^{\boxtimes l}$. Since an algebra in a monoidal category gives rise to a module category, our knowledge of NIM-representations for $\mathcal{F}ib_u^{\boxtimes l}$ allows us to prove complete anisotropy of $\mathcal{F}ib_u^{\boxtimes l}$. Actually our methods give a stronger result. A product of Fibonacci categories $\mathcal{F}ib_u$ is completely anisotropic as long as none of the indices $u$ is inverse to each other. The condition that $uv\not=1$ is not accidental. General category theory implies that the category $\mathcal{F}ib_u\boxtimes\mathcal{F}ib_{u^{-1}}$ always has a non-trivial simple, separable, ribbon, commutative algebra.
Complete anisotropy of Fibonacci categories implies that chiral algebras with categories of representations of the form $\mathcal{F}ib_u^{\boxtimes l}$ are maximal.
Among such chiral algebras are the chiral algebra $M(2,5)$ from the minimal series; that is, the chiral algebra of the Yang-Lee model (the category of representations is $\mathcal{F}ib_u$, with $u=e^{\frac{\pi i}{5}}$), a maximal extension of $M(3,5)^{\times 8}\times M(2,5)^{\times 7}$ (the category of representations is $\mathcal{F}ib_u$, with $u=e^{\frac{9\pi i}{5}}$), the affine chiral algebra $G_{2,1}$ (the category of representations is $\mathcal{F}ib_u$, with $u=e^{\frac{3\pi i}{5}}$) and the affine chiral algebra $F_{4,1}$ (with the category of representations $\mathcal{F}ib_u$ for $u=e^{\frac{7\pi i}{5}}$). Clearly tensoring with holomorphic algebras keeps the class of Fibonacci chiral algebras closed. Among examples of chiral algebras, which are not products of the above mentioned Fibonacci chiral algebras with holomorphic algebras is the coset $F_{4,1}/G_{2,1}$ (the category of representations is $\mathcal{F}ib_u\boxtimes\mathcal{F}ib_u$ with $u=e^{\frac{7\pi i}{5}}$). This coset can not be written as a tensor product of two chiral algebras of Fibonacci type.
\section{Preliminaries}\label{prel}
Throughout the paper we assume that the ground field $k$ is an algebraically closed field of characteristic zero (such as the field ${\bf C}$ of complex numbers).
\subsection{Modular categories}
Recall that a rigid monoidal category is called {\em fusion} when it is semi-simple $k$-linear together with a $k$-linear tensor product, finite-dimensional hom spaces and a finite number of simple objects (up to isomorphism).
We denote the set (of representatives) of isomorphism classes of simple objects in ${\cal C}$ by $Irr({\cal C})$.
Slightly changing the definition from \cite{tu}, we call a fusion category {\em modular} if it is rigid, braided, ribbon and satisfies the non-degeneracy ({\em modularity}) condition: for isomorphism classes of simple objects, the traces of the double braidings form a non-degenerate matrix
$$\tilde{S} = (\tilde{S}_{X,Y})_{X,Y\in Irr({\cal C})},\quad \tilde{S}_{X,Y} = tr(c_{X,Y}c_{Y,X}).$$
Here $c_{X,Y}:X\otimes Y\to Y\otimes X$ is the braiding (see \cite{tu,BaKi} for details).
Let ${\cal C}$ be a ribbon category.
Following \cite{tu} define $\overline{{\cal C}}$ to be ${\cal C}$ as a monoidal category equipped with a new braiding and ribbon twist:
$$\overline{c}_{X,Y}=c_{Y,X}^{-1},\quad \overline{\theta}_X = \theta_X^{-1}.$$
Again it is very easy to see that for a modular ${\cal C}$, $\overline{\cal C}$ is also modular.
Recall that the Deligne tensor product ${\cal C}\boxtimes{\cal D}$ of two fusion categories is a fusion category with simple objects $Irr({\cal C}\boxtimes{\cal D}) = Irr({\cal C})\times Irr({\cal D})$ and the tensor product defined by
$$(X\boxtimes Y)\otimes (Z\boxtimes W) = (X\otimes Z)\boxtimes(Y\otimes W).$$
It is straightforward to see that the Deligne tensor product of two modular categories is modular.
Let ${\cal C}$ be a full modular subcategory of a modular category ${\cal D}$. It was proved in \cite{mu1} that as modular categories
\begin{equation}\label{mudec}{\cal D} = {\cal C}\boxtimes{\cal C}_{\cal D}({\cal C}),\end{equation}
where the category ${\cal C}_{\cal D}({\cal C})$ (the {\em M\"uger's centraliser} of ${\cal C}$ in ${\cal D}$) is defined as the full subcategory of ${\cal D}$ of objects {\em transparent} with respect to objects of ${\cal C}$:
$${\cal C}_{\cal D}({\cal C}) = \{X\in{\cal D}|\ c_{Y,X}c_{X,Y} = 1_{X{\otimes} Y},\ \forall Y\in{\cal C}\}.$$
Let ${\cal Z} ({\cal C})$ be the monoidal centre of ${\cal C}$ \cite{js2}. Recall that a braiding $c$ in ${\cal C}$ gives rise to a braided monoidal functor $\iota_+:{\cal C}\to {\cal Z} ({\cal C})$.
Using the (conjugate) braiding $\overline c$ we can define another braided monoidal functor $\iota_-:\overline{\cal C}\to {\cal Z}({\cal C})$.
Taking the Deligne tensor product we can combine these two functors into one braided monoidal functor $$\iota:{\cal C}\boxtimes\overline{\cal C}\to {\cal Z}({\cal C}).$$
The following characterisation of modularity was proven in \cite{mu}.
A braided fusion category ${\cal C}$ is modular if and only if the functor $\iota$ is an equivalence.
\begin{example}
We call a fusion category ${\cal C}$ {\em pointed} if all its simple objects are invertible, i.e. $X\otimes X^*\cong 1$ for any simple $X\in{\cal C}$. In this case the set $Irr({\cal C})$ of isomorphism classes of simple objects is a group (with respect to the tensor product). Clearly for a braided ${\cal C}$ this group must be abelian.
It was shown in \cite{js} that (up to braided equivalence) braided structures on a pointed category ${\cal C}$ with the group $A = Irr({\cal C})$ are in one-to-one correspondence with functions $q:A\to k^*$ ($k^*$ denotes the multiplicative group of $k$) satisfying
$q(a^{-1}) = q(a)$ and such that $\sigma(a,b)= q(ab)q(a)^{-1}q(b)^{-1}$ is bilinear in $a$ and $b$ (a {\em bicharacter}). The correspondence assigns to a braiding $c$ the function $q$ such that $c_{a,a} = q(a)1_{a\otimes a}$. We will denote by ${\cal C}(A,q)$ the pointed category with the group of objects $A$ and the braiding corresponding to $q$. The following are straightforward
$${\cal C}(A\times B,q_A\times q_B) \simeq {\cal C}(A,q_A)\boxtimes{\cal C}(B,q_B),\quad \overline{{\cal C}(A,q)} \simeq {\cal C}(A,q^{-1}).$$
The braiding of ${\cal C}(A,q)$ is non-degenerate if and only if the form $\sigma$ is non-degenerate:
$$ker(\sigma) = \{a\in A|\ \sigma(a,b)=1\ \forall b\in A\} = 1.$$
Modular structures (ribbon twists) on the braided category ${\cal C}(A,q)$ correspond to homomorphisms $d:A\to {\bf Z}/2{\bf Z}$ (the dimension function). The ribbon twist $\theta$ corresponding to $d$ has the form
$$\theta_a = d(a)q(a)1_a.$$
\end{example}
\subsection{Module categories, algebras in monoidal categories and their modules}\label{algandmods}
Let $\cal{V}$ be a monoidal category.
A \emph{module category} (see \cite{qu,mkgj,mc}) is a category ${\cal{M}}$ together with a functor $\ast: \cal{V}\times{\cal{M}}\to{\cal{M}}$ and natural families of isomorphisms
$$\alpha_{X,Y,M}: (X\otimes Y)\ast M \to X\ast(Y\ast M)$$
and,
$$\lambda_{M}: I\ast M \to M$$
such that diagrams (1.1), (1.2) and (1.3) of \cite{mkgj} commute.
An (associative, unital) {\em algebra} in a monoidal category ${\cal C}$ is a triple $(A,\mu,\iota)$ consisting of an object $A\in{\cal C}$ together with a {\em multiplication} $\mu:A\otimes A\to A$ and a {\em unit} map $\iota:I\to A$, satisfying {\em associativiy}
$$\mu(\mu\otimes 1) = \mu(1\otimes\mu),$$
and {\em unit}
$$\mu(\iota\otimes 1) = 1 =\mu(1\otimes\iota)$$
axioms.
Where it will not cause confusion we will be talking about an algebra $A$, suppressing its multiplication and unit maps. A {\em homomorphism} of algebra is a morphism of underlying objects preserving algebra structures in the obvious way. An algebra is {\em simple} if any (non-zero) homomorphism out of it is injective.
A right {\em module} over an algebra $A$ is a pair $(M,\nu)$, where $M$ is an object of ${\cal C}$ and $\nu:M\otimes A\to M$ is a morphism ({\em action map}), such that
$$\nu(\nu\otimes 1) = \nu(1\otimes\mu).$$
A {\em homomorphism} of right $A$-modules $M\to N$ is a morphism $f:M\to N$ in ${\cal C}$ such that
$$\nu_N(f\otimes 1) = f\nu_M.$$
Right modules over an algebra $A\in{\cal C}$ together with module homomorphisms form a category ${\cal C}_A$.
The forgetful functor ${\cal C}_A\to{\cal C}$ has a right adjoint, which sends an object $X\in{\cal C}$ into the {\em free} $A$-module $X\otimes A$, with $A$-module structure defined by
$$\xymatrix{X\otimes A\otimes A \ar[r]^-{1\mu} & X\otimes A.}$$
Note that, since the action map $M\otimes A\to M$ is an epimorphism of right $A$-modules, any right $A$-module is a quotient of a free module.
More generally, for any right $A$-module $M$ and any $X\in{\cal C}$ the tensor product $X\otimes M$ has a structure of a right $A$-module
$$\xymatrix{X\otimes M\otimes A \ar[r]^-{1\mu} & X\otimes M.}$$
This makes the category of modules ${\cal C}_A$ a left module category over ${\cal C}$.
The adjoint pairing
$$\xymatrix{{\cal C}_A \ar@/^10pt/[rr]^-{U} &\perp & {\cal C} \ar@/^10pt/[ll]^-{-\otimes A} }$$
consisting of the forgetful and free $A$-module functors is an adjoint pair of ${\cal C}$-module functors.
The next notion provides algebras with semi-simple categories of modules.
An algebra $(A,\mu,\iota)$ in a rigid braided monoidal category ${\cal C}$ is called {\em separable} if the following composition (denoted $e:A\otimes A\to I$) is a non-degenerate pairing:
$$\xymatrix{A\otimes A \ar[r]^(.6)\mu & A \ar[r]^{\epsilon} & I.}$$
Here $\epsilon$ is the composite
$$\xymatrix{A \ar[r]^-{1{\otimes}\kappa_A} & A\otimes A\otimes A^* \ar[r]^-{\mu{\otimes} 1} & A\otimes A^* \ar[r]^-{c_{A,A^*}} & A^*\otimes A \ar[r]^-{ev_A} & I,}$$
where $\kappa_A$ and $ev_A$ are duality morphisms for $A$.
Non-degeneracy of $e$ means that there is a morphism $\kappa: I\to A\otimes A$ such that the composition
$$\xymatrix{A \ar[r]^-{1{\otimes}\kappa} & A^{\otimes 3} \ar[r]^-{e{\otimes} 1} & A}$$
is the identity.
It also implies that the similar composition
$$\xymatrix{A \ar[r]^-{\kappa{\otimes} 1} & A^{\otimes 3} \ar[r]^-{1{\otimes} e} & A}$$
is also the identity.
Using the graphical calculus for morphisms in a (rigid) monoidal category \cite{js1} one can represent morphisms between tensor powers of a separable algebra by graphs (one dimensional CW-complexes), whose end vertices are separated into incoming and outgoing.
For example, the multiplication map $\mu$ is represented by a trivalent graph with two incoming and one outgoing ends, the duality $\epsilon$ is an interval, with both incoming ends etc.
It turns out (e.g in symmetric case, it follows from the results of \cite{rsw}) that separability implies that we can contract loops in connected graphs with at least one end.
For a separable algebra $A$ the adjunction
$$\xymatrix{{\cal C} \ar@/^5pt/[r] & {\cal C}_A \ar@/^5pt/[l] }$$
splits.
Indeed, the splitting of the adjuction map $M\otimes A\to M$ is given by the projector $M\otimes A\to M\otimes A$:
$$\xymatrix{M\otimes A \ar[r]^-{I{\otimes}\epsilon} & M\otimes A^{\otimes 3} \ar[r]^-{I{\otimes}\mu{\otimes} I} & M\otimes A^{\otimes 2} \ar[r]^-{\nu{\otimes} I} & M\otimes A.}$$
For a separable algebra $A$ the effect on morphisms ${\cal C}_A(M,N)\to{\cal C}(M,N)$ of the forgetful functor ${\cal C}_A\to{\cal C}$ has a splitting $P:{\cal C}(M,N)\to{\cal C}_A(M,N)$:
$$P(f) = f,\quad f\in {\cal C}_A(M,N).$$
For $f\in{\cal C}(M,N)$ the image $P(f)$ is defined as the composition
$$\xymatrix{M \ar[r]^-{I{\otimes}\epsilon} & M\otimes A^{\otimes 2} \ar[r]^-{\nu_M{\otimes} I} & M\otimes A \ar[r]^-{f{\otimes} I} & N\otimes A \ar[r]^-{\nu_N} & N.}$$
Moreover, the splitting has the properties
$$P(fg)=fP(g)\ P(gh)=P(g)h,\quad f,h\in Mor\,({\cal C}_A),\ g\in Mor\,({\cal C} ).$$
This gives {\em Maschke's lemma} for separable algebras.
\begin{lemma}
Let $A$ be a separable algebra in a semi-simple rigid monoidal category ${\cal C}$. Then the category ${\cal C}_A$ of right
$A$-modules in ${\cal C}$ is also semi-simple.
\end{lemma}
\subsection{Commutative algebras and local modules}\label{coal}
Now let ${\cal C}$ be a braided monoidal category with the braiding $c_{X,Y}:X\otimes Y\to Y\otimes X$ (see \cite{js} for the definition).
An algebra $A$ in ${\cal C}$ is {\em commutative} if $\mu c_{A,A} = \mu$.
It was shown in \cite{pa} that the category ${_A}{{\cal C}}$ of left modules over a commutative algebra $A$ is monoidal with respect to the tensor product $M\otimes_AN$ over $A$, which can be defined by a coequaliser
$$\xymatrix{M\otimes_AN & M\otimes N \ar[l] && M\otimes A\otimes N \ar@/^5pt/[ll]^{(\nu_M1)(c_{M,A}1)} \ar@/_5pt/[ll]_{1\nu_N} }.$$
Moreover for commutative algebra $A$ the free functor ${\cal C}\to{\cal C}_A$ is (strong) monoidal which means the multiplication in $A$ induces an isomorphism
$$(X\otimes A)\otimes_A(Y\otimes A) \to (X\otimes Y)\otimes A.$$
A (right) module $(M,\nu)$ over a commutative algebra $A$ is {\em local} iff the diagram
$$\xymatrix{M\otimes A \ar[r]^\nu \ar[d]_{c_{M,A}} & M\\ A\otimes M \ar[r]^{c_{A,M}} & M\otimes A \ar[u]_\nu}$$
commutes. Denote by ${\cal C}_A^{loc}$ the full subcategory of ${\cal C}_A$ consisting of local modules. The following result was
established in \cite{pa}.
\begin{proposition}
The category ${\cal C}_A^{loc}$ is a full monoidal subcategory of ${\cal C}_A$. Moreover, the braiding in ${\cal C}$ induces a braiding in
${\cal C}_A^{loc}$.
\end{proposition}
The following statement was proved in \cite{ffrs}.
\begin{proposition}\label{algloc}
Let $(A,m,i)$ be a commutative algebra in a braided category ${\cal C}$. Let $B=(B,\mu,\iota)$ be an algebra in ${{\cal C}}_{A}$.
Define $\overline\mu$ and $\overline\iota$ as compositions $$\xymatrix{B\otimes B \ar[r] & B\otimes_AB \ar[r]^\mu & B,
& & & 1 \ar[r]^i & A \ar[r]^\iota & B.}$$ Then $\overline B=(B,\overline\mu,\overline\iota)$ is an algebra in ${\cal C}$.
\newline
The map $\iota:A\to B$ is a homomorphism of algebras in ${\cal C}$.
\newline
The algebra $\overline B$ in ${\cal C}$ is separable or commutative if and only if the algebra
$B$ in ${{\cal C}}_{A}$ is such.
\newline
The functor $({\cal C}_A^{loc})_B^{loc}\to {\cal C}_{\overline B}^{loc}$
\begin{equation}\label{locloc}
(M,m:B\otimes_A M\to M)\mapsto (M,\overline m:B\otimes M\to B\otimes_A M\stackrel{m}{\to} M)
\end{equation}
is a braided monoidal equivalence.
\end{proposition}
A commutative algebra $A$ in a ribbon category ${\cal C}$ (with the ribbon twist $\theta$) is called {\em ribbon} if $\theta_A=1_A$.
The next theorem is a part of theorem 4.5 from
\cite{KiO}.
\begin{theorem}\label{chiext}
Let $A$ be an indecomposable separable commutative, ribbon algebra in a modular category ${\cal C}$. Then
${\cal C}_A^{loc}$ is a modular category.
\end{theorem}
We call a commutative, separable, indecomposable, ribbon algebra $A\in{\cal C}$ {\em trivialising} (or {\em Lagrangian} in the terminology of \cite{dmno}) if ${\cal C}_A^{loc}$ is equivalent to the category ${\cal V}{\it ect}$ of vector spaces over the base field (that is, the only simple local $A$-module is $A$ itself).
For any modular category ${\cal C}$ the category ${\cal C}\boxtimes\overline{{\cal C}}$ always has a trivialising algebra $Z$, which we call the {\em diagonal} algebra (the {\em tube} algebra of \cite{mu}) with underlying object
$$Z = \oplus_{X\in Irr({\cal C})}X\boxtimes X^*.$$
\subsection{Fusion rules and modular data}
A set $R$ is called a \emph{fusion rule} if its integer span $\mathbb{Z}R$ is equipped with a structure of an associative unital ring such that the unit element of $\mathbb{Z}R$ belongs to $R$ and
$$r\cdot s \in \mathbb{Z}_{\geq 0}R$$
for any $r,s\in R$.
Here $ \mathbb{Z}_{\geq 0}R$ is the sub ring in $\mathbb{Z}R$ of linear combinations of elements of $R$ with non-negative coefficients.
We also require $\mathbb{Z}R$ to satisfy a \emph{rigidity condition}.
\newline
To formulate the rigidity condition equip $\mathbb{Z}R$ with a symmetric bilinear form $(-,-)$ defined by
$$(r,s) = \delta_{r,s}\quad r,s\in R.$$
Note that for $x\in\mathbb{Z}_{\geq 0}R$ we have
$$(x,x)=1 \Longleftrightarrow x\in R.$$
The rigidity condition is then the existence of an involution $(-)^{\ast}:R\to R$ such that
$$(r\cdot s,t) = (s, r^{\ast}t)\quad r,s,t\in R.$$
By \emph{a homomorphism of fusion rules} $R\to S$ we mean a map of sets which induces a homomorphism of rings $\mathbb{Z}R\to\mathbb{Z}S$.
If $R$ and $S$ are fusion rules then so is $R\times S$.
In particular $R^{\times\ell} = R^{\ell}$ has the structure of a fusion rule.
Let ${\cal{C}}$ be a semi-simple rigid monoidal category.
Then the set $Irr({\cal{C}})$ of isomorphism classes of simple objects in ${\cal{C}}$ has the structure of a fusion rule.
Note that $\mathbb{Z} Irr({\cal{C}})$ coincides with the Grothendieck ring $K_{0}({\cal{C}})$.
The {\em categorical dimension} of ${\cal C}$ is $\mbox{Dim}({\cal C}) = \sum_{X}d(X)^2$, where the sum is taken over isomorphism classes of simple objects $X$ of ${\cal C}$.
Here $d(X) = tr(I_X)$ is the dimension of $X$.
We use the definition of the {\em multiplicative central charge} of a modular category ${\cal C}$ given in \cite[Section~ 6.2]{DGNO}
$$\xi({\cal C}) = \frac{1}{\sqrt{\mbox{dim}({\cal C})}}\sum_{X}\theta(X)d(X)^2,$$
(the sum again is taken over simple objects) where $\theta_X = \theta(X)1_X$ is the twist on a simple object $X\in{\cal C}$.
We take the positive square root $\sqrt{\mbox{dim}({\cal C})}$ of the positive real cyclotomic number $\mbox{dim}({\cal C})$.
The following properties are well known, see for example \cite[Section~ 3.1]{BaKi}.
\begin{lemma} \label{xi}
\begin{enumerate}
\item[(i)] $\xi({\cal{C}})$ is a root of unity;
\item[(ii)] $\xi({\cal{C}}_1\boxtimes {\cal{C}}_2)=\xi({\cal{C}}_1)\xi({\cal{C}}_2)$;
\item[(iii)] $\xi(\overline{\cal{C}})=\xi({\cal{C}})^{-1}$.
\end{enumerate}
\end{lemma}
Let $SL_2({\bf Z})$ be the \emph{modular group}; that is, the group of determinant 1 integer $2\times 2$-matrices.
It is generated by the matrices
$$s = \left(\begin{array}{rr} 0 & -1\\ 1 & 0 \end{array}\right),\quad t = \left(\begin{array}{rr} 1 & 1\\ 0 & 1 \end{array}\right),$$
with the generating system of relations $s^4 = 1, (ts)^3 = s^2$.
Let ${\cal C}$ be a modular category and define
$$S = (\sqrt{Dim({\cal C})})^{-1}\tilde S,\quad T = \xi({\cal C})^{-\frac{1}{3}}diag(\theta_X)$$
where $X$ runs through isomorphism classes of simple objects of ${\cal C}$ and $\tilde S$ is the matrix defined in section (1.1).
The pair of matrices $S,T$ is often referred to as the {\em modular data} of ${\cal C}$.
The proof of the following result can be found in \cite{tu}.
\begin{theorem}
Let ${\cal C}$ be a modular category.
Then the operators $S$ and $T$ define an action of the modular group $SL_2({\bf Z})$ on the complexified Grothendieck group $K_0({\cal C})\otimes{\bf C}$.
\end{theorem}
The following is the second part of theorem 4.5 from
\cite{KiO}.
\begin{theorem}
The map $K_0({\cal C}_A^{loc})\otimes_{\bf Z}{\bf C}\to K_0({\cal C})\otimes_{\bf Z}{\bf C}$, induced by the
forgetful functor ${\cal C}_A^{loc}\to{\cal C}$ is $SL_2({\bf Z})$-equivariant.
\end{theorem}
\subsection{NIM-representations}\label{nimrepsec}
Let $R$ be a fusion rule.
\noindent A set $M$ is a {\em non-negative integer matrix} (or NIM-)representation of $R$ if $\mathbb{Z}M$ is equipped with a structure of an $\mathbb{Z}R$-module such that
$$r\cdot m \in \mathbb{Z}_{\geq 0}M\quad \forall r \in R,\ m\in M$$
and $\mathbb{Z}M$ also possesses the rigidity condition.
As before, to formulate the rigidity condition observe that $\mathbb{Z}R$ comes equipped with a symmetric bilinear form
$$(m,n) = \delta_{m,n};\, m,n\in M.$$
Again, it is obvious that for all $m\in\mathbb{Z}_{\geq 0}M$ we have
$$(m,m)=1 \Longleftrightarrow r\in M.$$
The rigidity condition for NIM-representations is then
$$(r\cdot m,n) = (m, r^{\ast}n)\quad r\in R,\, m,n\in M.$$
Note that a fusion rule $R$ is always a NIM-representation of itself.
More generally a homomorphism of fusion rules $R\to S$ turns $S$ into a NIM-representation of $R$.
A \emph{morphism of NIM-representations} $N\to M$ is a map of sets inducing a $\mathbb{Z}R$-module homomorphism $\mathbb{Z}N\to\mathbb{Z}M$.
The following is a complete reducibility statement for NIM-representations of a rigid fusion rule.
\begin{lemma}\label{cr}
Let $R$ be a fusion rule and let $N \subset M$ be an embedding of NIM-representations of $R$.
Then $$M\setminus N = \{m\in M|\quad m\not\in N\}$$ is a NIM-subrepresentation of $M$ and $M = N \sqcup (M\setminus N)$.
\end{lemma}
\begin{proof}
Note that $ \mathbb{Z}_{\geq 0}(M\setminus N)$ can be identified with the orthogonal complement of $\mathbb{Z}_{\geq 0}N$ in $\mathbb{Z}_{\geq 0}M$. Now for $m\in M\setminus N$, $r\in R$ and $n\in N$ we have
$$(r\cdot m, n) = (m, r^{\ast}n) = 0$$
since $r^{\ast}n\in\mathbb{Z}_{\geq 0}N$ and $(M\setminus N, N) = 0$.
Thus, $R\cdot(M\setminus N) \subset \mathbb{Z}_{\geq 0}(M\setminus N)$.
\end{proof}
Let ${\cal{M}}$ be a semi-simple module category over a semi-simple rigid monoidal category ${\cal{C}}$.
Then $Irr({\cal{M}})$ is a NIM-representation of the fusion rule $Irr({\cal{C}})$.
Note that the rigidity property for NIM-representations follows from the adjunction
$${\cal{M}} (X\ast M, N) \simeq {\cal{M}} (M, X^{\vee}\ast N).$$
\section{Fibonacci categories}\label{secfib}
In this section we describe modular categories with the {\em Fibonacci} fusion rule
$$\textrm{$\fibfrmod$} = \{1,x\} : x^2=1+x.$$
That is we classify all possible associativity constraints (F-matrices) and braidings (B-matrices). Although not written explicitly anywhere the results are known to specialists. We present them here for the sake of completeness.
We consider a semi-simple $k$-linear category $\mathcal{F}ib$ with simple objects $I$ and $X$.
The tensor product is defined by
$$X\otimes X = I \oplus X\, .$$
Under this definition there are two fundamental hom-spaces $\mathcal{F}ib(X^{2},X)$ and $\mathcal{F}ib(X^{2},I)$ for which the two respective basis vectors are
\begin{equation*}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+{\copy\Treetwo} [rr]*+{\copy\Treetwoid}
}
\end{equation*}
\subsection{Associativity}\label{fibassoc}
The only non-trivial component of the associativity constraint for $\mathcal{F}ib$ is $$\alpha_{X,X,X}:(X\otimes X)\otimes X\to X\otimes(X\otimes X).$$
On the level of hom-spaces this corresponds to two isomorphisms
$$\mathcal{F}ib(\alpha_{X,X,X},I):\mathcal{F}ib(X\otimes(X\otimes X),I)\to\mathcal{F}ib((X\otimes X)\otimes X,I)$$
and
$$\mathcal{F}ib(\alpha_{X,X,X},X):\mathcal{F}ib(X\otimes(X\otimes X),X)\to\mathcal{F}ib((X\otimes X)\otimes X,X).$$
Clearly $dim(\mathcal{F}ib(X^{3},I))=1$ and $dim(\mathcal{F}ib(X^{3},X))=2$ and so the associativity $k$-linear transformation for $\mathcal{F}ib(X^{3},X)$ is given by $A\in GL_2(k)$ and by $\alpha\in k^{\ast}$ for $\mathcal{F}ib(X^{3},I)$.
Graphically,
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \Treethreeid}="ta" ( [rrr]*+<0pt,6pt>{\copy \Treethreeidrev}="tb" !{+L*!R\txt\small{\phantom{aa}$\alpha$\phantom{aa}}."tb"="tb"} )
("ta" :@{|->} "tb")
}
\end{equation*}
\begin{equation}\label{assco}
\begin{array}{ccc}
\left[\begin{array}{c}
\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \TreethreeA}
[d]*+<0pt,6pt>{\copy \TreethreeB}
}
\end{array}\right]
&
\xymatrix{\ar@{|->}[rr]&&\quad A}
&
\left[\begin{array}{c}
\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \TreethreeArev}
[d]*+<0pt,6pt>{\copy \TreethreeBrev}
}
\end{array}\right]
\end{array}
\end{equation}
\noindent where
$$A = \left(\begin{array}{ccc} a & b \\ c & d \end{array}\right)\, .$$
The pentagon axiom of associativity coherence gives a set of equations (on $\alpha$ and matrix elements of $A$) for both $\mathcal{F}ib(X^{4},I)$ and $\mathcal{F}ib(X^{4},X)$.
The $k$-vector space $\mathcal{F}ib(X^{4},I)$ is two-dimensional and so there are two groups of equations, one for each choice of a basis tree:
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \TreefouridAtopa}="t1"(
:@{|->}[d]*+<0pt,6pt>{\copy \TreefouridAbota}="t1a" ( !{+L*!R\txt\scriptsize{$a$}."t1a"="t1a"},
[d]*+<0pt,6pt>{\copy \TreefouridAbotaid}="t1b" (:@{}[u]|{+})
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridAtopb}="t2" (!{+L*!R+\txt\scriptsize{$\alpha$}."t2"="tt2"} ("t1" :@{|->} "tt2") ,
[d]*+<0pt,6pt>{\copy \TreefouridAbotb}="t2a" ( !{+L*!R\txt\scriptsize{$\phantom{a.}\alpha a$}."t2a"="tt2a"},
[d]*+<0pt,6pt>{\copy \TreefouridAbotbid}="t2b" (:@{}[u]|{+}, !{+L*!R\txt\scriptsize{$b$}."t2b"="t2b"})
("t1a" :@{|->} "tt2a")
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridAtopc}="t3" (!{+L*!R+\txt\scriptsize{$\alpha^{2}$}."t3"="t3"} ("t2" :@{|->} "t3"),
:@{=}[d]*+<0pt,6pt>{\copy \TreefouridAtopc}="t3a" ( !{+L*!R\txt\scriptsize{$\phantom{a.}\alpha a^{2} + bc$}."t3a"="tt3a"},
[d]*+<0pt,6pt>{\copy \TreefouridAtopcid}="t3b" (:@{}[u]|{+}, !{+L*!R\txt\scriptsize{$\alpha ab +bd$}."t3b"="t3b"})
("t2a" :@{|->} "tt3a")
)
)
}
\end{equation*}
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \TreefouridAtopaid}="t1"(
:@{|->}[d]*+<0pt,6pt>{\copy \TreefouridAbota}="t1a" ( !{+L*!R\txt\scriptsize{$c$}."t1a"="t1a"},
[d]*+<0pt,6pt>{\copy \TreefouridAbotaid}="t1b" (:@{}[u]|{+}, !{+L*!R\txt\scriptsize{$d$}."t1b"="t1b"})
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridAtopbid}="t2" (!{+L*!R+\txt\scriptsize{$$}."t2"="tt2"} ("t1" :@{|->} "tt2") ,
[d]*+<0pt,6pt>{\copy \TreefouridAbotb}="t2a" ( !{+L*!R\txt\scriptsize{$\phantom{a.}\alpha c$}."t2a"="t2a"},
[d]*+<0pt,6pt>{\copy \TreefouridAbotbid}="t2b" (:@{}[u]|{+}="p2" , !{+L*!R\txt\scriptsize{$d$}."t2b"="t2b"})
("t1a" :@{|->} "t2a")
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridAtopcid}="t3" (!{+L*!R+\txt\scriptsize{$$}."t3"="t3"} ("t2" :@{|->} "t3"),
:@{=}[d]*+<0pt,6pt>{\copy \TreefouridAtopc}="t3a" ( !{+L*!R\txt\scriptsize{$\phantom{a.}\alpha ca + dc$}."t3a"="t3a"},
[d]*+<0pt,6pt>{\copy \TreefouridAtopcid}="t3b" (:@{}[u]|{+}, !{+L*!R\txt\scriptsize{$\alpha cb + d^{2}$}."t3b"="t3b"})
("t2a" :@{|->} "t3a")
)
)
}
\end{equation*}
\noindent which yields four equations
\begin{eqnarray*}
\alpha a^{2} + bc &=& \alpha^{2} \\
\alpha ab + bd &=& 0 \\
\alpha cb + d^{2} &=& 1 \\
\alpha ca + dc &=& 0
\end{eqnarray*}
There are then three calculations for $\mathcal{F}ib(X^{4},X)$ each corresponding to a different initial basis-tree
\begin{equation*}\scalebox{0.85}{\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \TreefouridBtopa}="t1a"(
:@{|->}[ddd]*+<0pt,6pt>{\copy \TreefouridBbota}="t1b" (!{+L*!R+\txt\scriptsize{$a$}}) (
[d]*+<0pt,6pt>{\copy \TreefouridBbotaid}="t1c" (!{+L*!R+\txt\scriptsize{$b$}}) :@{}[u]|{+}
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridBtopb}="t2a" (!{+L*!R+\txt\scriptsize{$a$}."t2a"="tt2a"} ("t1a" :@{|->} "tt2a") ) (
"t2a"[d]*+<0pt,6pt>{\copy \TreefouridBtopbid}="t2b" (!{+L*!R+\txt\scriptsize{$b$}}) :@{}[u]|{+} (
"t2a"[ddd]*+<0pt,6pt>{\copy \TreefouridBbotb}="t2c" (!{+L*!R+\txt\scriptsize{$a^{2}$}="t2cl"}) (
"t2a"[dddd]*+<0pt,6pt>{\copy \TreefouridBbotbid}="t2d" (!{+L*!R+\txt\scriptsize{$b$}}) :@{}[u]|{+} !{"t2c"."t2cl"="tt2c"} ("t1b" :@{|->} "tt2c") (
"t2a"[ddddd]*+<0pt,6pt>{\copy \TreefouridBbotbidid}="t2e" (!{+L*!R+\txt\scriptsize{$ab$}}) :@{}[u]|{+}
)
)
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridBtopc}="t3a" (!{+L*!R+\txt\scriptsize{$a^{2}$}."t3a"="tt3a"}) ("tt2a" :@{|->} "tt3a")
"t3a"[d]*+<0pt,6pt>{\copy \TreefouridBtopcid}="t3b" (!{+L*!R+\txt\scriptsize{$b$}}) :@{}[u]|{+}
"t3a"[dd]*+<0pt,6pt>{\copy \TreefouridBtopcidb}="t3c" (!{+L*!R+\txt\scriptsize{$ab$}}) :@{}[u]|{+}
"t3a"[ddd]*+<0pt,6pt>{\copy \TreefouridBtopc}="t3d" ("t3d" :@{=} "t3c") (!{+L*!R+\txt\scriptsize{$a^{3} + cb$}."t3d"="t3dl"}) ("tt2c" :@{|->} "t3dl")
"t3a"[dddd]*+<0pt,6pt>{\copy \TreefouridBtopcid}="t3e" (!{+L*!R+\txt\scriptsize{$a^{2}b +bd$}}) :@{}[u]|{+}
"t3a"[ddddd]*+<0pt,6pt>{\copy \TreefouridBtopcidb}="t3f" (!{+L*!R+\txt\scriptsize{$\alpha ab$}}) :@{}[u]|{+}
}
}\end{equation*}
and,
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \TreefouridBtopaid}="t1a"(
:@{|->}[ddd]*+<0pt,6pt>{\copy \TreefouridBbota}="t1b" (!{+L*!R+\txt\scriptsize{$c$}}) (
[d]*+<0pt,6pt>{\copy \TreefouridBbotaid}="t1c" (!{+L*!R+\txt\scriptsize{$d$}}) :@{}[u]|{+}
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridBtopbidid}="t2a" (!{+L*!R+\txt\scriptsize{$$}."t2a"="tt2a"} ("t1a" :@{|->} "tt2a") ) (
"t2a"[ddd]*+<0pt,6pt>{\copy \TreefouridBbotb}="t2c" (!{+L*!R+\txt\scriptsize{$ca$}="t2cl"}) (
"t2a"[dddd]*+<0pt,6pt>{\copy \TreefouridBbotbid}="t2d" (!{+L*!R+\txt\scriptsize{$d$}}) :@{}[u]|{+} !{"t2c"."t2cl"="tt2c"} ("t1b" :@{|->} "tt2c") (
"t2a"[ddddd]*+<0pt,6pt>{\copy \TreefouridBbotbidid}="t2e" (!{+L*!R+\txt\scriptsize{$cb$}}) :@{}[u]|{+}
)
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridBtopc}="t3a" (!{+L*!R+\txt\scriptsize{$c$}."t3a"="tt3a"}) ("tt2a" :@{|->} "tt3a") (
"t3a"[d]*+<0pt,6pt>{\copy \TreefouridBtopcidb}="t3b" (!{+L*!R+\txt\scriptsize{$d$}}) :@{}[u]|{+} (
"t3a"[ddd]*+<0pt,6pt>{\copy \TreefouridBtopc}="t3d" ("t3d" :@{=} "t3b") (!{+L*!R+\txt\scriptsize{$ca^{2} + cd$}."t3d"="t3dl"}) ("tt2c" :@{|->} "t3dl") (
"t3a"[dddd]*+<0pt,6pt>{\copy \TreefouridBtopcid}="t3e" (!{+L*!R+\txt\scriptsize{$acb + d^{2}$}}) :@{}[u]|{+} (
"t3a"[ddddd]*+<0pt,6pt>{\copy \TreefouridBtopcidb}="t3f" (!{+L*!R+\txt\scriptsize{$\alpha cb$}}) :@{}[u]|{+}
)
)
)
)
}
\end{equation*}
and,
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+<0pt,6pt>{\copy \TreefouridBtopaidid}="t1a"(
:@{|->}[ddd]*+<0pt,6pt>{\copy \TreefouridBbotaid}="t1b" (!{+L*!R+\txt\scriptsize{$\alpha$}})
)
[rr]*+<0pt,6pt>{\copy \TreefouridBtopb}="t2a" (!{+L*!R+\txt\scriptsize{$c$}."t2a"="tt2a"} ("t1a" :@{|->} "tt2a") ) (
"t2a"[d]*+<0pt,6pt>{\copy \TreefouridBtopbid}="t2b" (!{+L*!R+\txt\scriptsize{$d$}}) :@{}[u]|{+} (
"t2a"[ddd]*+<0pt,6pt>{\copy \TreefouridBbotb}="t2c" (!{+L*!R+\txt\scriptsize{$\alpha c$}="t2cl"}) (
"t2a"[dddd]*+<0pt,6pt>{\copy \TreefouridBbotbidid}="t2d" (!{+L*!R+\txt\scriptsize{$\alpha d$}}) :@{}[u]|{+} !{"t2c"."t2cl"="tt2c"} ("t1b" :@{|->} "tt2c")
)
)
)
[rr]*+<0pt,6pt>{\copy \TreefouridBtopc}="t3a" (!{+L*!R+\txt\scriptsize{$ca$}."t3a"="tt3a"}) ("tt2a" :@{|->} "tt3a")
"t3a"[d]*+<0pt,6pt>{\copy \TreefouridBtopcid}="t3b" (!{+L*!R+\txt\scriptsize{$cb$}}) :@{}[u]|{+}
"t3a"[dd]*+<0pt,6pt>{\copy \TreefouridBtopcidb}="t3c" (!{+L*!R+\txt\scriptsize{$d$}}) :@{}[u]|{+}
"t3a"[ddd]*+<0pt,6pt>{\copy \TreefouridBtopc}="t3d" ("t3d" :@{=} "t3c") (!{+L*!R+\txt\scriptsize{$\alpha ca$}."t3d"="t3dl"}) ("tt2c" :@{|->} "t3dl")
"t3a"[dddd]*+<0pt,6pt>{\copy \TreefouridBtopcid}="t3e" (!{+L*!R+\txt\scriptsize{$\alpha cb$}}) :@{}[u]|{+}
"t3a"[ddddd]*+<0pt,6pt>{\copy \TreefouridBtopcidb}="t3f" (!{+L*!R+\txt\scriptsize{$\alpha^{2} d$}}) :@{}[u]|{+}
}
\end{equation*}
\noindent These diagrams provide a further eight equations
\begin{eqnarray*}
a^{3}+bc &=& a^{2} \\
a^{2}b+bd &=& b \\
ca^{2}+cd &=& c \\
abc+d^{2} &=& 0 \\
\alpha ab &=& ab \\
\alpha cb &=& d \\
\alpha ca &=& ca \\
\alpha^{2}d &=& cb.
\end{eqnarray*}
Solving all the equations proves the following lemma.
\begin{lemma}\label{assoclemma} The associativity constraint for the Fibonacci category is given by (\ref{assco}), where
$$\alpha = 1,\quad\quad{A=\left(\begin{array}{cc} a & b \\ -ab^{-1} & -a\end{array}\right)}$$
with $a$ being a solution of $a^2=a+1$.
\end{lemma}
Note that $det(A)=-1$ and ${A^{-1}=A}$.
\subsection{Braiding}
The only non-trivial component of a braiding on $\mathcal{F}ib$ is the isomorphism $c_{X,X}: X^{2}\to X^{2}$.
On the level of hom-spaces $\mathcal{F}ib (X^{2},1)$ and $\mathcal{F}ib (X^{2},X)$ this isomorphism is given by
\begin{equation}\label{braid}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*+{\copy\Treetwo}="t1" [rr]*+{\copy\Treetwo}="t2" (!{+L*!R+\txt\scriptsize{$u$}."t2"="t2"}) ("t1" :@{|->} "t2")
[rr]*+{\copy\Treetwoid}="t3" [rr]*+{\copy\Treetwoid}="t4" (!{+L*!R+\txt\scriptsize{$w$}."t4"="t4"}) ("t3" :@{|->} "t4") ,
}
\end{equation}
where $w,u\in k^{\ast}$.
The coherence condition gives the following calculations
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy \Treethreeidrevs}="t1"
"t1"[urr]*+<0pt,6pt>{\copy \Treethreeids}="t2" (!{+L*!R+\txt\scriptsize{$$}."t2"="t2"}) ("t1" :@{|->} "t2")
"t1"[urrrr]*+<0pt,6pt>{\copy \Treethreeidrevs}="t3" (!{+L*!R+\txt\scriptsize{$w$}."t3"="t3"}) ("t2" :@{|->} "t3")
"t1"[urrrrrr]*+<0pt,6pt>{\copy \Treethreeids}="t4" (!{+L*!R+\txt\scriptsize{$w$}."t4"="t4"}) ("t3" :@{|->} "t4")
"t1"[drr]*+<0pt,6pt>{\copy \Treethreeidrevs}="t5" (!{+L*!R+\txt\scriptsize{$u$}."t5"="t5"}) ("t1" :@{|->} "t5")
"t1"[drrrr]*+<0pt,6pt>{\copy \Treethreeids}="t6" (!{+L*!R+\txt\scriptsize{$u$}."t6"="t6"}) ("t5" :@{|->} "t6")
"t1"[drrrrrr]*+<0pt,6pt>{\copy \Treethreeids}="t7" (!{+L*!R+\txt\scriptsize{$u^{2}$}."t7"="t7"}) ("t6" :@{|->} "t7")
("t4" :@{=} "t7")
}
\end{equation*}
\noindent and,
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy \TreethreeArevs}="t1"
"t1"[uurr]*+<0pt,6pt>{\copy \TreethreeAs}="t2" (!{+L*!R+\txt\scriptsize{$a$}="tt2"}) (
[d]*+<0pt,6pt>{\copy \TreethreeBs}="t2a" ("t2a" :@{}|{+} "t2") (!{"t2"."tt2"="t2"}) (!{+L*!R+\txt\scriptsize{$b$}."t2a"="t2a"}),
[ddd]*+<0pt,6pt>{\copy \TreethreeArevs}="t2b" (!{+L*!R+\txt\scriptsize{$u$}."t2b"="t2b"}),
("t1" :@{|->} "t2") ("t1" :@{|->} "t2b")
)
"t1"[uurrrr]*+<0pt,6pt>{\copy \TreethreeArevs}="t3" (!{+L*!R+\txt\scriptsize{$ua$}="tt3"}) (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t3a" ("t3a" :@{}|{+} "t3") (!{"t3"."tt3"="t3"}) (!{+L*!R+\txt\scriptsize{$b$}."t3a"="t3a"}),
[ddd]*+<0pt,6pt>{\copy \TreethreeAs}="t3b" (!{+L*!R+\txt\scriptsize{$ua$}="tt3b"}),
[dddd]*+<0pt,6pt>{\copy \TreethreeBs}="t3c" ("t3c" :@{}|{+} "t3b") (!{"t3b"."tt3b"="t3b"}) (!{+L*!R+\txt\scriptsize{$ub$}."t3c"="t3c"})
("t2" :@{|->} "t3") ("t2b" :@{|->} "t3b")
)
"t1"[uurrrrrr]*+<0pt,6pt>{\copy \TreethreeAs}="t4" (!{+L*!R+\txt\scriptsize{$ua^{2}+bc$}="tt4"}) (
[d]*+<0pt,6pt>{\copy \TreethreeBs}="t4a" ("t4a" :@{}|{+} "t4") (!{"t4"."tt4"="t4"}) (!{+L*!R+\txt\scriptsize{$uab-ab$}="tt4a"}),
[ddd]*+<0pt,6pt>{\copy \TreethreeAs}="t4b" ("t4a" :@{=} "t4b") (!{"t4a"."tt4a"="t4a"}) (!{+L*!R+\txt\scriptsize{$u^{2}a$}="tt4b"}),
[dddd]*+<0pt,6pt>{\copy \TreethreeBs}="t4c" ("t4c" :@{}|{+} "t4b") (!{"t4b"."tt4b"="t4b"}) (!{+L*!R+\txt\scriptsize{$wub$}."t4c"="t4c"})
("t3" :@{|->} "t4") ("t3b" :@{|->} "t4b")
)
}
\end{equation*}
\noindent and,
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy \TreethreeBrevs}="t1"
"t1"[uurr]*+<0pt,6pt>{\copy \TreethreeAs}="t2" (!{+L*!R+\txt\scriptsize{$c$}="tt2"}) (
[d]*+<0pt,6pt>{\copy \TreethreeBs}="t2a" ("t2a" :@{}|{+} "t2") (!{"t2"."tt2"="t2"}) (!{+L*!R+\txt\scriptsize{$-a$}."t2a"="t2a"}),
[ddd]*+<0pt,6pt>{\copy \TreethreeBrevs}="t2b" (!{+L*!R+\txt\scriptsize{$w$}."t2b"="t2b"}),
("t1" :@{|->} "t2") ("t1" :@{|->} "t2b")
)
"t1"[uurrrr]*+<0pt,6pt>{\copy \TreethreeArevs}="t3" (!{+L*!R+\txt\scriptsize{$cu$}="tt3"}) (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t3a" ("t3a" :@{}|{+} "t3") (!{"t3"."tt3"="t3"}) (!{+L*!R+\txt\scriptsize{$-a$}."t3a"="t3a"}),
[ddd]*+<0pt,6pt>{\copy \TreethreeAs}="t3b" (!{+L*!R+\txt\scriptsize{$wc$}="tt3b"}),
[dddd]*+<0pt,6pt>{\copy \TreethreeBs}="t3c" ("t3c" :@{}|{+} "t3b") (!{"t3b"."tt3b"="t3b"}) (!{+L*!R+\txt\scriptsize{$-wa$}."t3c"="t3c"})
("t2" :@{|->} "t3") ("t2b" :@{|->} "t3b")
)
"t1"[uurrrrrr]*+<0pt,6pt>{\copy \TreethreeAs}="t4" (!{+L*!R+\txt\scriptsize{$cua-ac$}="tt4"}) (
[d]*+<0pt,6pt>{\copy \TreethreeBs}="t4a" ("t4a" :@{}|{+} "t4") (!{"t4"."tt4"="t4"}) (!{+L*!R+\txt\scriptsize{$cub+a^{2}$}="tt4a"}),
[ddd]*+<0pt,6pt>{\copy \TreethreeAs}="t4b" ("t4a" :@{=} "t4b") (!{"t4a"."tt4a"="t4a"}) (!{+L*!R+\txt\scriptsize{$uwc$}="tt4b"}),
[dddd]*+<0pt,6pt>{\copy \TreethreeBs}="t4c" ("t4c" :@{}|{+} "t4b") (!{"t4b"."tt4b"="t4b"}) (!{+L*!R+\txt\scriptsize{$-w^{2}a$}."t4c"="t4c"})
("t3" :@{|->} "t4") ("t3b" :@{|->} "t4b")
)
}
\end{equation*}
\noindent We obtain the following equations
\begin{eqnarray*}
u^{2} &=& w \\
u^{2}a &=& ua^{2} - a \\
wu &=& ua -a \\
-w^{2} &=& a- u
\end{eqnarray*}
\noindent Solving these proves the following lemma.
\begin{lemma}\label{braidlemma} A braiding (\ref{braid}) on a Fibonacci category is completely determined by $u, w=u^2$ such that
$${u^{2} = ua -1}$$
\end{lemma}
\begin{remark} Note that $a = u + u^{-1}$ together with $a^{2} - a = 1$ implies that $u$ is a primitive root of unity of order $10$.
Indeed, replacing $a$ by $u+u^{-1}$ in $a^2 = 1+a$ we get
$$u + u^{-1} +1 = (u + u^{-1})^2 = u^2 + 2 + u^{-2}$$
or
$$0 = u^2 - u + 1 - u^{-1} + u^{-2} = u^{-2}(u^4 - u^3 + u^2 - u + 1).$$
Thus the field of definition of a braided Fibonacci category is the cyclotomic field $\mathbb{Q}(\sqrt[10]{1})=\mathbb{Q}(\sqrt[5]{1})$.
\end{remark}
\subsection{Twist}
For both simple objects $X$ and $I$ the space of endomorphisms is one dimensional (generated by the identity morphism) and thus the twist automorphisms of $I$ and $X$ are simply scalar multiples of the identities on that object:
\begin{eqnarray*} \theta_{1}&=&\operatorname{id}_{I} \\ \theta_{X}&=&\rho\cdot\operatorname{id}_{X} \end{eqnarray*}
where $\rho\in k^{*}$ and naturality demands the scalar for $\theta_{I}$ to be $1$ (see \cite{tu}).
\noindent Using the coherence axiom for twists on the basis trees of $\mathcal{F}ib (X^{2},X)$ and $\mathcal{F}ib (X^{2},I)$ respectively gives
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy\Treetwo}="t1" (
[d]*+<0pt,6pt>{\copy\Treetwo}="t2" (!{+L*!R+\txt\scriptsize{$u$}}) ("t1" :@{|->} "t2")
)
"t1"[rrd]*+<0pt,6pt>{\copy\Treetwo}="t3" (!{+L*!R+\txt\scriptsize{$u^{2}$}="tt3"}) (!{"t3"."tt3"="t3"}) ("t2" :@{|->} "t3")
"t1"[rrrr]*+<0pt,6pt>{\copy\Treetwo}="t4" (!{+L*!R+\txt\scriptsize{$\rho$}="tt4"}) (!{"t4"."tt4"="t4"}) ("t1" :@{|->} "t4") (
[d]*+<0pt,6pt>{\copy\Treetwo}="t5" (!{+L*!R+\txt\scriptsize{$\rho^{2}u^{2}$}="tt5"}) (!{"t5"."tt5"="t5"}) ("t3" :@{|->} "t5") ("t4" :@{=} "t5")
)
}
\end{equation*}
\noindent and
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy\Treetwoid}="t1" (
[d]*+<0pt,6pt>{\copy\Treetwoid}="t2" (!{+L*!R+\txt\scriptsize{$u^{2}$}}) ("t1" :@{|->} "t2")
)
"t1"[rrd]*+<0pt,6pt>{\copy\Treetwoid}="t3" (!{+L*!R+\txt\scriptsize{$u^{4}$}="tt3"}) (!{"t3"."tt3"="t3"}) ("t2" :@{|->} "t3")
"t1"[rrrr]*+<0pt,6pt>{\copy\Treetwoid}="t4" (!{+L*!R+\txt\scriptsize{$1$}="tt4"}) (!{"t4"."tt4"="t4"}) ("t1" :@{|->} "t4") (
[d]*+<0pt,6pt>{\copy\Treetwoid}="t5" (!{+L*!R+\txt\scriptsize{$u^{4}$}="tt5"}) (!{"t5"."tt5"="t5"}) ("t3" :@{|->} "t5") ("t4" :@{=} "t5")
)
}
\end{equation*}
\begin{lemma} Twist structures on $\mathcal{F}ib$ are completely determined by the braiding as
$${\rho = u^{-2}}.$$
\end{lemma}
\subsection{Monoidal Equivalences}
Let $\mathcal{F}ib_{A}$ denote the monoidal category $\mathcal{F}ib$ with the associativity matrix $A$ as first prescribed in \ref{fibassoc} with entries $a,b,c,d\in k^{\ast}$ and $\alpha = 1$.
Suppose there is another associativity matrix for $\mathcal{F}ib$ written
$$A'=\left(\begin{array}{cc} a' & b' \\ c' & d'\end{array}\right)$$
such that there is a monoidal equivalence $F: \mathcal{F}ib_{A} \to \mathcal{F}ib_{A'}$.
Due to the semi-simple nature of $\mathcal{F}ib$ the only possibility for the underlying endo-functor of this equivalence is the identity (on objects).
Thus the monoidal functor is completely determined by the automorphisms on $\mathcal{F}ib(X^{2},I)$ and $\mathcal{F}ib(X^{2},X)$ given by
\begin{equation}\label{moneq}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy\Treetwoid}="t1" (
[d]*+<0pt,6pt>{\copy\Treetwo}="t3"
)
"t1"[rr]*+<0pt,6pt>{\copy\Treetwoid}="t2" (!{+L*!R+\txt\scriptsize{$f$}="tt2"}) (!{"t2"."tt2"="t2"}) ("t1" :@{|->} "t2") (
[d]*+<0pt,6pt>{\copy\Treetwo}="t4" (!{+L*!R+\txt\scriptsize{$g$}="tt4"}) (!{"t4"."tt4"="t4"}) ("t3" :@{|->} "t4")
)
}
\end{equation}
\noindent respectively, where $f,g\in k^{*}$.
\noindent The coherence condition for the monoidal equivalence gives the following calculations
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy \Treethreeids}="t1"
"t1"[urr]*+<0pt,6pt>{\copy \Treethreeidrevs}="t2" (!{+L*!R+\txt\scriptsize{$\phantom{f}$}="tt2"}) (!{"tt2"."t2"="t2"}) ("t1" :@{|->} "t2") (
[dd]*+<0pt,6pt>{\copy \Treethreeids}="t3" (!{+L*!R+\txt\scriptsize{$f$}="tt3"}) (!{"tt3"."t3"="t3"}) ("t1" :@{|->} "t3")
)
"t1"[urrrr]*+<0pt,6pt>{\copy \Treethreeidrevs}="t4" (!{+L*!R+\txt\scriptsize{$f$}="tt4"}) (!{"tt4"."t4"="t4"}) ("t2" :@{|->} "t4") (
[dd]*+<0pt,6pt>{\copy \Treethreeids}="t5" (!{+L*!R+\txt\scriptsize{$fg$}="tt5"}) (!{"tt5"."t5"="t5"}) ("t3" :@{|->} "t5")
)
"t1"[urrrrrr]*+<0pt,6pt>{\copy \Treethreeidrevs}="t6" (!{+L*!R+\txt\scriptsize{$fg$}="tt6"}) (!{"t6"."tt6"="t6"}) ("t4" :@{|->} "t6") (
[dd]*+<0pt,6pt>{\copy \Treethreeidrevs}="t7" (!{+L*!R+\txt\scriptsize{$fg$}="tt7"}) (!{"t7"."tt7"="t7"}) ("t5" :@{|->} "t7") ("t7" :@{=} "t6")
)
}
\end{equation*}
\noindent and
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy \TreethreeAs}="t1"
"t1"[urr]*+<0pt,6pt>{\copy \TreethreeArevs}="t2" (!{+L*!R+\txt\scriptsize{$a$}="tt2"}) (!{"t2"."tt2"="t2"}) ("t1" :@{|->} "t2") (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t3" (!{+L*!R+\txt\scriptsize{$b$}="tt3"}) (!{"t3"."tt3"="t3"}) ("t3" :@{}|{+} "t2")
[d]*+<0pt,6pt>{\copy \TreethreeAs}="t4" (!{+L*!R+\txt\scriptsize{$g$}="tt4"}) (!{"t4"."tt4"="t4"}) ("t1" :@{|->} "t4")
)
"t1"[urrrr]*+<0pt,6pt>{\copy \TreethreeArevs}="t5" (!{+L*!R+\txt\scriptsize{$ga$}="tt5"}) (!{"t5"."tt5"="t5"}) ("t2" :@{|->} "t5") (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t6" (!{+L*!R+\txt\scriptsize{$b$}="tt6"}) (!{"t6"."tt6"="t6"}) ("t6" :@{}|{+} "t5")
[d]*+<0pt,6pt>{\copy \TreethreeAs}="t7" (!{+L*!R+\txt\scriptsize{$g^{2}$}="tt7"}) (!{"t7"."tt7"="t7"}) ("t4" :@{|->} "t7")
)
"t1"[urrrrrr]*+<0pt,6pt>{\copy \TreethreeArevs}="t8" (!{+L*!R+\txt\scriptsize{$ag^{2}$}="tt8"}) (!{"t8"."tt8"="t8"}) ("t5" :@{|->} "t8") (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t9" (!{+L*!R+\txt\scriptsize{$bf$}="tt9"}) (!{"t9"."tt9"="t9"}) ("t9" :@{}|{+} "t8")
[d]*+<0pt,6pt>{\copy \TreethreeArevs}="t10" (!{+L*!R+\txt\scriptsize{$g^{2}a'$}="tt10"}) (!{"t10"."tt10"="t10"}) ("t7" :@{|->} "t10") ("t10" :@{=} "t9")
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t11" (!{+L*!R+\txt\scriptsize{$g^{2}b'$}="tt11"}) (!{"t11"."tt11"="t11"}) ("t11" :@{}|{+} "t10")
)
}
\end{equation*}
\noindent and
\begin{equation*}\def6pt{6pt}
\renewcommand{\objectstyle}{\labelstyle}
\xygraph{ !{0;/r4.0pc/:;/u4.0pc/::}
[]*++<0pt,6pt>{\copy \TreethreeBs}="t1"
"t1"[urr]*+<0pt,6pt>{\copy \TreethreeArevs}="t2" (!{+L*!R+\txt\scriptsize{$c$}="tt2"}) (!{"t2"."tt2"="t2"}) ("t1" :@{|->} "t2") (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t3" (!{+L*!R+\txt\scriptsize{$d$}="tt3"}) (!{"t3"."tt3"="t3"}) ("t3" :@{}|{+} "t2")
[d]*+<0pt,6pt>{\copy \TreethreeBs}="t4" (!{+L*!R+\txt\scriptsize{$$}="tt4"}) (!{"t4"."tt4"="t4"}) ("t1" :@{|->} "t4")
)
"t1"[urrrr]*+<0pt,6pt>{\copy \TreethreeArevs}="t5" (!{+L*!R+\txt\scriptsize{$cg$}="tt5"}) (!{"t5"."tt5"="t5"}) ("t2" :@{|->} "t5") (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t6" (!{+L*!R+\txt\scriptsize{$d$}="tt6"}) (!{"t6"."tt6"="t6"}) ("t6" :@{}|{+} "t5")
[d]*+<0pt,6pt>{\copy \TreethreeBs}="t7" (!{+L*!R+\txt\scriptsize{$f$}="tt7"}) (!{"t7"."tt7"="t7"}) ("t4" :@{|->} "t7")
)
"t1"[urrrrrr]*+<0pt,6pt>{\copy \TreethreeArevs}="t8" (!{+L*!R+\txt\scriptsize{$cg^{2}$}="tt8"}) (!{"t8"."tt8"="t8"}) ("t5" :@{|->} "t8") (
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t9" (!{+L*!R+\txt\scriptsize{$df$}="tt9"}) (!{"t9"."tt9"="t9"}) ("t9" :@{}|{+} "t8")
[d]*+<0pt,6pt>{\copy \TreethreeArevs}="t10" (!{+L*!R+\txt\scriptsize{$fc'$}="tt10"}) (!{"t10"."tt10"="t10"}) ("t7" :@{|->} "t10") ("t10" :@{=} "t9")
[d]*+<0pt,6pt>{\copy \TreethreeBrevs}="t11" (!{+L*!R+\txt\scriptsize{$fd'$}="tt11"}) (!{"t11"."tt11"="t11"}) ("t11" :@{}|{+} "t10")
)
}
\end{equation*}
\noindent Put
$$G=\left(\begin{array}{cc} f & 0 \\ 0 & g^{2}\end{array}\right).$$
Monoidal coherence equations are equivalent to the matrix conjugation equation
$$A' = G^{-1} A G\,\, .$$
\begin{proposition}
Up to monoidal equivalence associativity constraints
\ref{assco} for Fibonacci category correspond to solutions of the equation $a^2 = a+1$: $$\alpha = 1,\quad {A=\left(\begin{array}{cc} a & 1 \\ -a & -a \end{array}\right)}.$$
For each associativity constraint on $\mathcal{F}ib$ braided (balanced) structures up
to braided equivalence correspond to solutions of $u^2 = au - 1$.
\end{proposition}
\begin{proof}
For any $A$ as found in lemma \ref{assoclemma} Choose $f$ and $g$ such that $f^{-1}g^{2} = b$ so that
$$G^{-1}AG=\left(\begin{array}{cc} f^{-1} & 0 \\ 0 & g^{-2}\end{array}\right) \left(\begin{array}{cc} a & b \\ -ab^{-1} & -a \end{array}\right) \left(\begin{array}{cc} f & 0 \\ 0 & g^{2}\end{array}\right) = \left(\begin{array}{cc} a & 1 \\ -a & -a \end{array}\right)\, . $$
That there are exactly two braided structures is given by lemma \ref{braidlemma}. Clearly monoidal equivalences (\ref{moneq}) can not identify different braided (balanced) structures.
\end{proof}
\begin{remark}
We write $\mathcal{F}ib_{u}$ for $\mathcal{F}ib$ with a particular choice of parameterizing $u$ (and $a = u + u^{-1}$).
\end{remark}
\subsection{Duality and dimensions}
Here we show that Fibonacci categories are rigid; that is, any object has a dual, and that they are modular.
All we need to check is that $X$ has a dual.
From the fusion rule it is clear that if $X^*$ exists it must be $X$.
Thus, all we need to construct is the evaluation and coevaluation maps
$$ev:X\otimes X\to I,\quad coev:I\to X\otimes X,$$
such that the compositions
$$\xymatrix{X\ar[r]^-{1\otimes coev} & X\otimes(X\otimes X) \ar[rr]^-{\alpha_{X,X,X}^{-1}} && (X\otimes X)\otimes X \ar[r]^-{ev\otimes 1} & X}$$
\begin{equation}\label{dual}
\xymatrix{X\ar[r]^-{coev\otimes 1} & (X\otimes X)\otimes X \ar[rr]^-{\alpha_{X,X,X}} && X\otimes (X\otimes X) \ar[r]^-{1\otimes ev} & X}
\end{equation}
are identity.
Since such evaluation and coevaluation maps are unique up to a constant we can assume that $ev$ is the basic element in $\mathcal{F}ib(X^2,I)$.
To describe $coev$ note that the composition in $\mathcal{F}ib$ gives a non-degenerate pairing
$$\mathcal{F}ib(I,X^2)\otimes\mathcal{F}ib(X^2,I)\to \mathcal{F}ib(I,I) = k.$$
We can choose a basic element in $\mathcal{F}ib(I,X^2)$ to be the dual to the basic element in $\mathcal{F}ib(X^2,I)$ as in \ref{fibassoc}.
Then $coev$ is proportional to the basic element with the coefficient $\gamma$.
The compositions (\ref{dual}) are both equal to $-a\gamma$ and so
$$\gamma = -a^{-1} = 1-a.$$
\begin{lemma}
The dimension $dim(X)$ of $X$ is $1-a$.
\end{lemma}
\begin{proof}
The composition
$$\xymatrix{I\ar[r]^-{coev} & X\otimes X \ar[rr]^-{(\theta_{X}\otimes 1)c_{X,X}} && X\otimes X \ar[r]^-{ev} & I}$$
coincides with $(1-a)1_X$.
\end{proof}
\begin{corollary}
The categorical dimension $Dim(\mathcal{F}ib_{u})$ of the ribbon category $\mathcal{F}ib_{u}$ is $3 - a$.
The square of the multiplicative central charge $\xi(\mathcal{F}ib_{u})^2$ is $u^{-1}$.
\end{corollary}
\begin{proof}
For the categorical dimension
$$Dim(\mathcal{F}ib_{u}) = dim(I)^2 + dim(X)^2 = 1 + (1-a)^2 = 1 + 1 - 2a + a^2 = 3 - a.$$
Note that, since $1 - u^{-1} + u^{-2} - u^{-3} + u^{-4}=0$,
$$\tau_+(\mathcal{F}ib_{u}) = 1 + u^{-2}(1-a)^2 = 1 + u^{-2}(2-u-u^{-1}) = 1 + 2u^{-1} - u^{-1} - u^{-3}$$
coincides with $u^{-2}-u^{-4}$.
Similarly,
$$\tau_-(\mathcal{F}ib_{u}) = 1 + u^2(1-a)^2 = 1 + u^2(2-u-u^{-1}) = 1 + 2u^{2} - u^{3} - u = u^2-u^4.$$
Thus for the square of the multiplicative central charge
$$\xi(\mathcal{F}ib_{u})^2 = \frac{\tau_+}{\tau_-} = \frac{u^{-4}(u^2-1)}{u^2(1-u^2)} = -u^{-6} = u^{-1}.$$
\end{proof}
\begin{proposition}
The ribbon category $\mathcal{F}ib_{u}$ is modular.
The $S$- and $T$-matrices are
$$S = \frac{1}{\sqrt{3-a}}\left(\begin{array}{cc}1& 1-a \\ 1-a & -1 \end{array}\right),\quad T = u^\frac{1}{6}\left(\begin{array}{cc} 1 & 0 \\ 0 & u^2 \end{array}\right).$$
\end{proposition}
\begin{proof}
The first row (and column) of the $S$-matrix is filled with dimensions (in our case $1$ and $d$).
The only non-trivial entry is the lower right corner, which is the trace of the square of the braiding on $X$.
Since the braiding on $X$ has a form $(u^21_I,u1_X)$ with respect to the decomposition $X\otimes X = I\oplus X$.
Its square has eigenvalues $(u^41_I,u^21_X)$.
Then the trace $tr(c_{X,X}^2)$ can be written as
$$u^4+u^2dim(X) = u^4+u^2(1-u-u^{-1}) = u^4-u^3+u^2-u = -1.$$
By the definition the matrix $T$ up to the factor $\xi(Fib_u)^{-\frac{1}{3}} = u^\frac{1}{6}$ is a diagonal matrix with diagonal entries being $\theta_1$ and $\theta_X$.
\end{proof}
The results of this section are summarized by the following theorem.
\begin{theorem}
Every braided balanced structure on a Fibonacci category is modular.
Thus there are four non-equivalent Fibonacci modular categories $\mathcal{F}ib_{u}$, parameterized by primitive roots of unity u of order 10.
\end{theorem}
Here is a simple consequence of the above theorem.
\begin{corollary}\label{decomp}
Let ${\cal C}$ be a modular category and let $X\in{\cal C}$ be such that $X^{\otimes 2} = I\oplus X$. Then
$${\cal C}\simeq \mathcal{F}ib_u\boxtimes{\cal D}$$ for some $u$ and a modular category ${\cal D}$.
\end{corollary}
\begin{proof}
Since any braided balanced structure on a Fibonacci category is modular, $X$ generates a modular subcategory $\mathcal{F}ib_u$ for some $u$.
Then by M\"uger's decomposition formula (\ref{mudec}) $${\cal C}\simeq \mathcal{F}ib_u\boxtimes{\cal C}_{\cal C}(\mathcal{F}ib_u).$$
\end{proof}
Note that it is the specific feature of the Fibonacci category that inforces the decomposition in corollary \ref{decomp}. For example an object $X$ in a modular category ${\cal C}$ with the property $X^{\otimes 2}=I$ does not necessarily generates a modular subcategory in ${\cal C}$.
\section{NIM-representations of $Fib^{\times\ell}$ and algebras in $\mathcal{F}ib^{\boxtimes\ell}$}
In this section we study commutative, ribbon algebras in $\mathcal{F}ib^{\boxtimes \ell}$.
We do this by classifying NIM-representations of the Fibonacci fusion rule $Fib$ and its tensor powers $Fib^{\boxtimes \ell}$.
We encode NIM-representations by certain types of oriented graphs.
Nodes correspond to elements of a NIM-set $M$.
Edges are colored in $\ell$ colours.
Two nodes $m$ and $n$ are the source and the target of an $i$-th coloured edge respectively iff the multiplicity $(x*m,n)$ of $n$ in $x_i*m$ is non-zero.
Here $x_i=1\otimes...\otimes 1\otimes X\otimes 1\otimes...\otimes 1$, where $X$ is in the $i$-th component.
\subsection{NIM-representation of $Fib^{\boxtimes \ell}$}
We begin by analyzing NIM-representations of $Fib$. Here we have only one colour for the edges.
\begin{lemma}\label{nimfib}
Up to isomorphism there is only one connected NIM-graph for the Fibonacci fusion rule
$$\xymatrix{m\ar@{-}[rr] && n \ar@(ur,dr) }\quad\,\,\, .$$
\end{lemma}
\begin{proof}
Let $m$ be a node of the graph of $M$.
Write
\begin{equation}\label{dec}
x*m = \sum_{i=1}^{k} m_{i}.
\end{equation}
\noindent An initial requisite observation is the following
$$x^{2}*m = x*(x*m) = x* \sum\limits_{i=1}^{k} m_{i} = k\cdot m + \sum\limits_{i=1}^{k} (x*m_{i} - m),$$
where $x*m_{i} - m$ is a non-negative linear combination of elements of the NIM-set $M$.
\noindent By the fusion rule $x^2=1+x$ we also have
$$x^{2}*m = x*m+ m$$
and thus
\begin{equation}x*m = (k-1)\cdot m + \sum\limits_{i=1}^{k} (xm_{i} - m) \end{equation}
Comparing this with (\ref{dec}) we see that all but one summand in (\ref{dec}) are equal to $m$ with the one remaining being say $n$.
Thus
$$x*m = (k-1)m+n.$$
Now applying the fusion rule again we see that
$$k\cdot m +n = x*m+m = x^2*m = (k-1)x*m+x*n = (k-1)^2m+(k-1)n+x*n$$
or
$$(k-(k-1)^2)m+(2-k)n = x*n.$$
If $m=n$ then $x*m=(2-(k-1)^2)m$, which is in contradiction with the initial assumption $x*m=k\cdot m$.
Thus $m\not=n$ and $k$ is at most 2.
We end up with two possibilities:
\newline
$k=1$ and $x*m=n$, or
\newline
$k=2$ and $x*m=m+n$.
\noindent Note that in the first case
$$x*n = x^{2}*m = (x+1)*m = xm + m = n + m.$$
\noindent While in the second case we have
$$x*n = x^{2}*m -x*m = (x+1)*m - x*m = m.$$
\noindent Thus up to the permutation $m\mapsto n$ we have only one possible indecomposable NIM-representation and one possible NIM-graph for $\mathcal{F}ib$ fusion rule.
\end{proof}
We now treat the general case of NIM-representations over tensor powers of $Fib$.
Let $M$ be a NIM-representation of $\textrm{$\fibfrmod$}^{\boxtimes \ell}$.
It follows from lemma \ref{nimfib} that for any $m\in M$ and any $i=1,...,\ell$ the multiplicity $(x_{i}*m,m)$ can not be greater than 1.
Define a map $\Gamma: M \to \{ 0 , 1 \}^\ell$ by $m \longmapsto \overline{m}$ where we write $\overline{m}_{i} = (x_{i}*m,m)$.
Let $<$ be the natural partial order on $\{ 0 , 1 \}^\ell$.
\begin{lemma}\label{gammalemma}
If nodes, corresponding to $n,m\in M$, are connected in the graph of the NIM representation $M$, then $\overline{m} < \overline{n}$ or $\overline{n} < \overline{m}$ in $\{ 0 , 1 \}^{\mathfrak{n}}$.
\end{lemma}
\begin{proof}
Suppose that $n$ and $m$ are connected by the $j$-th colored edge.
Then we have $(x_{j}*n,n)=0$, $(x_{j}*n,m)=1$, and $(x_{j}*m,m)=(x_{j}*m,n)=1$.
\noindent In other words $x_{j}*m= m+n$ and so $n = x_{j}*m - m$.
\noindent We need to show that for any (other) $i$ that we have $\overline n_i\leq\overline m_i$.
For each $i$ there are two possibilities, either $x_{i}*m = k$ ($\overline{m}_{i}=0$) or $x_{i}*m = k + m$ ($\overline{m}_{i}=1$).
While it is obvious that $\overline n_i\leq\overline m_i$ in the second case, in the first case we have
$$x_{j}*k = (x_{j}x_{i})*m = (x_{i}x_{j})*m = x_{i}*n + x_{i}*m = x_{i}*n + k$$
\noindent and so $(x_{j}*k,x_{i}*n)=1$ which implies $(x_{i}*n,n)=0$ and $\overline n_i=\overline m_i$.
\end{proof}
\begin{remark} The image of $\Gamma$ is a sub-lattice in $\{ 0 , 1 \}^\ell$.\end{remark}
\begin{lemma}\label{embedlemma}
Let ${\cal{M}}$ be a NIM-representation of $\textrm{$\fibfrmod$}^{k}$.
Let $m\in{\cal{M}}$ such that $(x_{i}m,m) = (x_{i}x_{j} \ast m,m)=0$ for all $i,j = 1,...,k$.
Then the assignment $y\in\textrm{$\fibfrmod$}^{k} \mapsto y \ast m$ defines an embedding of NIM-representations $\textrm{$\fibfrmod$}^{k}\hookrightarrow{\cal{M}}$.
\end{lemma}
\begin{proof}
We are required to show that $y\in\textrm{$\fibfrmod$}^{k} \mapsto y \ast m$ is an embedding, i.e. for $\epsilon ,\eta\in\{ 0,1 \}^{k}$
\begin{equation}\label{eqn39}(x_{1}^{\epsilon_{1}}...x_{k}^{\epsilon_{k}} \ast m,x_{1}^{\eta_{1}}...x_{k}^{\eta_{k}} \ast m) = \delta_{\epsilon_{1}\eta_{1}}...\delta_{\epsilon_{k}\eta_{k}}.\end{equation}
Assume (by induction) that this is true for any $\epsilon$ and $\eta$ with supports (sets of indexes of non-zero coordinates) in some proper subset of $[k] = \{ 1...k \}$.
We start by proving $(x_{1}...x_{k} \ast m,m) = 0$.
\noindent If $(x_{1}...x_{k} \ast m,m) \neq 0$ then by the assumption that $x_{1}...x_{k-1} \ast m$ and $x_{k} \ast m$ are simple we have
$$(x_{1}...x_{k-1} \ast m,x_{1}...x_{k-1} \ast m) = 1$$
and
$$(x_{k} \ast m,x_{k} \ast m) = 1$$
hence
$$(x_{1}...x_{k} \ast m,m) = (x_{1}...x_{k-1} \ast m,m)$$
should be equal to $1$, or $x_{1}...x_{k-1} \ast m = x_{k} \ast m$.
Similarly $x_{1} = x_{2}...x_{k}\ast m$.
\noindent But then
\begin{eqnarray*}
x_{1}\ast m & = & x_{2}...x_{k}\ast m \\
& = & (x_{2}...x_{k-1})\ast(x_{2}...x_{k-1}\ast m) \\
& = & x_{1}(x_{2} +1)...(x_{k-1}+1)\ast m \\
& = & x_{1}\ast m + x_{1}x_{2}\ast m + ... + x_{1}x_{k-1}\ast m + ...
\end{eqnarray*}
which contradicts, for example, that $(x_{1}x_{2}\ast m, x_{1}x_{2}\ast m)=1$.
\noindent Thus $(x_{1}^{\epsilon_{1}}...x_{k}^{\epsilon_{k}}\ast m , m) = 0$ for any non-zero $\epsilon\in \{ 0,1 \}^{k}$.
So we have
$$(x_{1}^{\epsilon_{1}}...x_{k}^{\epsilon_{k}} \ast m,x_{1}^{\eta_{1}}...x_{k}^{\eta_{k}} \ast m) = (m,m) \sum\limits_{i=1}^{k} \delta_{\epsilon_{i}\eta_{i}} + \textrm{ sum of } (x_{1}^{\xi_{1}}...x_{k}^{\xi_{k}}\ast m,m)$$
for non-zero $\xi_{1}...\xi_{k}$.
This proves (\ref{eqn39}).
That the image of $y\in\textrm{$\fibfrmod$}^{k} \mapsto y\ast m$ is a NIM-subrepresentation is obvious (see section \ref{nimrepsec}).
\end{proof}
\begin{theorem}\label{descriptiontheorem}
Any indecomposable NIM-representation of $\textrm{$\fibfrmod$}^{\ell}$ is of the form $\textrm{$\fibfrmod$}^{\lambda}$ for some set theoretic partition $\lambda$ of $[\ell] = \{ 1...\ell \}$.
\end{theorem}
\begin{proof}
Let $M$ be an indecomposable NIM-representation of $\textrm{$\fibfrmod$}^{\ell}$.
Our first step is to use lemma \ref{gammalemma} to show that there exists a $m\in M$ such that
$$(x_{i} \ast m ,m) = 0\quad\forall i = 1,...,\ell.$$
Indeed, let $m$ be an element with a minimal (with respect to the partial order on $\{ 0,1 \}^{\ell}$) $\Gamma (m)$.
If $\Gamma (m)\neq 0$, then $\exists i$ such that $\Gamma (m)_{i} = (x_{i}\ast m,m)=1$ and then
$$((x_{i}-1)\ast m , (x_{i}-1)\ast m) = ((x_{i}-1)^{2}\ast m,m) = 2(m,m) - (x_{i}\ast m,m) = 1.$$
Thus $(x_{i}-1)\ast m\in M$.
By lemma \ref{gammalemma}, $\Gamma ((x_{i}-1)\ast m) < \Gamma (m)$ which contradicts the assumption.
Hence $\Gamma (m) = 0$ or $(x_{i}\ast m,m) = 0$ $\forall i = 1,...,\ell$.
This in particular implies that
$$(x_{i}\ast m , x_{i}\ast m) = (x_{i}^{2}\ast m , m) = (x_{i}\ast m , m) + (m,m) = 1$$
and so $x_{i}\ast m \in M$.
Now define a set-theoretic partition $\lambda$ of $[\ell]$ by putting $i,j\in [\ell]$ in a given partition if and only if $x_{i}\ast m = x_{j}\ast m$.
Let $k$ be the number of parts of $\lambda$.
By permuting elements of $[\ell]$ we can assume that $x_{1},...,x_{k}$ lie in different parts of $\lambda$.
By lemma \ref{embedlemma} the assignment $y\mapsto y\ast m$ defines an injective map $\textrm{$\fibfrmod$}^{k} \hookrightarrow M$ of NIM-representations of $\textrm{$\fibfrmod$}^{k}$.
This obviously extends to the injective map $\textrm{$\fibfrmod$}^{\ell} \hookrightarrow M$ of NIM-representations of $\textrm{$\fibfrmod$}^{\ell}$.
By lemma \ref{cr} and the indecomposability of $M$, this map must be iso.
\end{proof}
\begin{example}NIM-representations of $\mathcal{F}ib^{\boxtimes 2}$
\end{example}
We have two colours in this case, which we depict by a solid line and a dashed line.
We have two set-theoretical partitions of $[2]$. The first $\{1\}\cup\{2\}=[2]$ corresponds to the square
$$
\xymatrix{
&& \cdot \ar@{.}[ddrr] \ar@{-}@(u,r) && \\
&&&& \\
\cdot \ar@{-}[uurr] \ar@{.}[ddrr] &&&& \cdot \ar@{-}@(u,r) \ar@{.}@(d,r) \\
&&&& \\
&& \cdot \ar@{-}[uurr] \ar@{.}@(d,r) &&
}
$$
\newline
while the second, $\{1,2\}=[2]$, corresponds to the double interval
\newline
$$\renewcommand{\objectstyle}{\labelstyle}
\xymatrix{
\cdot \ar@{-}@<+0.5ex>[rr] \ar@<-0.5ex>@{.}[rr] && \cdot \ar@{-}@(ul,ur) \ar@{.}@(ur,dr)
}$$
\newline
\begin{example}NIM-representations of $\mathcal{F}ib^{\boxtimes 3}$
\end{example}
We have three colours in this case, which we depict by a solid line, a broken line and a dashed line.
We have four set-theoretical partitions of $[3]$.
An example is $\{1\}\cup\{2\}\cup\{3\}=[3]$ which corresponds to the cube
\newline
$$
\xymatrix{
&& \cdot \ar@{.}[ddrr] \ar@{-}@(ul,ur) \ar@{--}[rr] && \cdot \ar@{.}[ddrr] \ar@{-}@(ul,ur) \ar@{--}@(ur,dr) &&& \\
&&&& & \\
\cdot \ar@{-}[uurr] \ar@{.}[ddrr] \ar@{--}[rr] && \cdot \ar@{-}[uurr] \ar@{.}[ddrr] \ar@{--}@(ur,dr) && \cdot \ar@{-}@(u,r) \ar@{.}@(d,r) \ar@{--}[rr] && \cdot \ar@{-}@(ul,ur) \ar@{--}@(ur,dr) \ar@{.}@(dl,dr) & \\
&&&&& \\
&& \cdot \ar@{-}[uurr] \ar@{.}@(dl,dr) \ar@{--}[rr] && \cdot \ar@{-}[uurr] \ar@{--}@(ur,dr) \ar@{.}@(dl,dr) &&&
}
$$
\newline
\newline
\newline
Although we do not classify all possible module categories of $\mathcal{F}ib^{\boxtimes\ell}$, the description of their NIM-representations (obtained in theorem \ref{descriptiontheorem}) is enough to prove that there are no non-trivial ribbon algebras in $\mathcal{F}ib^{\boxtimes\ell}$ which we establish in the next section.
\subsection{Commutative algebras in $\mathcal{F}ib^{\boxtimes\ell}$}
Here we look at commutative, ribbon algebras in products of Fibonacci modular categories.
\begin{theorem}\label{noalg}
There are no ribbon algebras in $\mathcal{F}ib_u^{\boxtimes\ell}$.
\end{theorem}
\begin{proof}
Let $A$ be an indecomposable algebra in ${\cal{C}} = \mathcal{F}ib_u^{\boxtimes\ell}$.
Then ${\cal{C}}_{A}$ is an indecomposable ${\cal{C}}$-module category (see section \ref{algandmods}) of $A$-modules in ${\cal{C}}$.
As was noted in section \ref{algandmods} the forgetful functor $F:{\cal{C}}_{A}\to{\cal{C}}$ (forgetting the module structure) is a functor of module categories over ${\cal{C}}$ and has the right adjoint $G:{\cal{C}}\to{\cal{C}}_{A}$ which is again a ${\cal{C}}$-module functor.
Note that the adjoint sends the monoidal unit $I$ to $A$ as a module over itself.
Hence for the NIM-representation $M$ of ${\cal{C}}_{A}$ we have maps of NIM-representations
$$f:M\to\textrm{$\fibfrmod$}^{\boxtimes\ell}\,\, \textrm{ and }\,\, g:\textrm{$\fibfrmod$}^{\boxtimes\ell}\to M$$
which are adjoint, i.e.
$$(g(y),m)_{M} = (g,f(m))_{\textrm{$\fibfrmod$}^{\boxtimes\ell}}.$$
\noindent Since ${\cal{C}}_{A}$ is indecomposable as a ${\cal{C}}$-module category, so is its fusion rule $M$.
According to theorem \ref{descriptiontheorem} we should have $M \simeq \textrm{$\fibfrmod$}^{\lambda}$ for some set-theoretic partition $\lambda$ of $[\ell]$.
\noindent Assume that $\lambda$ has only one part $\lambda = (\ell)$.
In particular $M=\textrm{$\fibfrmod$}^{(\ell)}$ has just two simple objects: $m$ and $n$.
Assume that $m=g(1)$.
Since $g$ is a map of NIM-representations we have $g(x_{i})=n$ for all $i=1,...,\ell$ such that
$$g(x_{i}\ast 1) = x_{i} g(1) = x_{i}\ast m =n.$$
Hence for an arbitrary element $x_{i_{1}}...x_{i_{s}}$ of $\textrm{$\fibfrmod$}^{\boxtimes\ell}$
$$g(x_{i_{1}}...x_{i_{s}}\ast 1) = f_{s}\ast n + f_{s-1}\ast m$$
where $f_{s}$ is the $s$-th Fibonacci number.
Thus the adjoint map $f$ has the form
\begin{eqnarray}\label{f}
f(m) & = & 1 + \sum\limits_{s=1}^{\ell} f_{s-1}\ast\sum\limits_{i_{1}<...<i_{s}}x_{i_{1}}...x_{i_{s}} \\
f(n) & = & \sum\limits_{s=1}^{\ell} f_{s}\ast \sum\limits_{i_{1}<...<i_{s}}x_{i_{1}}...x_{i_{s}}
\end{eqnarray}
Obviously $f(m)$, which is the class of the algebra $A$ in $K_{0}(\mathcal{F}ib_u^{\boxtimes\ell})=\mathbb{Z}[\textrm{$\fibfrmod$}^{\boxtimes\ell}]$, cannot be ribbon.
\noindent Assume now that $\lambda$ is a set-theoretic partition of $[\ell]$ into ordered parts
$$\lambda = [1...\ell_{1}][\ell_{1}...\ell_{2}]...[\ell_{n-1}...\ell_{n}].$$
In this case according to theorem \ref{descriptiontheorem}, $M$ as a NIM-representation of $\mathcal{F}ib_u^{\boxtimes\ell} = \mathcal{F}ib_u^{\boxtimes\ell_{1}}\boxtimes ...\boxtimes\mathcal{F}ib_u^{\boxtimes\ell_{s}}$ has the form
$$M = \textrm{$\fibfrmod$}^{(\ell_{1})}\boxtimes ... \boxtimes \textrm{$\fibfrmod$}^{(\ell_{s})}$$
By the above $\mathbb{Z}[\textrm{$\fibfrmod$}^{(\ell_{j})}] = K_0((\mathcal{F}ib_u^{\boxtimes\ell_{i}})_{A})$ for some non-ribbon algebra $A_{i}\in\mathcal{F}ib_u^{\boxtimes\ell_{i}}$.
Then $\mathbb{Z}M = K_0(\boxtimes^{s}_{j=1} (\mathcal{F}ib_u^{\boxtimes\ell_{i}})_{A_{i}}) = K_0((\mathcal{F}ib_u^{\boxtimes\ell})_{A})$ where $A=\boxtimes^{s}_{j=1}A_{i}$.
Since $A_{i}$ are non-ribbon then so is $A$.
\newline
The case for general $\lambda$ can be reduced to the above by a permutation of $[n]$.
\end{proof}
\begin{remark}
Note that the product $\mathcal{F}ib_u\boxtimes\mathcal{F}ib_{u^{-1}}$ has a commutative ribbon algebra, whose class in $K_0(\mathcal{F}ib_u\boxtimes\mathcal{F}ib_{u^{-1}}) = Fib^{\otimes 2}$ is $1 + x_1x_2$ (see section \ref{coal}). The corresponding NIM-representation is $Fib^{(2)}$.
\newline
At the same time the argument of the proof of theorem \ref{noalg} works well for $\mathcal{F}ib_u^{\boxtimes\ell}\boxtimes\mathcal{F}ib_v^{\boxtimes m}$ as long as $uv\not= 1$, thus proving that there are no commutative ribbon algebras in $\mathcal{F}ib_u^{\boxtimes\ell}\boxtimes\mathcal{F}ib_v^{\boxtimes m}$ for such $u,v$.
\end{remark}
\section{Applications}
Here we look at vertex operator algebras (=chiral algebras) whose categories of modules are tensor products of Fibonacci categories.
\subsection{Rational vertex operator algebras} \label{vex}
We start by recalling basic facts about rational vertex operator algebras and their representations.
Let ${\cal{V}}$ be a rational vertex operator algebra (or VOA for short) of central charge $c$; that is, a vertex algebra satisfying conditions 1-3 from \cite[Section~ 1]{Hu}.
To each ${\cal{V}}$-module ${\cal{M}}$ one can associate its {\em character}
$$\chi_{\cal{M}}(q) = tr_{\cal{M}}(q^{L_0-\frac{c}{24}}).$$
Here $tr_{\cal{M}}$ is the trace of an operator on a vector space ${\cal{M}}$, $L_0$ is the zero Virasoro mode and $c$ is the central charge of ${\cal{V}}$.
The action of $L_0$ on ${\cal{M}}$ is diagonalisable so that
$${\cal{M}} = \bigoplus_{l}{\cal{M}}_l,\quad\quad {\cal{M}}_l = \{m\in{\cal{M}},\ L_0(m) = lm\}.$$
Moreover if ${\cal{M}}$ is irreducible then
$${\cal{M}} = \bigoplus_{n\geq0}{\cal{M}}_{h+n}$$
for some rational number $h=h_M$, the {\em conformal weight} of ${\cal{M}}$.
Thus the character of an irreducible ${\cal{M}}$ can be written as
$$\chi_{\cal{M}}(q) = q^{h-\frac{c}{24}}\sum_{n\geq 0}dim({\cal{M}}_{h+n})q^n.$$
The (complex) span of characters of a rational VOA is a module over the modular group $SL_2(\mathbb{Z})$ with respect to the action
$$S(\chi)(\tau) = \chi(-\frac{1}{\tau}),\quad\quad T(\chi) = \chi(\tau+1).$$
Here we make the change of variable $q = e^{2\pi i\tau}$.
It is proved in \cite{Hu} that the category ${\cal R}{\it ep}(V)$ of $V$-modules of finite length has the natural structure of a modular tensor category.
We will denote the fusion of $V-$modules $M$ and $N$ by $M*N$.
The relation between the central charge $c_V$ of a unitary rational VOA $V$ and the central charge of the category of its modules ${\cal R}{\it ep}(V)$ is given by (e.g. see \cite{ms,re}): $$\xi({\cal R}{\it ep}(V)) = e^{\frac{2\pi ic_V}{8}}.$$ In the non-unitary case only the square of the above identity is true. The ribbon twist on an irreducible $V$-module $M$ is related to its conformal weight $h_M$ as follows (see \cite{Hu}):
$$\theta_M = e^{2\pi ih_M}I_M.$$
Note that a rational vertex algebra has to be {\em simple} (i.e. has no non-trivial ideals).
This, in particular, means that VOA maps between rational vertex algebras are monomorphisms.
The category of modules ${\cal R}{\it ep}({\cal{V}}\otimes {\cal{U}})$ of the tensor product of two (rational) vertex algebras is ribbon equivalent to the tensor product ${\cal R}{\it ep}({\cal{V}})\boxtimes{\cal R}{\it ep}({\cal{U}})$ of the categories of modules (see, for example \cite{fhl}). Sometimes following physics tradition we will write ${\cal{V}}\times {\cal{U}}$ for the tensor product ${\cal{V}}\otimes {\cal{U}}$ of VOAs. For ${\cal{V}}$-module $M$ and ${\cal{U}}$-module $N$ the ${\cal{V}}\otimes {\cal{U}}$-module with underlying vector space $M\otimes N$ will be denoted by $M\boxtimes N$.
Now consider an extension of vertex operator algebras $V\subset W$, where $V$ is a rational vertex operator algebra of $W$. Then $W$ viewed as a $V$-module decomposes into a finite direct sum of irreducible $V$-modules.
Moreover $W$ considered as an object $A\in {\cal R}{\it ep}(V)$ has a natural structure of simple, separable, commutative, ribbon algebra, see \cite[Theorem 5.2]{KiO}.
The converse is also true if the conformal weights of the irreducible components of the underlying $V$-representation of the algebra are all positive. In particular it is always true if $V$ is a unitary VOA.
\medskip
{\bf Holomorphic VOAs}
\newline
A VOA is called {\em holomorphic} if it has only one irreducible module, namely itself. It is known (see e.g. \cite{sch,dlm}) that the central charge of a holomorphic VOA is divisible by 8.
\medskip
{\bf Diagonal extensions}
\newline
Suppose that ${\cal{V}}$ is a VOA whose category of representations has a form ${\cal C}\boxtimes\overline{\cal C}$ for some modular category ${\cal C}$. The diagonal algebra $Z\in {\cal C}\boxtimes\overline{\cal C}$ (from section \ref{coal}) defines a holomorphic extension of ${\cal{V}}$, which is automatically a VOA in the unitary case.
\medskip
{\bf Simple current extensions}
\newline
Recall that a module over a VOA is called {\em simple current} if it is invertible.
Let $S$ be a subgroup of the group of simple currents of a rational VOA such that conformal weights of elements of $S$ are positive integer (in the non-unitary case we should also assume that the braiding $c_{s,s}=1$ for any $s\in S$). The object $\oplus_{s\in S}s$ of the category ${\cal R}{\it ep}({\cal{V}})$ has a structure of simple, separable, commutative, ribbon algebra and the corresponding VOA extension of ${\cal{V}}$ is called the {\em simple current} extension (see e.g. \cite{fss}).
\medskip
{\bf Conformal embeddings}
\newline
Let $\bigoplus_i{\frak g}^i \subset {\frak g}'$ be an embedding (here ${\frak g}^i$ and ${\frak g}'$ are finite dimensional simple Lie algebras).
We will symbolically write $\oplus_i({\frak g}^i)_{k_i}\subset {\frak g}'_{k'}$ if the restriction of a $\hat {\frak g}'-$module of level $k'$ to $\hat {\frak g}^i$ has level $k_i$ (in this case the numbers $k_i$ are multiples of $k'$).
Such an embedding defines an embedding of vertex algebras $\otimes_iV({\frak g}^i,k_i)\subset V({\frak g}',k')$; but in general this embedding does not preserve the Virasoro element.
In the case when it does, the embedding $\oplus_i({\frak g}^i)_{k_i}\subset {\frak g}'_{k'}$ is called {\em conformal}, see \cite{bb, SW, KW}.
\medskip
{\bf Cosets}
\newline
Let ${\cal{U}}\subseteq {\cal{V}}$ be an embedding of rational vertex algebras which does not preserve conformal vectors $\omega_{\cal{U}},\omega_{\cal{V}}$ (only operator products are preserved).
The {\em centralizer} $C_{\cal{V}}({\cal{U}})$ is a vertex algebra with the conformal vector $\omega_{\cal{V}}-\omega_{\cal{U}}$ (see \cite{fz}).
Note that the tensor product ${\cal{U}}\otimes C_{\cal{V}}({\cal{U}})$ is mapped naturally to ${\cal{V}}$ and this map is a map of vertex operator algebras.
In the case when ${\cal{V}},{\cal{U}}$ and $C_{\cal{V}}({\cal{U}})$ are rational this map is an embedding (by simplicity of ${\cal{U}}\otimes C_{\cal{V}}({\cal{U}})$) and ${\cal{V}}$ is a commutative algebra in the category of modules
$${\cal R}{\it ep}({\cal{U}})\boxtimes{\cal R}{\it ep}(C_{\cal{V}}({\cal{U}}))\simeq {\cal R}{\it ep}({\cal{U}}\otimes C_{\cal{V}}({\cal{U}})).$$
The coset construction allows us to describe VOAs with a given completely anisotropic category of representations.
\begin{proposition}
Let ${\cal{V}}$ be a rational VOA with a completely anisotropic category of representations $Rep({\cal{V}})$. Then any rational VOA whose category of representations is $\overline{Rep({\cal{V}})}$ is a coset ${\cal{U}}/{\cal{V}}$ of some holomorphic ${\cal{U}}$.
\end{proposition}
\begin{proof}
Let ${\cal{W}}$ be a VOA such that $Rep({\cal{W}}) = \overline{Rep({\cal{V}})}$. Then $$Rep({\cal{V}}\otimes {\cal{W}}) = Rep({\cal{V}})\boxtimes\overline{Rep({\cal{V}})}$$ contains a commutative ribbon separable indecomposable algebra with trivial category of local modules. Thus the corresponding VOA extension ${\cal{U}}$ of ${\cal{V}}\otimes{\cal{W}}$ is holomorphic. The coset ${\cal{U}}/{\cal{V}}$ is an extension of ${\cal{W}}$ and by complete anisotropy of ${\cal R}{\it ep}({\cal{W}})$ must coincide with ${\cal{W}}$.
\end{proof}
\medskip
{\bf Lattice VOAs}
\newline
Let $L$ be a positive definite integral even lattice of rank d, $(\ ,\ )$ the pairing. Its
associated vertex operator algebra ${\cal{V}}_L$ is rational and unitary and has central charge d. Its irreducible representations
are in one-to-one correspondence with the cosets in $L^\# /L$, where $L^\#$ is the dual lattice
of $L$ \cite{do}.
Fusion rule are given by the addition in the finite abelian group $L^\# /L$.
\medskip
{\bf Affine VOAs}
\newline
Let ${\frak g}$ be a finite dimensional simple Lie algebra and let $\hat {\frak g}$ be the corresponding affine Lie algebra. For a positive integer $k$ let ${\frak g}_k = V({\frak g},k)$ be the simple vertex operator algebra associated with the vacuum $\hat {\frak g}-$module of level $k$ \cite{fz}. This VOA is rational of central charge
\begin{equation} \label{cformula}
c({\frak g}_k)=\frac{k\dim {\frak g}}{k+h^\vee},
\end{equation}
where $h^\vee$ is the dual Coxeter number of the Lie algebra ${\frak g}$.
In particular the central charge of $A_{1,k}$ is $\frac{3k}{k+2}$.
The irreducible $V(A_1,k)$-modules are irreducible highest weight modules $L_{A_{1,k}}(i)$ with $i=0,...,k$.
The conformal weight of $L_{A_1}(k,i)$ is $\frac{i(i+2)}{4(k+2)}$.
The fusion rule is given by
$$L_{A_{1,k}}(i)*L_{A_{1,k}}(j) = \left\{\begin{array}{cc}\bigoplus\displaylimits^{min(i,j)}_{s=0} L_{A_{1,k}}(i+j-2s), & i+j<k\\ \\ \bigoplus\displaylimits^{min(i,j)}_{s=i+j-k} L_{A_{1,k}}(i+j-2s), & i+j\geq k\end{array}\right.$$
In particular $L_{A_{1,k}}(k)$ is invertible (a {\em simple current}):
$$L_{A_{1,k}}(k)*L_{A_{1,k}}(k) = L_{A_{1,k}}(0).$$ Its conformal weight is $k/4$.
Its action on other irreducible modules is
$$L_{A_{1,k}}(k)*L_{A_{1,k}}(i) = L_{A_{1,k}}(k-i).$$
\newline
\medskip
{\bf Minimal models}
\newline
Let $1<p<q$ be a coprime integers. Following \cite{bpz,fms,wa} consider the minimal (Virasoro) VOA $M(p,q)={\cal V}{\it ir}_{c_{p,q}}$ of
central charge
$$c_{p,q} = 1-6\frac{(p-q)^2}{pq}.$$
Irreducible representations of $M(p,q)$ are irreducible Virasoro highest weight modules $L_{Vir_{c_{p,q}}}(h_{(r,s)})=L_{p,q}(r,s)$ with conformal weights
$$h_{(r,s)} = \frac{(qr-ps)^2-(p-q)^2}{4pq}$$
labelled by
$$(r,s),\quad 1\leq r\leq p-1,\quad 1\leq s\leq q-1$$ modulo identification $(p-r,q-s) = (r,s)$.
Fusion rules have the following compact presentation:
$$L_{p,q}(r,s)L_{p,q}(r',s') = \osum\displaylimits_{i=1+|r-r'|}^{min(r+r'-1,2p-r-r'-1)}\ \ \osum\displaylimits_{j=1+|s-s'|}^{min(s+s'-1,2q-s-s'-1)}\ L_{p,q}(i,j),$$ where the primed summation indicates that $i$ and $j$ increment in twos.
Note that for $p>2$ the minimal model $M(p,q)$ has a simple current $L_{p,q}(p-1,1)$ of conformal weight $\frac{(p-2)(q-2)}{4}$. Its fusion is
$$L_{p,q}(p-1,1)L_{p,q}(r,s) = L_{p,q}(p-r,s).$$ In particular $L_{p,q}(p-1,1)$ generates a cyclic group of order 2.
\subsection{Fibonacci chiral algebras}
Here we look at examples where the Fibonacci categories are categories of representations of rational vertex operator algebras (VOAs of {\em Fibonacci} type). In particular we present all four modular categories $\mathcal{F}ib_u$ as categories of representations of rational VOAs.
Note that the formula
\begin{equation}\label{cch}
u^{-1} = \xi(\mathcal{F}ib_u)^2 = e^{\frac{\pi ic_V}{2}}
\end{equation}
allows us to determine the structure of a modular category for the category of representations of a Fibonacci VOA. The relation $\xi(\mathcal{F}ib_u)^4 = u^{-2} = \theta_X$ implies that for a Fibonacci VOA ${\cal{V}}$
$$\frac{c_{\cal{V}}}{2}\equiv h\ \ (1),$$ where $h$ is the conformal weight of the non-trivial irreducible representation of ${\cal{V}}$.
\begin{example}({Affine VOA} $G_{2,1}$)
The central charge of $G_{2,1}$ is $14/5$.
The conformal weights of irreducible $G_{2,1}$-modules are $0$ and $2/5$.
\newline
Since the fusion rule of $G_{2,1}$ is of Fibonacci type, the value of the central charge (or conformal weights) for $G_{2,1}$ implies that
$${\cal R}{\it ep}(G_{2,1})\cong \mathcal{F}ib_{e^{\frac{3\pi i}{5}}}.$$
The conformal embedding $A_{1,28}\subset G_{2,1}$ allows us to write two irreducible $G_{2,1}$-modules as sums of irreducible $A_{1,28}$-modules:
$$L_{G_{2,1}}(0) = L_{A_{1,28}}(0)\oplus L_{A_{1,28}}(10)\oplus L_{A_{1,28}}(18)\oplus L_{A_{1,28}}(28),$$ $$L_{G_{2,1}}(1) = L_{A_{1,28}}(6)\oplus L_{A_{1,28}}(12)\oplus L_{A_{1,28}}(16)\oplus L_{A_{1,28}}(22).$$
Another conformal embedding $A_{1,3}\times A_{1,1}\subset G_{2,1}$ allows us to identify $G_{2,1}$ with the simple current extension of $A_{1,3}\times A_{1,1}$.
Indeed, the $A_{1,3}\times A_{1,1}$-module $L_{A_{1,3}}(3)\boxtimes L_{A_{1,1}}(1)$ is invertible (simple current) and has conformal weights 1.
The commutative algebra
$$A = L_{A_{1,3}}(0)\boxtimes L_{A_{1,1}}(0)\oplus L_{A_{1,3}}(3)\boxtimes L_{A_{1,1}}(1)$$
(the simple current extension) in ${\cal R}{\it ep}(A_{1,3}\times A_{1,1}) = {\cal R}{\it ep}(A_{1,3})\boxtimes {\cal R}{\it ep}(A_{1,1})$ coincides with $G_{2,1}$.
In particular, the non-trivial irreducible local $A$-module (the irreducible ${\cal{V}}$-module) has the following decomposition into irreducible $A_{1,3}\times A_{1,1}$-modules:
$$L_{A_{1,3}}(2)\boxtimes L_{A_{1,1}}(0)\oplus L_{A_{1,3}}(1)\boxtimes L_{A_{1,1}}(1).$$
\end{example}
\begin{example}({Affine VOA} $F_{4,1}$)
The central charge of $F_{4,1}$ is $26/5$.
The conformal weights of irreducible $F_{4,1}$-modules are $0$ and $3/5$.
\newline
The fusion rule of $F_{4,1}$ is of Fibonacci type. Thus by the formula (\ref{cch})
$${\cal R}{\it ep}(F_{4,1})\cong \mathcal{F}ib_{e^{\frac{7\pi i}{5}}}.$$
\newline
The conformal embedding $A_{2,2}\times A_{2,1}\subset F_{4,1}$ allows us to identify $F_{4,1}$ with the simple current extension of $A_{2,2}\times A_{2,1}$.
\newline
The conformal embedding $G_{2,1}\times F_{4,1}\subset E_{8,1}$ allows us to identify $F_{4,1}$ with the centraliser (coset) of $G_{2,1}$ in $E_{8,1}$:
$$F_{4,1} = \frac{E_{8,1}}{G_{2,1}}.$$
Note that the converse is also true: $G_{2,1}$ is the centraliser (coset) of $F_{4,1}$ in $E_{8,1}$.
\end{example}
\begin{example}({Simple current extension of} $A_{1,8}$)
The irreducible $A_{1,8}$-module $L_{A_{1,8}}(8)$ is invertible (simple current) and has conformal weight 1.
The commutative algebra $A = L_{A_{1,8}}(0)\oplus L_{A_{1,8}}(8)$ (the simple current extension) is maximal in ${\cal R}{\it ep}(A_{1,8})$ and ${\cal R}{\it ep}(A_{1,8})_A^{loc}$ has the type $\mathcal{F}ib^{\boxtimes 2}$. Thus ${\cal{V}} = L_{A_{1,8}}(0)\oplus L_{A_{1,8}}(8)$ has a structure of VOA (a VOA extension of $A_{1,8}$) such that ${\cal R}{\it ep}({\cal{V}})$ is equivalent to $\mathcal{F}ib^{\boxtimes 2}$ as a monoidal category.
Indeed it follows from the $A_{1,8}$-fusion rule that there are five non-isomorphic induced $A$-modules
$$A\otimes L_{A_{1,8}}(0) = A\otimes L_{A_{1,8}}(8),\quad A\otimes L_{A_{1,8}}(1) = A\otimes L_{A_{1,8}}(7),$$ $$A\otimes L_{A_{1,8}}(2) = A\otimes L_{A_{1,8}}(6),\quad A\otimes L_{A_{1,8}}(3) = A\otimes L_{A_{1,8}}(5),\quad A\otimes L_{A_{1,8}}(4)$$
and all but the last one are irreducible as $A$-modules.
The last one is a sum of two (non-isomorphic) $A$-modules.
By looking at conformal weights it can be seen that the $A$-modules induced from $L_{A_{1,8}}(1)$ and $L_{A_{1,8}}(3)$ are non-local and the rest is local.
Thus irreducible local $A$-modules (the irreducible ${\cal{V}}$-modules) have the following decompositions into irreducible $A_{1,8}$-modules:
$$L_{A_{1,8}}(0)\oplus L_{A_{1,8}}(8),\quad L_{A_{1,8}}(2)\oplus L_{A_{1,8}}(6),\quad L_{A_{1,8}}(4),\quad L_{A_{1,8}}(4).$$
The central charge of ${\cal{V}}$ coincides with the central charge of $A_{1,8}$ and is equal to $12/5$.
The conformal weights of irreducible ${\cal{V}}$-modules are $0$, $1/5$ and $3/5$ (the last one appearing twice).
This in particular shows that although
$${\cal R}{\it ep}(V)\cong \mathcal{F}ib_{e^{\frac{7\pi i}{5}}}\boxtimes\mathcal{F}ib_{e^{\frac{7\pi i}{5}}}.$$
${\cal{V}}$ is not a tensor product of two VOAs of type $\mathcal{F}ib$ (in that case the conformal weight of the second irreducible module would be twice the conformal weight of the last two irreducibles).
Conformal embeddings allow us to represent ${\cal{V}}$ as cosets. The conformal embedding $A_{1,8}\times G_{2,1}\times G_{2,1}\subset E_{8,1}$ factors through
$$A_{1,8}\times G_{2,1}\times G_{2,1}\subset{\cal{V}}\times G_{2,1}\times G_{2,1}\subset E_{8,1}$$
and gives the following coset presentation for ${\cal{V}}$:
$${\cal{V}} = \frac{E_{8,1}}{G_{2,1}\times G_{2,1}}.$$
The conformal embedding $A_{1,8}\times G_{2,1}\subset F_{4,1}$ factors through
$$A_{1,8}\times G_{2,1}\subset{\cal{V}}\times G_{2,1}\subset F_{4,1}$$
and gives another coset presentation for ${\cal{V}}$:
$${\cal{V}} = \frac{F_{4,1}}{G_{2,1}}.$$
\end{example}
Note that although the chiral algebras
$\frac{F_{4,1}}{G_{2,1}}\times E_{8,1}$ and $F_{4,1}\times F_{4,1}$ (here $E_{8,1}$ is the holomorphic chiral algebra of central charge 8) have equivalent categories of representations and the same central charge they are not isomorphic. The way to see it is to compare their degree 1 components.
\begin{example}({Minimal VOA} $M(2,5) = {\cal V}{\it ir}_{-\frac{22}{5}}$)
The minimal VOA with the central charge $c_{2,5} = -22/5$.
The irreducible modules $L_{2,5}(1,1)$ and $L_{2,5}(1,2)$ have conformal weights $0$ and $-1/5$.
The fusion rule of $M(2,5)$ is of Fibonacci type. Thus
$${\cal R}{\it ep}(M(2,5))\cong \mathcal{F}ib_{e^{\frac{\pi i}{5}}}.$$
Being minimal and maximal at the same time $M(2,5)={\cal V}{\it ir}_{-\frac{22}{5}}$ is the only VOA of central charge $-22/5$ (which contains ${\cal V}{\it ir}_{-\frac{22}{5}}$).
\end{example}
\begin{example}({Simple current extension of minimal VOA} $M(3,10) = {\cal V}{\it ir}_{-\frac{44}{5}}$)
The minimal VOA $M(3,10)$ has central charge $c_{3,10} = -44/5$. Moreover we have an embedding
$${\cal V}{\it ir}_{-\frac{44}{5}}\subset {\cal V}{\it ir}_{-\frac{22}{5}}\otimes {\cal V}{\it ir}_{-\frac{22}{5}}$$
Note that ${\cal V}{\it ir}_{-\frac{22}{5}}\otimes {\cal V}{\it ir}_{-\frac{22}{5}}$ is a simple current extension of ${\cal V}{\it ir}_{-\frac{44}{5}}$. Indeed its decomposition as an ${\cal V}{\it ir}_{-\frac{44}{5}}$-module is
$L_{3,10}(1,1)\oplus L_{3,10}(2,1).$
\end{example}
\begin{example}{Maximal extension of $M(3,5)^{\times 8}\times M(2,5)^{\times 7}$}
The central charge of the minimal model $M(3,5)$ is $-3/5$. The irreducible representations are
$$1=L_{3,5}(1,1), \quad x=L_{3,5}(2,1),\quad y=L_{3,5}(1,2),\quad z=L_{3,5}(2,2).$$
Their conformal weights are
$$h_1=0,\quad h_x=\frac{3}{4},\quad h_y=-\frac{1}{20},\quad h_z=\frac{1}{5}.$$
Fusion rules of $M(3,5)$ have the form:
$$
\begin{array}{c|cccc}
\times & x &y & z \\ \hline x & 1 \\ y & z & 1+z \\ z & y & x+y & 1+z \end{array}$$
Thus the category $Rep(M(3,5))$ is a product of a Fibonacci category and a pointed category with the ${\bf Z}/2{\bf Z}$-fusion rule. The values of conformal weights of representations of $M(3,5)$ imply that
$$Rep(M(3,5)) \cong \mathcal{F}ib_{e^{\frac{9\pi i}{5}}}\boxtimes {\cal C}({\bf Z}/2{\bf Z},q,\theta),$$ with $\theta(1) = -i$. In particular the product $M(2,5)\times M(3,5)$ has an extension
$${\cal{V}} = L_{2,5}(1,1)\boxtimes L_{3,5}(1,1)\oplus L_{2,5}(1,2)\boxtimes L_{3,5}(2,2).$$ By looking at the characters it can be seen that although ${\cal{V}}$ is non-negatively graded ${\cal{V}} = \oplus_{n\geq 0}{\cal{V}}_n$, the degree zero component ${\cal{V}}_0$ is not one-dimensional.
The representation category of ${\cal{V}}$ is $Rep({\cal{V}}) = {\cal C}({\bf Z}/2{\bf Z},q,\theta)$. The 8-th power ${\cal C}({\bf Z}/2{\bf Z},q,\theta)$ has a trivialising algebra. This implies that the 8-th power ${\cal{V}}^{\times 8}$ has a holomorphic extension (of central charge $-40$). Similarly the category ${\cal C} = Rep(M(3,5)^{\times 8}\times M(2,5)^{\times 7})$ has a maximal algebra $A$ such that the category of local modules ${\cal C}_A^{loc}$ is $\mathcal{F}ib_{e^{\frac{9\pi i}{5}}}$. The corresponding maximal extension ${\cal{U}}$ of $M(3,5)^{\times 8}\times M(2,5)^{\times 7}$ is a non-negatively graded VOA (although with $dim({\cal{U}}_0)>1$) of central charge $-178/5$.
\end{example}
|
\section{Results}
We performed simulations of fracturing ranked surfaces on square
lattices. It is worth noting that, despite the similarities with random
percolation, the suppressing of connectivity poses a statistically
different problem. For example, while for $p\rightarrow1$ there is only
a single configuration in random percolation (all bonds occupied), in
ranked percolation there are $N!$, evenly weighted, possible
configurations, where $N$ is the total number of bonds. In classical
percolation the total number of configurations is $2^N$.
Figure~\ref{fig::mbbsize} shows the dependence of the number of bridge
bonds $N_{BB}$ on system size, for different fractions of occupied
bonds, namely, $p=p_c=0.5$, $p=0.51$, and $p=0.8$. As expected, at the
percolation threshold of classical percolation ($p=p_c$), the number of
bridge bonds diverges with system size as $N_{BB}\sim L^{1/\nu}$, where
$\nu$ is the correlation length exponent, with $\nu=4/3$ in $2D$; while
for $p=0.8$, $N_{BB}\sim L^{d_{BB}}$, with $d_{BB}=1.215\pm0.003$. The
latter is in fact observed at any $p>p_c$. We found the same result
(see {\it Appendix}) for site percolation and on other
lattices (star, triangular, and honeycomb), which provides strong
evidence for the universality of this exponent. For $p\gtrsim p_c$,
like $p=0.51$ we can observe, as depicted in Fig.~\ref{fig::mbbsize}, a
crossover between the two different regimes. The inset of
Fig.~\ref{fig::bbscale} shows $N_{BB}$, rescaled by $L^{d_{BB}}$, as a
function of $p$, for different system sizes. The number of bridge bonds
grows with $p$, such that, $N_{BB}\sim(p-p_c)^\zeta$, where
$\zeta=0.50\pm0.03$ is a novel exponent, which we call bridge-growth
exponent. The overlap of the different curves confirms that the fractal
dimension of the bridge bonds above $p_c$ is $d_{BB}$, for all $p>p_c$.
This result differs from classical percolation where fractality is
solely observed at criticality \cite{Broadbent57,Stauffer94} while,
above $p_c$, bridge bonds are only observed for finite systems
\cite{Coniglio89}.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{fig3.pdf}
\end{center}
\caption{
Crossover in the size dependence.
Number of bridge bonds, $N_{BB}$, for different $p$, namely, $p=p_c=0.5$ (circles), $p=0.51$ (stars), and $p=0.8$ (triangles).
The solid lines stand for the best fit.
At $p=p_c$, as conjectured by Coniglio \cite{Coniglio89}, $N_{BB}\sim L^{1/\nu}$, where $\nu=4/3$ is the critical exponent of the correlation length in $2D$.
For $p>p_c$, the number of bridge bonds scales with $L^{d_{BB}}$.
A crossover between the two regimes in system size is observed (stars) for $p$ in the neighborhood of $p_c$.
Systems of size $L^2$ have been considered, with $L$ ranging from $32$ to $4096$.
All results have been averaged over $10^4$ samples.
Error bars are smaller than the symbol size.
\label{fig::mbbsize}}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{fig4.pdf}
\end{center}
\caption{
Tricritical scaling, crossover, and data collapse.
Number of bridge bonds, $N_{BB}$, as a function of the fraction
of occupied bonds, $p$, for $2D$ with different system sizes $L=\{256,512,1024,2048,4096\}$.
The scaling function given by equation~(\ref{eq::scaling}) is applied, with $\theta=0.93$, obtaining $\zeta=0.50\pm0.03$.
In the inset, $N_{BB}$ has been rescaled by $L^{d_{BB}}$, where $d_{BB}=1.215$.
All results have been averaged over $10^4$ samples.
\label{fig::bbscale}}
\end{figure}
For polymer chains, at high temperatures, the excluded volume prevails
over attractive forces and the chain can be described as a self-avoiding
walk. When the temperature is reduced, the attractive forces become
relevant leading, at a theta-temperature, to a new exponent at
the crossover between two dimensions \cite{deGennes79,Chang91,Poole89}.
Analogously, in ranked percolation the fractal dimension of the bridge
bonds is $1/\nu$, at $p=p_c$, between $d_{BB}$ above $p_c$ and zero
below $p_c$. For the tricritical scaling we verify the following
\textit{ansatz},
\begin{equation}\label{eq::scaling}
N_{BB}=L^{1/\nu}\mathcal{F}\left[\left(p-p_c\right)L^\theta\right],
\end{equation}
where $\mathcal{F}[x]\sim x^{\zeta}$ for $x\neq0$, and is nonzero at
$x=0$; and the exponent $\theta$ is the crossover exponent. Therefore,
the following relation is obtained,
\begin{equation}\label{eq::scaling.relation}
\theta=\zeta^{-1}\left(d-\varphi-\frac{1}{\nu}\right),
\end{equation}
where $\varphi=d-d_{BB}$. In the main plot of Fig.~\ref{fig::bbscale}
we see, for $2D$, the scaling given by equation~(\ref{eq::scaling}),
with $\theta=0.93$.
The results above disclose a tricritical $p_c$ below which the fraction
of bridges in the bridge line vanishes in the thermodynamic limit. This
is identical to a minimum height in the bridge line, $H=h_\mathrm{min}$,
in the context of landscapes (Figs.~\ref{fig::rank}(a)-(b)). For a
cumulative distribution function of heights $H$, $P(H\leq h)$, the
minimum is given by $P(H\leq h_\mathrm{min})=p_c$. Note that, for
uncorrelated landscapes, regardless the distribution of heights, the set
of bridges only depends on their position in the rank. To observe the
new set of exponents on discretized landscapes
(Fig.~\ref{fig::rank}(c)), $N_{BB}$ is the number of sites in the bridge
line with both neighbors (one at each side) having height lower than
$h$, where $P(H\leq h)=p$.
To study the dependence of exponents $\zeta$ and $\varphi$ on the
spatial dimension, we analyze the same problem up to dimension six. On
a simple-cubic lattice ($3D$), above $p_c$, the set of all bridge bonds
has a fractal dimension $d_{BB}=2.50\pm0.05$ and grows with
$\zeta=1.0\pm0.1$ (see {\it Appendix}).
Figure~\ref{fig::nbbdim} shows the size dependence of $N_{BB}$, in the
limit $p=1$, for lattices with size $L^d$, where $2\leq d\leq 6$ is the
spatial dimension. In the inset, we plot the exponent $\varphi$
as a function of $d$. Since the set of bridge bonds blocks connectivity
from one side to the other, its fractal dimension must
follow $d-1\leq d_{BB} \leq d$, i.e., $0\leq\varphi\leq 1$. With
increasing dimension $\varphi$ decreases. At the upper-critical
dimension of percolation, $d_c=6$, $\varphi=0.0\pm0.1$ and the set of
bridge bonds becomes dense having the spatial dimension $d$.
Table~\ref{tab::exponents} summarizes the exponents for dimensions $2$
to $6$. The bridge-growth exponent grows with $d$ and converges to
$\zeta=1.5\pm0.7$ at the upper-critical dimension. For $d>6$, above the
critical dimension of percolation, the exponent $\varphi$ remains zero
and the dimension of the set of bridges is equal to $d$. This can be
understood from the fact that above $d_c$ one has an infinity of
spanning clusters \cite{Newman81} and thus many more possible bridges. Since the
dimension of the set of bridges at $p=p_c$ is $1/\nu$ and above $p_c$ is
$d$, the crossover exponent increases with $d$, with the relation given
by equation~(\ref{eq::scaling.relation}), where $\varphi=0$.
\begin{table}
\caption{
Table of exponents.
Values of the exponents $\zeta$ and
$\varphi$ up to dimension six. With increasing dimension, the $\zeta$
exponent converges to $\zeta=1.5\pm0.7$ and $\varphi$ goes to zero,
revealing that the set of bridge bonds is dense.
\label{tab::exponents} }
\begin{tabular}{ccc}
\hline\hline $d$ & $\zeta$ & $\varphi$ \\
\hline $2$ & $0.50\pm0.03$ & $0.785\pm0.003$ \\
$3$ & $1.0\pm0.1$ & $0.50\pm0.02$ \\
$4$ & $1.3\pm0.5$ & $0.26\pm0.08$ \\
$5$ & $1.4\pm0.6$ & $0.1\pm0.2$ \\
$6$ & $1.5\pm0.7$ & $0.0\pm0.1$ \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{fig5.pdf}
\end{center}
\caption{
Dependence on the spatial dimension and approach to the
mean-field limit.
Size dependence of the number of bridge bonds, $N_{BB}$, in
the limit $p=1$, up to dimension six. The solid lines stand for the
best fit. The mass of the shortest path in loopless percolation (SP) in
$2D$ is also included for comparison. Results have been averaged over
$10^4$ samples for $2D$ and $3D$, and $10^2$ samples for higher
dimensions. Error bars are smaller than the symbol size. The inset
shows the dependence on the spatial dimension of the exponent $\varphi$.
At the upper-critical dimension of percolation, $d=6$, the set of bridge
bonds becomes dense having the spatial dimension.
\label{fig::nbbdim}}
\end{figure}
Since one can interchange occupied and empty bonds, there exists a
symmetry between bridges and cutting bonds (red bonds), so one can raise
the question of what happens when bonds are removed from a percolating
system with the constraint that connectivity cannot be broken.
Initially all bonds are occupied and at the end, since cutting bonds are
never removed, a line of cutting bonds is obtained, which we denote here as
the {\it cutting line}. Above $p_c$, as in the classical case, the percolation
cluster is compact and there are no cutting bonds. For $p<p_c$, the set
of cutting bonds is fractal with the same fractal dimension as the
bridge-bond set, $d_{CB}=d_{BB}$, whereas at $p_c$ it is $1/\nu$. For
the crossover scaling a similar \textit{ansatz} to the one given by
equation~(\ref{eq::scaling}) is verified, where the argument of the
scaling function is then $(p_c-p)L^\theta$. The same hyperscaling of
equation~(\ref{eq::scaling.relation}) is obtained with
$\varphi=d-d_{CB}$. At $d=2$, our numerical results corroborate the
hypothesis of the same $\zeta$ and $\varphi$, for cutting and bridge
bonds (see {\it Appendix}). For $d>2$, the set of
cutting bonds is a line with $d_{CB}\leq2$ and the one of bridge bonds
has a dimension above $d-1$, so that the fractal dimensions differ.
Above the critical dimension, the set of cutting bonds has dimension
two, like the shortest path at $p_c$ \cite{Havlin84}, for any $p\leq
p_c$.
Fisher \cite{Fisher61} proposed a bond-site transformation to map bond
percolation on a lattice onto site percolation on a covering graph. For
example, the covering graph of the square lattice is obtained by adding
all diagonal edges to every other face \cite{Wierman95}. Considering
such a mapping for ranked percolation and, following the analogy with a
random landscape, where sites are sequentially removed from the lowest
number (height) and suppressing global disconnection, the obtained
cutting line is identical to the bridge one with the constraint applied
in the perpendicular direction. The same is also observed for
site-ranked percolation on a square lattice. Given the relation between
connectivity and disconnection, in both cases, the cutting version needs
to be defined on different topologies, namely, the star lattice for the
square (on sites) and the same lattice for its covering graph, but with
swapped cells of diagonal bonds. For the triangular lattice, the
relation between cutting and bridges sites is straightforward without
requiring additional connections.
\begin{figure}
\includegraphics[width=\columnwidth]{fig1supmat.pdf}
\caption{
Following the rank, bonds are sequentially removed except
the cutting bonds.
Size dependence of the number of cutting bonds, $N_{CB}$, for
different fractions of occupied bonds $p$, namely
$p=p_c=0.5$ (circles), $p=p=0.49$
(stars), and $p=0.2$ (triangles). The solid lines are guides to the
eye. At $p=p_c$, $N_{CB} \sim
L^{1/\nu}$, where $\nu = 4/3$ is the critical exponent related with
the correlation length. For $p<p_c$, the number of cutting bonds
scales with $L^{d_{CB}}$. A crossover between the two regimes with the
system size is observed (stars) for $p$ in the neighborhood
of $p_c$. Systems of size $L^2$ have been considered, with $L$ ranging
from $8$ to $512$. Results have been averaged over $10^3$ samples for
most system sizes and over $50$ samples for the largest one. Error
bars are smaller than the symbol size.
The inset shows the number of cutting bonds, $N_{CB}$, as a function
of occupied bonds, $p$, for two-dimensional lattices with $L$ ranging
from $64$ to $256$ (averaged over $10^2$ samples).
The scaling is given by Eq.~($1$) in the article, where the argument is
then $(p_c-p)L^\theta$, with $\theta=0.93$.
We obtain $\zeta=0.56\pm0.08$, which is compatible with the one for bridge bonds.
\label{fig::cb}}
\end{figure}
\section{Discussion}
For several different models in 2D the same fractal dimension as the one
of bridge bonds has been reported. Here we discuss how some of them can
be related with the process of fracturing ranked surfaces. Let us
construct ranked percolation configurations on a random landscape
starting with the bond with the largest number (height) and then
occupying bonds sequentially in order of decreasing number. Each time a
chosen bond closes a loop, it is not occupied (loopless). One stops the
procedure when, for the first time, a path of connected sites spans from
one side to the other (percolation threshold). Since all clusters are
trees, the backbone and the shortest path are the same and their fractal
dimension is identical to the bridge-bond line, as verified in
Fig.~\ref{fig::nbbdim}. In fact, given the connectivity/disconnection
relation discussed in the previous section, this line corresponds to the
bridge line when bonds are occupied from the lowest to the largest
number in the covering graph. This line also corresponds to the line of
cutting bonds when bonds are removed from the smallest to the largest
number. A similar procedure was also proposed to obtain the minimum
spanning tree (MST), for which the same fractal dimension is found for
the paths between all pairs of sites \cite{Dobrin01}. There bonds are
sequentially removed (from the highest to the lowest value) under the
constraint that all sites remain connected. If the constraint is relaxed
such that only connectivity between two opposite borders is imposed, the
cutting-bond line is obtained. Therefore, the cutting-bond line is the
path between borders on the MST for which the largest value (height) is
minimum, a valid relation in any spatial dimension $d$.
\begin{figure}
\includegraphics[width=\columnwidth]{fig2supmat.pdf}
\caption{
Size dependence of the number of bridges, $N_B$, on several different lattices.
The solid lines are guides to the eye, all with slope $1.215$.
The best fit to each data set gives the same fractal dimension of $1.215\pm0.005$.
Systems of linear size $L$ have been considered, with $L$ ranging from $32$ to $4096$.
Results have been averaged over $10^5$ samples.
\label{fig::difnet}}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{fig3supmat.pdf}
\caption{
Number of bridge bonds, $N_{BB}$, as a function of the fraction
of occupied bonds, $p$, for $3D$ (simple-cubic lattices) with different sizes $L=\{256,512,1024\}$.
The scaling function proposed in this article is applied, with $\theta=1.36$, obtaining $\zeta=1.0\pm0.1$.
In the inset, $N_{BB}$ has been rescaled by $L^{d_{BB}}$, where $d_{BB}=2.50$.
All results have been averaged over $10^4$ samples.
\label{fig::bbscale3d}}
\end{figure}
The fractal dimension $d_{BB}$ was also observed for the backbone of the
optimal path crack (OPC) \cite{Andrade09,Oliveira11}. For a random
landscape, the sequence of optimal paths between opposite borders is
obtained and their highest site removed. Every such path crosses the
bridge line. The highest site on successive paths is either on the
bridge line itself or is higher than the lowest bridge-line site.
Therefore, as in the loopless percolation described before, the removed
sites percolate when all sites on the bridge line are removed, which is
the backbone of the crack, giving the same fractal dimension $d_{BB}$.
This is in fact the case for the crack of any sequence of self-avoiding
paths between opposite borders. Since in this case the duality between
cutting and bridges is not used (only valid in 2D), the relation between
OPC and bridge bonds is still true in higher dimensions.
Let us now define the optimal minimax path (OMP) in the following way.
One starts selecting the set of paths on the landscape for which the
highest-ranked site is minimum (minimax paths). Such set is then reduced
to include only the paths for which the second highest site is also
minimum and one proceeds iteratively to the following sites, until a
unique path is obtained. This path is the optimal path in strong
disorder \cite{Porto97,Porto99}, since under such disorder strength each
site is higher than the sum over the height of all sites with lower
rank. This path is also identical to the backbone of the
loopless-ranked-percolation cluster when occupied from the lowest to the
largest number and, therefore, the cutting line in any dimension.
In summary, suppressing connectivity (disconnection) between opposite
borders on ranked surfaces leads to a fractal set of bridge (cutting)
bonds, with a universal fractal dimension, even far away from the critical
point of classical percolation. In $2D$, there is an
equivalence between bridges and cutting bonds and $d_{BB}=d_{CB}$,
whereas for $d>2$ cutting bonds are still a fractal line but bridge
bonds form a surface. The discussed models are then split into two groups. The
ones suppressing connectivity (e.g., watershed \cite{Cieplak94,Fehr09}
and optimal path crack \cite{Andrade09,Oliveira11}) are in the bridges
universality class, while the ones keeping connectivity (e.g., optimal
path on the MST or in strong disorder media \cite{Porto97}) are in the
universality class of cutting bonds. For $d>6$, $d_{BB}=d$ and
$d_{CB}=2$, so we conjecture that the upper-critical dimension of
bridges and cutting bonds is also $d_c=6$. Finally, we show that, at the
percolation threshold of classical percolation, ranked percolation
displays a theta-point-like crossover.
\begin{figure}
\includegraphics[width=\columnwidth]{fig4supmat.pdf}
\caption{
Size of the largest cluster of sites on the edge of bridge bonds, $M_{BB}$, rescaled by $L^{d_{BB}}$, as a function of the fraction of occupied bonds $p$.
The inset shows $\chi$ defined by Eq.~(\ref{eq::def.chi}).
Results have been obtained on a square lattice with $L$ ranging from $256$ to $4096$.
All results have been averaged over $10^4$ samples.
\label{fig::bbp}}
\end{figure}
This work opens up several challenges. Besides the need for a more
precise numerical estimation of the bridge-growth exponent, it would be
interesting to formulate a renormalization group scheme and to obtain
the new set of exponents from analytic treatments such as, e.g., exact
results in the mean-field limit and a Schramm-Loewner evolution in two
dimensions. Another interesting possibility is to find the corresponding
exponents in other related problems with different universality classes
like, e.g., the Kasteleyn-Fortuin clusters in the $q$-state Potts model
\cite{Coniglio89,Fortuin72,Wu78,Coniglio80,Deng04} with or without
magnetic field. Regarding the fractal dimension of the surface of
discontinuous percolation clusters \cite{Araujo10,Schrenk11}, it would
also be interesting to understand how it relates with the herein
introduced bridge-bonds universality class. For the cutting bonds, the
study of the crossover for higher dimensions represents another computational
challenge. Finally, it would also be interesting to try to identify the
third scaling field of our theta-like point.
\section{Methods}
All numerical results have been obtained with Monte Carlo simulations.
Results have been averaged over $10^4$ samples for $2D$ and $3D$, and
$10^2$ samples for higher dimensions.
\section{Author contributions}
K.J.S. and N.A.M.A. carried out the numerical experiments. All authors
conceived and designed the research, analyzed the data, worked out the
theory and wrote the manuscript.
\begin{acknowledgments}
We acknowledge financial support from the ETH
Risk Center. We also acknowledge the Brazilian agencies CNPq, CAPES and
FUNCAP, and the Pronex grant CNPq/FUNCAP, for financial support.
\end{acknowledgments}
|
\section{Introduction}
Wiener's Tauberian Lemma \cite{W32} is a classical result in harmonic analysis which states that if a periodic function $f$ has an absolutely convergent Fourier series and never vanishes
then the function $1/f$ also has an absolutely convergent Fourier series. This result has many extensions (see \cite{ABK08, B08, BK10, B90, B97Sib, B97Izv, GKW89, GK10, GL06, J90, K11, K90, L53, S95, S78, S05, S07} and references therein),
which have been used, for example, in such diverse areas as differential equations \cite{K99}, pseudo-differential operators \cite{GR08, GS07},
frame theory \cite{ABK08, BCHL06I, F09, G04, KO08, S07, S08ACM}, time-frequency analysis \cite{GL04, GL06, KO08}, sampling theory \cite{AAK09, ABK08, SS09, S10}, finite-section method \cite{GRS10, RRS98}, etc.
A standard reformulation of Wiener's lemma states that if an invertible operator is defined by a (bi-infinite) Laurent matrix with
summable diagonals then the inverse has the same property. In \cite{B90, GKW89, K90} it was shown (independently) that the matrix
does not have to be Laurent. Hence, if one interprets a matrix entry $a_{ij}$ as a ``memory cell'', that is information on what impact an event at ``time'' $j$ has on the state of the system at ``time'' $i$, the Wiener property for an operator roughly means that such an operator does not spread the information too far or has a localized memory. In \cite{K11} one of us showed
that Wiener's lemma
is really a statement about the preservation of memory localization by inverse operators, and the way the memory is defined is
irrelevant to a large extent. In this paper we continue and expand this line of research paying particular attention to the case of ``continuous
memory'' and obtaining specific memory estimates for the inverse operators.
The following Wiener's lemma extension developed in \cite{B97Izv} serves as a starting point and an example for the research in this paper.
Let $\mathcal{X}, \mathcal{Y}$ be infinite dimensional complex Banach spaces and ${\mathcal P}=({\mathcal P}_k)$ and ${\mathcal Q} =({\mathcal Q}_k)$, $k\in S\subseteq{\mathbb Z}$, be two sequences of (continuous) idempotents from the Banach algebras $B(\mathcal{X})$ and $B(\mathcal{Y})$ of bounded linear operators on $\mathcal{X}$ and $\mathcal{Y}$, respectively. We assume that each of the sequences is disjunctive, i.e. ${\mathcal P}_k{\mathcal P}_j = \delta_{jk}{\mathcal P}_k$, $k,j\in S$, and similarly for the sequence ${\mathcal Q}_k$, $k\in S$. As usual, by $\delta_{jk}$ we mean the Kronecker delta. We also assume that for every $x\in \mathcal{X}$ and $y\in\mathcal{Y}$ we have
\[x = \sum_{k\in S} {\mathcal P}_kx\quad\mbox{ and }\quad y = \sum_{k\in S} {\mathcal P}_ky,\]
where the series converge unconditionally, and that
\begin {equation}\label{crp}
C({\mathcal P}) = \sup_{\a_k\in{\mathbb T}}\norm{\sum_{k\in S} \a_k{\mathcal P}_k}<\infty\mbox{ and }
C({\mathcal Q}) = \sup_{\a_k\in{\mathbb T}}\norm{\sum_{k\in S} \a_k{\mathcal Q}_k}<\infty,\end{equation}
where ${\mathbb T} = \{\gamma\in{\mathbb C}:\, |\gamma|=1\}$ is the unit circle.
A sequence of idempotents with the above properties will be called a \emph{disjunctive resolution of the identity}. Given such ${\mathcal P}$ and ${\mathcal Q}$ one can define a matrix $\mathcal A = (A_{mn})$ for each operator $A$ in the vector space $L(\mathcal{X},\mathcal{Y})$ of bounded linear operators from $\mathcal{X}$ to $\mathcal{Y}$. The matrix is a mapping
$\mathcal A: S\times S\to L(\mathcal{X},\mathcal{Y})$ given by
\[\mathcal A(m,n) =A_{mn} = {\mathcal Q}_mA{\mathcal P}_n,\ m,n\in S,\]
and the operators $A_{mn}$, $m,n\in S$, are called \emph{operator blocks}. If $A$ is continuously invertible, then the operator blocks of the operator $B = A^{-1}\in L(\mathcal{Y},\mathcal{X})$ are defined by
$B_{mn} = {\mathcal P}_mB{\mathcal Q}_n$, $m,n\in S$.
In \cite{B97Izv} the class of operators with summable diagonals, which in this paper we will denote by $\mathcal{W}$, was defined using the sequence
\[d_A(k) = \sup_{m-n=k}\norm{A_{mn}},\ k\in S-S,\, \mbox{ and } d_A(k)=0,\mbox{ if } k\in{\mathbb Z}\backslash (S-S).\]
More precisely, we call
\[\mathcal{W} = \mathcal{W}(\mathcal{X},\mathcal{Y}) = \{A\in L(\mathcal{X},\mathcal{Y}):\, \sum_{k\in{\mathbb Z}}d_A(k)< \infty\}\]
the \emph{Wiener class} of operators (with respect to the resolutions of the identity ${\mathcal P}$ and ${\mathcal Q}$).
Wiener's lemma extension proved in \cite{B97Izv} states that if $A\in \mathcal{W}(\mathcal{X},\mathcal{Y})$ is
invertible, that is
$B = A^{-1}\in L(\mathcal{Y},\mathcal{X})$, then $B\in\mathcal{W}(\mathcal{Y},\mathcal{X})$.
\begin {rem}
It was shown in \cite{K11} that if $\mathcal{X}=\mathcal{Y}={\mathcal H}$ is a Hilbert space, then the result remains valid if the resolutions of the identity ${\mathcal P}$ and ${\mathcal Q}$ are not assumed to be disjunctive.
\end {rem}
Although we have avoided mentioning it until now, the key technique for estimating the norms of the operator blocks of the inverse operator in \cite{B97Izv} is based on the spectral properties of the representations of locally compact Abelian (LCA-) groups.
Indeed, the resolutions of the identity ${\mathcal P}$ and ${\mathcal Q}$ give rise to two $2\pi$-periodic representations
\[U_\mathcal{X}:{\mathbb R}\to B(\mathcal{X}) \mbox{ and } U_\mathcal{Y}:{\mathbb R}\to B(\mathcal{Y})\]
defined by
\begin {equation}\label{exres}
U_\mathcal{X}(t) x = \sum_{k\in S} e^{ikt}P_kx,\ U_\mathcal{Y}(t) y = \sum_{k\in S} e^{ikt}Q_ky, \ x\in\mathcal{X}, y\in\mathcal{Y}, t\in{\mathbb R}.
\end{equation}
The above representations, in turn, produce another $2\pi$-periodic representation
$U_{\mathcal{X}\mathcal{Y}}: {\mathbb R}\to B(L(\mathcal{X},\mathcal{Y}))$ via
\begin {equation}\label{typop1}
U_{\mathcal{X}\mathcal{Y}} (t)A= U_\mathcal{Y}(t)AU_\mathcal{X}(-t).
\end{equation} The \emph{Fourier series} of the operator $A$ is then given by
\[U_{\mathcal{X}\mathcal{Y}}(t) A \simeq \sum_{k\in{\mathbb Z}} e^{ikt}A_k,\]
where the \emph{Fourier coefficients} satisfy \cite{B97Izv}
\[A_k = \frac1{2\pi}\int_0^{2\pi} e^{-ikt}U_{\mathcal{X}\mathcal{Y}}(t) A \,dt = \sum_{m-n = k} {\mathcal Q}_mA{\mathcal P}_{n},\ k\in{\mathbb Z}.\]
The above integral converges in the strong operator topology and the series converges strongly and unconditionally.
If $C({\mathcal P}) = C({\mathcal Q}) = 1$, which is always true for some equivalent renormalization of $\mathcal{X}$ and $\mathcal{Y}$, we have $d_k(A) = \|A_k\|$ and the connection between the Wiener's class $\mathcal{W}$ and summability of Fourier coefficients becomes apparent.
Often it is natural to refer to the Fourier coefficients of an operator as its \emph{diagonals} and the set of indices of non-zero diagonals as \emph{memory}. Obviously both notions are dependent upon the choice of the representation.
Although the paper \cite{B97Izv} deals only with representations \eqref{exres}, it is easily seen that the technique works for any periodic representation (or, equivalently, a representation of the LCA-group ${\mathbb T}$).
In \cite{BK10} it is shown how to prove analogous results in the almost periodic case (for representations of the Bohr compactification ${\mathbb R}_c$ of ${\mathbb R}$). In this paper we are primarily interested in theory based on representations of ${\mathbb R}^d$, $d\in{\mathbb N}$. In this case, memory is no longer a discrete set, the representations in \eqref{exres} are usually not well defined, and we have to use far more sophisticated methods of the spectral theory
of Banach modules (representations of LCA-groups). Fortunately, most of the required methods
have already been developed in \cite{B04, BK05}.
The remainder of the paper is organized as follows. In section \ref{nota1} we summarize the
notions and results of the spectral theory of representations of LCA-groups.
In Section \ref{Cla} we introduce different classes of vectors with spectral decay such as Wiener Class, Beurling Class and others. We believe that some of the classes (one-sided exponential spectral decay, Sobolev-type classes) have not been previously considered in conjunction with Wiener's lemma type results. We also don't believe that any of these classes have been introduced at this level of generality.
Section \ref{Clop} introduces standard \cite{B04, BK05, BR75} Banach module structures on the space of operators which allow us to represent memory decay of linear operators as the spectral decay of vectors in Banach modules. In Sections \ref{Wcomp}--\ref{sobol} we prove results about the memory decay of inverses to operators in different classes. The main Theorems are \ref{1sidexpwin}, \ref{invexp}, \ref{maint}, \ref{absB}, and \ref{sobolt}. It is worth noting here that some of these results are quantitative. While we use a combination of traditional methods of proof
(holomorphic extensions, Neumann series ``boot-strap'', Brandenburg trick, etc. \cite{G10}),
we also employ a closed operator functional calculus that has not been used before. More importantly,
the framework of Banach modules (group representations) makes our results extremely general as illustrated by
a variety of examples in Section \ref{examp}. We show that many of the previously known results of this kind can be obtained using our technique. On the other hand, an absolute majority of these examples cannot be obtained at this level of generality in other frameworks known to us. In particular, the block-matrix technique of \cite{B97Izv} fails because
in general the representation in \eqref{exres} is not well defined (the series may not converge unconditionally). Other matrix-based techniques \cite{S05, S07, S10ca} do not apply essentially for the same reason. Known results for integral operators \cite{K99, K01, S08} apply to narrower classes of operators. Finally,
the abstract technique based on the Bochner-Phillips theorem \cite{B97Sib, BP42} has either not been developed for this level of generality or
cannot be used directly, because we consider operators in $L(\mathcal{X},\mathcal{Y})$ as well as Banach algebras with a group of automorphisms.
\section{Banach modules over weighted group algebras and their spectral properties}\label{nota1}
In this section we introduce the notation and develop the necessary tools from representation theory. The presentation is largely
analogous to that of \cite{BK05}, where isometric representations are used. Our interest in differential operators, however, forces
the use of unbounded representations, which requires an extra effort. Although, most of the results
in Sections \ref{Wcomp}--\ref{sobol} are formulated for bounded representations we prefer to
present the prerequisite theory more generally in order to use it in the sequel to this paper.
As above, $\mathcal{X}$ will denote a complex Banach space and $B(\mathcal{X})$ will be the Banach algebra of
bounded linear operators on $\mathcal{X}$. By $\mathbb G $ we will denote an LCA-group
and by $\widehat{\G}$ -- its Pontryagin dual, the group of continuous unitary characters of $\mathbb G $.
We write the operation additively on all LCA-groups, except when the group ${\mathbb T} = \{\theta\in{\mathbb C}:\ |\theta| = 1\}$ is used.
Next, we introduce the following definition of weights.
\begin{defn} \label{wegt}
A \emph{weight} is an even function $\nu: \mathbb G \to [1,\infty)$ such that
\[ \nu(g_1+g_2)\leq \nu(g_1)\nu(g_2),\ \mbox{ for all }\ g_1,g_2\in\mathbb G . \]
A weight is \emph{admissible} if it satisfies the \emph{GRS-condition}
\[\lim\limits_{n\to\infty}n^{-1}\ln\nu(ng)=0,\quad \mbox{ for all }\quad g\in\mathbb G ,\quad ng = \underbrace{g+g+\ldots+ g}_{n\, {\rm times}}. \]
A weight is \emph{exponential} if $\mathbb{G}$ is a subgroup of ${\mathbb R}^d$ (possibly, with discrete topology) and
$\nu(g) = e^{a|g|}$ for some $a > 0$.
For
$g = (g_1,g_2,\ldots,g_d)\in{\mathbb R}^d$ we will use the notation $|g|_p = \left(\sum\limits_{k=1}^d |g_k|^p\right)^{\frac1p}$, $1\le p< \infty$, $|g| = |g|_1$, and $|g|_\infty = \max\limits_{1\le k\le d} |g_k|$.
\end{defn}
We denote by $L_\nu(\mathbb G )$ the Banach algebra of (equivalence classes of) complex functions
on $\mathbb G $ integrable with weight $\nu$ with respect to the Haar measure on $\mathbb G $. If $\nu \equiv 1$, we obtain the standard space $L^1(\mathbb G ) \equiv L_1(\mathbb G )$. We shall use both notations interchangeably, hinting at the possibility of using more general weights when the subscript is used. The role of
multiplication is played by the convolution of functions and the norm is given by
\[\norm{f}_\nu = \int_\mathbb G |f(g)|\nu(g)dg .\]
We refer to \cite{K09, L85} for the various properties of the algebra $L_\nu(\mathbb G )$ used below.
In this paper we assume that the weight $\nu$ is non-quazi-analytic \cite{D56, LMF73, L85}, that is
\[\sum_{n=0}^\infty \frac{\ln \nu(ng)}{1+n^2} < \infty \quad \mbox{for all } g\in\mathbb G .\]
This condition ensures \cite{D56, K09} that the group algebra $L_\nu(\mathbb G )$ is regular and its spectrum
is isomorphic to the dual group $\widehat{\G}$.
We denote by $\hat{f} : \widehat{\G} \to {\mathbb C}$ the Fourier transform of $f \in L_\nu(\mathbb G )$. The inverse Fourier
transform of a function $h: \widehat{\G} \to {\mathbb C}$ is denoted by $\check h$ or $h^\vee$. If $\mathbb{G} = {\mathbb R}^d$, we use the Fourier transform in the form
\[\hat f(\l) = \int_{{\mathbb R}^d} f(t)e^{-it\cdot\l} dt,\ \l\in{\mathbb R}^d.\]
The memory of the operators will eventually be defined using the properties of group representations
$\mathcal{T}: \mathbb G \to B(\mathcal{X})$. Together with the representation in \eqref{exres}, the following representations are used most often.
If $\mathcal{X}$ is an appropriate space of functions (or distributions) on $\mathbb G $, the \emph{translation} representation $\mathcal{T} = T: \mathbb G \to B(\mathcal{X})$ is given by
\begin {equation}\label{trans1}
T(t)x(s) = x(s+t),\ x\in\mathcal{X}, \ s,t\in\mathbb G ,
\end{equation}
and the \emph{modulation} representation $\mathcal{T} = M: \widehat{\G} \to B(\mathcal{X})$ is defined by
\begin {equation}\label{mod1}
M(\xi)x(s) = s(\xi)x(s),\ x\in\mathcal{X}, \ \xi\in\widehat{\G}, \ s\in \mathbb G .
\end{equation}
Other important examples of representations used in the paper are introduced below.
We assume that $\mathcal{X}= (\mathcal{X},\mathcal{T})$ is a non-degenerate Banach $L_\nu(\mathbb G )$-module \cite{B04, BK05},
with the module structure associated with a representation $\mathcal{T}: \mathbb G \to B(\mathcal{X})$.
In particular, in this paper we consider only modules for which the following assumption is satisfied.
\begin{ass}\label{assu}
The following three conditions hold for the Banach $L_\nu(\mathbb G )$-module $\mathcal{X}$:
\begin{enumerate}
\item the module $\mathcal{X}$ is non-degenerate, that is, if $fx = 0$ for all $f \in L_\nu(\mathbb G )$ then $x = 0$;
\item the module structure on $\mathcal{X}$ is associated with a representation $\mathcal{T}: \mathbb G \to B(\mathcal{X})$, that is,
for all $f \in L_\nu(\mathbb G )$, $x \in \mathcal{X}$, and $g\in\mathbb G $,
\begin {equation}\label{redef}
\mathcal{T}(g)(fx) = \left(T(g)f\right)x = f\left(\mathcal{T}(g)x\right),
\end{equation}
where $T$ is the translation representation (\ref{trans1});
\item
$\norm{fx} \leq \|f\|_\nu\|x\|$, for all $f \in L_\nu(\mathbb G )$ and $x \in \mathcal{X}$.
\end{enumerate}
\end{ass}
\begin {rem}\label{cnepr}
We observe that to ensure that $x=0$ it is enough to show that $fx = 0$
for every $f\in L_\nu(\mathbb{G})$ with $\text{ supp\!} \hat f$ compact. Indeed, since every $h\in L_\nu(\mathbb{G})$
is a limit of a net $(f_\a)$ of functions with $\text{ supp\!}\hat f_\a$ compact \cite{K09, L85}, we would get
\[hx = \lim_{\a} (h*f_\a)x = 0\]
and, hence, $x = 0$ since $\mathcal{X}$ is non-degenerate.
\end {rem}
If $\mathcal{T}$ is a \emph{strongly continuous} representation, i.e. the function $x_\mathcal{T}: \mathbb G \to\mathcal{X}$ given by
$x_\mathcal{T}(g) =\mathcal{T}(g)x$ is continuous for every $x\in\mathcal{X}$, and $\nu(g) \geq \norm{\mathcal{T}(g)}$ for a weight $\nu$, then
a non-degenerate Banach $L_\nu(\mathbb G )$-module structure on $\mathcal{X}$ satisfying assumption \ref{assu} may be defined via
\begin {equation}\label{scdef}
fx = \mathcal{T}(f)x = \int_\mathbb G f(g)\mathcal{T}(-g)x dg,\quad f\in L_\nu(\mathbb G ), \ x\in\mathcal{X}.
\end{equation}
It is easy to show (\cite[Lemma 2.2]{BK05}) that every non-degenerate $L_\nu(\mathbb G )$-module has at most one representation
associated with it. Therefore, if $\mathcal{T}$ is strongly continuous, the module structure has to be defined via (\ref{scdef}). If $\mathcal{T}$ is not strongly continuous, we shall make use of the submodule $\mathcal{X}_{c}\subset \mathcal{X}$ of $\mathcal{T}$-\emph{continuous} vectors of $\mathcal{X}$. We let $x\in\mathcal{X}_c$ if the function
$x_\mathcal{T}$ is continuous.
The symbol $\mathcal{T}(f)$ in \eqref{scdef} may seem like an abuse of notation. We justify this by regarding $\mathcal{T}$ also as a representation of the algebra $M_\nu(\mathbb{G})$ of finite complex Borel measures on $\mathbb{G}$ with respect to convolution and
the norm
\[\|\mu\|_\nu = \int_\mathbb{G} \nu(g) d|\mu|(g).\]
The group $\mathbb{G}$ is then identified with Dirac measures in $M_\nu(\mathbb{G})$ in an appropriate way and $L_\nu(\mathbb{G})$ -- with the algebra of measures in $M_\nu(\mathbb{G})$ that are absolutely continuous with respect to the Haar measure on $\mathbb{G}$. For $x \in\mathcal{X}_c$ we then have
\begin {equation}\label{tmes}
\mu x = \mathcal{T}(\mu)x = \int_\mathbb G \mathcal{T}(-g)x d\mu(g),\quad \mu\in M_\nu(\mathbb G ), \ x\in\mathcal{X},
\end{equation}
and $\|\mu x\|\le\|\mu\|_\nu\|x\|$. We refer to \cite{BK05} for more details.
\begin {rem}\label{tenp}
We observe that if $\mathbb G ={\mathbb R}^d$ and we have an $L_1({\mathbb R}^d)$-module structure associated with a representation $\mathcal{T}:{\mathbb R}^d\to B(\mathcal{X})$ we also have $d$ module structures over $L_1({\mathbb R})$ associated with the representations $\mathcal{T}^{(k)}: {\mathbb R}\to B(\mathcal{X})$, $k=1,\ldots, d$, given by
\[\mathcal{T}^{(k)}(t) x = \mathcal{T}(0,\ldots, 0,t,0,\ldots, 0)x,\ x\in\mathcal{X},\]
where $t$ is the $k$-th component of the vector. Obviously, in this case,
\[\mathcal{T}(t_1,\ldots, t_d) = \prod_{k=1}^d \mathcal{T}^{(k)}(t_k)\quad\mbox{and}\quad
\mathcal{T}(f_1\otimes\ldots\otimes f_d) = \prod_{k=1}^d \mathcal{T}^{(k)}(f_k),\]
$t_k\in{\mathbb R}$, $f_k\in L_1({\mathbb R})$, $k=1,\ldots, d$. A similar observation is clearly valid in the case of tensor-product weights and Cartesian products of arbitrary LCA-groups.
\end {rem}
\subsection{Spectral properties} The role of memory of operators on Banach modules is fulfilled by the notion of the Beurling spectrum. We begin by defining this notion for vectors in Banach modules.
\begin {defn}
The \emph{Beurling spectrum} of a set $N \subseteq (\mathcal{X},\mathcal{T})$ is the subset $\L(N)= \L(N,\mathcal{T})$ of the dual group $\widehat{\G}$ the compliment of which is given by
\[\{\gamma\in\widehat{\G}: \mbox{ there is } f\in L_\nu(\mathbb{G}) \mbox{ such that } \hat f(\gamma) \ne 0 \]
\[\mbox{ and } fx =\mathcal{T}(f)x = 0 \mbox{ for all } x\in N\}.\]
\end {defn}
When $N = \{x\}$ is a singleton, we shall write $\Lambda(x)$ instead of $\L(\{x\})$.
Given a closed set $\sigma\subset \widehat{\G}$ we shall denote by $\mathcal{X}(\sigma)$ the (closed) \emph{spectral submodule} of
all vectors $x\in\mathcal{X}$ such that $\L(x)\subseteq \sigma$. The symbol $\mathcal{X}_{Comp}$ will stand for the set
of all vectors such that $\L(x)$ is compact.
\begin {rem}\label{otherspec}
The notion of Beurling spectrum has a lot of very different guises (see \cite{B04, BR75, D56, EN06} among many others). All of these were shown to be equivalent at the level of generality of this paper \cite{B04}; we especially note the equivalent notions of the spectrum used in \cite{GK10} and in
\cite[Theorem XI.11.24]{DS88II}.
\end {rem}
Observe that if $\mathcal{X}$ is an appropriate space of functions on $\mathbb{G}$ and $\mathcal{T}=T$ is defined via
\eqref{trans1}, then $T(f)x = f*x$ and $\L(x) = \text{ supp\!} \hat x$, $x\in\mathcal{X}$, $f \in L_\nu(\mathbb{G})$, possibly, in the sense of distributions.
Alternatively, if $\mathcal{X}$ is an appropriate space of functions on $\widehat{\G}$ and $\mathcal{T}=M$ is defined via
\eqref{mod1}, then $M(f)x = \hat f x$ and $\L(x) = \text{ supp\!} x$, $x\in\mathcal{X}$, $f \in L_\nu(\mathbb{G})$.
If $\mathcal{T} = \mathcal{T}_\mathcal{X}$ is defined via a resolution of the identity as in \eqref{exres}, then
\[\L(x) = \{n\in{\mathbb Z}: {\mathcal P}_n x\ne 0\}.\]
In the next lemma we present basic properties of the Beurling spectrum that will be heavily used throughout the paper. We refer to \cite{B04, BK05} and references therein for the proof.
\begin {lem}\label{sprop}
Let $\mathcal{X}$ be a non-degenerate Banach $L_\nu(\mathbb{G})$-module with the structure associated with a
representation $\mathcal{T}$.
Then
\begin{description}
\item[ (i)] $\L(M)$ is closed for every $M \subseteq \mathcal{X}$ and $\L(M)=\emptyset$ if and only if $M =\{0\}$;
\item[(ii)] $\L(Ax + By) \subseteq \L(x) \cup \L(y)$ for all $A$, $B \in B(\mathcal{X})$ that commute with all
operators $\mathcal{T}(f)$,
$f \in L_\nu(\mathbb{G})$;
\item[(iii)] $\L(f x)\subseteq (\text{ supp\!} \hat f)\cap \L(x)$ for all $f \in L_\nu(\mathbb{G})$ and $x\in\mathcal{X}$;
\item[(iv)] $f x=0$ if $(\text{ supp\!} \hat f)\cap\L(x)=\emptyset$, where $f \in L_\nu(\mathbb{G})$ and $x\in\mathcal{X}$;
\item[(v)] $f x = x$ if $\L(x)$ is a compact set, and $\hat f \equiv 1$ in some neighborhood of $\L(x)$, $f\in L_\nu(\mathbb{G})$, $x\in\mathcal{X}$;
\item[(vi)] if $M_0$ is dense in $M\subseteq X$, then $\L(M) = \overline{\bigcup_{x\in M_0}\L(x)}$.
\end{description}
\end {lem}
From the above lemma and examples we see that the Beurling spectrum $\L(x,\mathcal{T})$ is, indeed, well suited to represent the support of the ``Fourier transform'' of the function $x_\mathcal{T}(g)=\mathcal{T}(g)x$. The goal of this paper is to study the behavior of the decay of this ``Fourier transform'' (memory decay in the case of operators), relate it to
smoothness properties of the function $x_\mathcal{T}$, and, in particular, to prove Wiener-type theorems for $x$ when it is an invertible linear operator or an element in a Banach algebra. Appropriate module structures on the spaces of operators will be discussed in Section \ref{Clop}. Presently we shall introduce the notions of approximate identities and $\gamma$-nets which are essential for our analysis of the memory decay.
\subsection{\texorpdfstring{Approximate identities and $\gamma$-nets}{Approximate identities and gamma-nets}}
Bounded approximate identities (b.a.i.) and $\gamma$-nets are often used to obtain ergodic theorems in Banach modules \cite{B78, B88}. We shall use b.a.i.~ to approximate $\mathcal{T}$-continuous vectors with vectors that have compact Beurling spectrum and $\gamma$-nets to approximate the ``$\gamma$-Fourier coefficient'' of $x_\mathcal{T}$. We index all our b.a.i.~ and $\gamma$-nets by some net $\Omega$.
\begin {defn}
A bounded net $(f_\a)$ in $L_\nu(\mathbb{G})$ is called a \emph{bounded approximate identity} (b.a.i) in the algebra $L_\nu(\mathbb{G})$ if the following two conditions hold:
\begin{description}
\item[ (i)] $\hat f_\a(0) = 1$ for all $\a\in\Omega$.
\item[(ii)] $\lim f_\a*f = f$ for all $f\in L_\nu(\mathbb{G})$.
\end{description}
\end {defn}
The functions
\begin {equation}\label{znet}
f_\a(t) = \frac1{\pi^d}\prod_{k=1}^d \frac{\a_k}{t_k^2+\a_k^2},
\end{equation}
give an example of a b.a.i.~ in $L_1({\mathbb R}^d)$ when $\a \to 0^+$. A construction of b.a.i.~ in $L_1(\mathbb{G})$ with the compact support of the Fourier transform is presented, for example, in \cite{BK05}. Existence of b.a.i.~ in $L_\nu(\mathbb G )$ is, to the best of our knowledge, an open question; we refer to \cite[Lemmas 3.3.10, 3.4.9]{B04} and \cite{L85, SW71} for more details on the subject. Here we will note only that the algebra $L_\nu({\mathbb R}^d)$ with a non-quazi-analytic weight $\nu$ contains a b.a.i.~ if $\nu$ is majorized by a monotonic non-quazi-analytic weight $\tilde\nu$, that is
$\tilde\nu(u) \le \tilde\nu(v)$ whenever $|u|_2\le|v|_2$, $u,v\in{\mathbb R}^d$.
The following useful result is well-known \cite[and references therein]{B04, BK05, HR79}. There
by $\mathcal{X}_\Phi$ we mean the submodule of vectors admitting factorization: $\mathcal{X}_\Phi = \{x = fy\in\mathcal{X}, y\in\mathcal{X}, f\in L_\nu(\mathbb{G})\}$.
\begin {thm}\label{CoHe}
Let $(\mathcal{X},\mathcal{T})$ be a non-degenerate Banach $L_\nu(\mathbb{G})$-module. Then
$\mathcal{X}_{Comp} \subseteq \mathcal{X}_\Phi \subseteq {\mathcal{X}_{c}}$.
Moreover, if the algebra $L_\nu(\mathbb{G})$ contains a b.a.i.~ then
\[\mathcal{X}_c = \overline{\mathcal{X}_{Comp}} = \{x = \lim_\a f_\a x,\ f_\a \mbox{ -- a b.a.i.~ in } L_\nu(\mathbb{G})\}\]
and if there is a b.a.i.~ with $\|f_\a\|\le 1$ then also $\mathcal{X}_c = \mathcal{X}_\Phi$.
\end {thm}
We will need the following class of subsets of the module $\mathcal{X}_\Phi$.
\begin {defn}\label{modorbit}
Given $x\in (\mathcal{X}, \mathcal{T})$ the \emph{modular orbit} $\Omega(x)\subset \mathcal{X}_\Phi$ is the set
$\Omega(x) =\{fx,\ f\in L_\nu(\mathbb{G})\}$.
\end {defn}
Observe that by Domar's Theorem \cite{D56}, \cite[Ch. 5.1]{L85} we have that the set
$\{f\in L_\nu(\mathbb{G}):\, \text{ supp\!}\hat f \mbox{ compact}\}$ is dense in $L_\nu(\mathbb{G})$ and, therefore,
due to Lemma \ref{sprop}, for any open neighborhood $\mathcal U$ of $\L(x)$ we have
\begin {equation}\label{goodf}
\overline{\Omega(x)} = \overline{\{fx,\ f\in L_\nu(\mathbb{G}), \mbox{ with }\text{ supp\!}\hat f\subset \mathcal{U} \mbox{ compact}\}}.
\end{equation}
\begin {lem}\label{goodx}
If $x\in\mathcal{X}_c$ and $L_\nu(\mathbb{G})$ contains a b.a.i.~ then $x\in \overline{\Omega(x)}$.
\end {lem}
\begin {proof}
Follows from Domar's theorem and Theorem \ref{CoHe}.
\end {proof}
\begin {defn}\label{gnet}
For $\gamma\in\widehat{\G}$ a bounded net $(f_\a)$ in the algebra $L_\nu(\mathbb{G})$ is called a $\gamma$-net if the following two conditions hold:
\begin{description}
\item[ (i)] $\hat f_\a(\gamma) = 1$ for all $\a$;
\item[(ii)] $\lim f_\a * f = 0$ for every $f \in L_\nu(\mathbb{G})$ with $\hat f(\gamma) = 0$.
\end{description}
\end {defn}
The functions \eqref{znet} provide an example of a $0$-net in $L_1({\mathbb R}^d)$ as $\a\to \infty$. Other
examples of $\gamma$-nets in $L_1(\mathbb{G})$ are furnished in \cite{BK05}. Unfortunately, existence of such nets in $L_\nu(\mathbb{G})$ presents a formidable restriction on the growth of the weight $\nu$. In particular, in case $\mathbb G ={\mathbb R}$ it is known \cite{B04} that $\gamma$-nets do not exist already when $\nu$ is a linear weight.
The following proposition is a version of the results stated in \cite{B78, BK05}.
\begin {prop}
Assume that $(f_\a)$ is a $\gamma$-net and $x\in\mathcal{X}$ satisfies $0\ne x_\gamma = \lim f_\a x$. Then
$\L(x_\gamma) = \{\gamma\}$ and the limit does not depend on the particular choice of the $\gamma$-net.
\end {prop}
The above proposition shows why $\gamma$-nets are used to approximate ``Fourier coefficients'' of the function $x_\mathcal{T}$. Of course, in general the limit $x_\gamma$ does not exist. That is why in the following section (see \eqref{triangled}) we use elements
of a $\gamma$-net to define ``parts'' of a vector with the Beurling spectrum in a neighborhood of $\gamma$.
The following result uses $\gamma$-nets in its proof and allows us to sharpen some of the properties of the Beurling spectra we mentioned in Lemma \ref{sprop}.
\begin {lem}\label{ssprop}\cite[Lemma 3.7.32]{B04}.
Let
$\mathcal{X}$ be a non-degenerate Banach $L_1(\mathbb{G})$-module with the structure associated with a
representation $\mathcal{T}$.
Then
\begin{description}
\item[(i)] $f x=0$ if $(\text{ supp\!} \hat f)\cap\L(x)$ is countable and $\hat f(\gamma) = 0$ for all $\gamma \in (\text{ supp\!} \hat f)\cap\L(x)$,
$f \in L_1(\mathbb{G})$, $x\in\mathcal{X}$;
\item[(ii)] $f x = x$ if $\L(x)$ is a compact set, the boundary of $\L(x)$ is countable, and $\hat f \equiv 1$ on $\L(x)$, $f\in L_1(\mathbb{G})$, $x\in\mathcal{X}$.
\end{description}
\end {lem}
\subsection{Extended module structure}\label{closed}
In the following section we define various classes of spectral decay for vectors in Banach modules. To characterize some of these classes we need to extend the Banach module structure on $\mathcal{X}$ by means of a closed operator functional calculus.
In the following definition we use the spaces
$\widehat{L_{\nu}}(\mathbb G )=\{\hat\varphi$: $\varphi\in L_\nu(\mathbb G ) \}$ and
$\widehat{L_{\nu}}^{loc}(\mathbb G )=\{h$: $\widehat{\G} \to {\mathbb C}$ such that
$h\hat\varphi\in\widehat{L_{\nu}}(\mathbb G )$ for all $\varphi\in L_\nu(\mathbb G )$
with
$\text{ supp\!}{\hat\varphi}$ compact$\}$. Observe that $\widehat{L_{\nu}}(\mathbb G )\subset\widehat{L_{\nu}}^{loc}(\mathbb G )$.
For $h\in\widehat{L_{\nu}}^{loc}(\mathbb G )$ we define a (closed) operator $\check\mathcal{T}(h) = h\diamond : D(h)=D(h,\mathcal{T})\subseteq\mathcal{X}\to\mathcal{X}$ in the following way.
First, let $x\in\mathcal{X}_{comp}$ and
\begin {equation}\label{cl}
\check\mathcal{T}(h)x = h\diamond x : = (h\hat\varphi)^{\vee} x = \mathcal{T}((h\hat\varphi)^{\vee})x,
\end{equation}
where $\varphi\in L_\nu(\mathbb G )$ is such that $\text{ supp\!} \hat\varphi$ is compact and $\hat\varphi\equiv 1$ in a neighborhood of $\L(x)$.
The vector $\check\mathcal{T}(h)x$ is well-defined in this way because it is independent of the choice of $\varphi$. Indeed, if $\phi\in L_\nu(\mathbb G )$ is another function with the same properties, then
$h\hat\phi-h\hat\varphi\in\widehat{L_{\nu}}(\mathbb G )$ is $0$ in a neighborhood of $\Lambda(x)$ and, hence,
\[ \mathcal{T}((h\hat\phi)^{\vee})x= \mathcal{T}((h\hat\varphi)^{\vee})x\]
by \eqref{cl} and Lemma \ref{sprop}. The same lemma ensures also that $h\diamond x\in\mathcal{X}_{comp}$ whenever $x\in\mathcal{X}_{comp}$.
Next, we extend the definition of $\check\mathcal{T}(h)$ by taking the closure of the just defined operator on $\mathcal{X}_{comp}$. In other words,
if $x_n\in \mathcal{X}_{comp}$, $n\in{\mathbb N}$, $x = \lim\limits_{n\to \infty} x_n$, and
$y = \lim\limits_{n\to \infty} h\diamond x_n$ exists, we let $h\diamond x = y$.
In the following lemma we show that the above definition makes sense.
\begin {lem}
The operator $\check\mathcal{T}(h) = h\diamond : D(h)=D_{\mathcal{T}}(h)\subseteq\mathcal{X}\to\mathcal{X}$
is a well-defined closed operator.
\end {lem}
\begin {proof}
We need to prove that the restriction of $\check\mathcal{T}(h)$ onto $\mathcal{X}_{Comp}$ is closable.
In particular, we need to show that if
$x_n\in \mathcal{X}_{comp}$, $n\in{\mathbb N}$, $ \lim\limits_{n\to \infty} x_n = 0$, and
$y = \lim\limits_{n\to \infty} h\diamond x_n$ exists, then $y=0$. Consider an arbitrary
$f\in L_\nu(\mathbb G )$ with $\text{ supp\!}\hat f$ compact and let $\varphi\in L_\nu(\mathbb G )$ satisfy
$\hat\varphi \equiv 1$ in a neighborhood of $\text{ supp\!} \hat f$. Let also $\varphi_n\in L_\nu(\mathbb G )$ be such that $\hat\varphi_n \equiv 1$ in a neighborhood of $\L(x_n)$, $n\in{\mathbb N}$. Then
\begin{equation*
\begin {split}
fy & = \lim_{n\to\infty} f(h\diamond x_n) = \lim_{n\to\infty} (h\hat f\hat\varphi_n)^\vee x_n
= \lim_{n\to\infty} (h\hat\varphi_n)^\vee (fx_n)
\\ &= \lim_{n\to\infty} h\diamond (fx_n)
=\lim_{n\to\infty} (h\hat \varphi)^\vee (fx_n) =0.
\end{split}
\end{equation*}
By Remark \ref{cnepr}, this implies $y = 0$.
\end {proof}
From the definition we immediately have $\mathcal{T}(f) = \check\mathcal{T}(\hat f)$, $f\in L_\nu(\mathbb G )$.
We also observe the following slightly less obvious properties of the extended module structure.
\begin {prop}\label{spro} If $g\in\mathbb G $, $f\in L_\nu(\mathbb G )$, $h\in \widehat{L_{\nu}}^{loc}(\mathbb G )$, and $x\in D(h)$, then
$\mathcal{T}(g)x$, $\mathcal{T}(f)x \in D(h)$ and we have $\mathcal{T}(g)(h\diamond x) = h\diamond (\mathcal{T}(g)x))$ and
$\mathcal{T}(f)(h\diamond x) = h\diamond (\mathcal{T}(f)x))$.
\end {prop}
\begin {proof}
In case $x\in\mathcal{X}_{Comp}$, we get $\mathcal{T}(g)x$, $\mathcal{T}(f)x \in\mathcal{X}_{Comp}$ and the assertions are obvious. Otherwise, let $x =\lim\limits_{n\to\infty} x_n$ with $x_n\in\mathcal{X}_{Comp}$, $n\in{\mathbb N}$. Then,
by the definition of $\check\mathcal{T}(h)$ we have
\[h\diamond (\mathcal{T}(g)x)) = \lim_{n\to\infty} h\diamond(\mathcal{T}(g)x_n) = \lim_{n\to\infty} \mathcal{T}(g)(h\diamond x_n)
=\mathcal{T}(g)(h\diamond x)\]
and similar equations hold for $\mathcal{T}(f)(h\diamond x)$.
\end {proof}
Some basic properties of the Beurling spectrum outlined in Lemmas \ref{sprop}, \ref{ssprop} remain valid for the extended module structure in the following way.
\begin {lem}\label{sssprop}
Assume $\L(x)$ is compact, has countable boundary, and $\varphi, \psi \in \widehat{L_{\nu}}^{loc}(\mathbb G )$ satisfy
$\varphi\equiv\psi$ on $\L(x)$. Then $\check\mathcal{T}(\varphi)x = \check\mathcal{T}(\psi)x$.
\end {lem}
\begin {proof}
Let $f\in L_1(\mathbb G )$ satisfy $\hat f\equiv 1$ in a neighborhood of $\L(x)$. Applying
Lemma \ref{ssprop} to the function $h = (\hat f(\psi-\varphi))^\vee \in L_1(\mathbb G )$ we get
$0 = h x = \check\mathcal{T}(\psi)x - \check\mathcal{T}(\varphi)x$.
\end {proof}
The following proposition will play an important role in estimating the spectral decay.
\begin {prop}\label{inmeas}
Assume that $h\in\widehat{L_{\nu}}^{loc}(\mathbb G )$ is such that \[\hat h(\gamma) = \sum_{n\in{\mathbb Z}} c_n \gamma(g_n),\] $g_n\in\mathbb G $,
$n\in{\mathbb Z}$, and $\sum_{n\in{\mathbb Z}} |c_n|\nu(g_n) < \infty$. Then
\[\check\mathcal{T}(h)x = \sum_{n\in{\mathbb Z}} c_n \mathcal{T}(g_n)x = \mathcal{T}(\mu)x, \ x\in\mathcal{X}_c,\]
where the measure $\mu \in M_\nu(\mathbb{G})$ is given by $\mu = \sum_{n\in{\mathbb Z}} c_n \d_{-g_n}$, $\d_g$ -- Dirac measure concentrated at $g\in\mathbb{G}$.
\end {prop}
\begin {proof}
Assume $x\in \mathcal{X}_{Comp}$ and $f\in L_\nu(\mathbb G )$ is such that $fx = x$ and $\text{ supp\!} \hat f$ is compact. Then, using \eqref{redef},
\[\left(\check\mathcal{T}(h) - \sum_{n\in{\mathbb Z}} c_n \mathcal{T}(g_n)\right)x =\left(\check\mathcal{T}(h) - \sum_{n\in{\mathbb Z}} c_n \mathcal{T}(g_n)\right)\mathcal{T}(f)x=\]
\[\sum_{n\in{\mathbb Z}} c_n \check\mathcal{T}(\hat f(\gamma)\gamma(g_n))x - \sum_{n\in{\mathbb Z}} c_n \mathcal{T}(T(g_n)f)x = 0,\]
and the proposition follows from the definition of $\check\mathcal{T}(h)$ and Theorem \ref{CoHe}.
\end {proof}
The following two types of functions in $\widehat{L_1}^{loc}({\mathbb R}^d)$ are especially useful. For $\l=(\l_1,\ldots,\l_d)\in{\mathbb R}^d$ and $\a=(\a_1,\ldots,\a_d)\in {{\mathbb R}^d}$ we let
\begin {equation}\label{expfun}
{\textbf{e}}_\alpha (\l) = e^{\a\cdot\l} = \exp\left(\sum\limits_{k=1}^d{\a_k \l_k}\right)
\end{equation}
\begin {equation}\label{modexp}
{\textbf{h}}_\a (\l) =e^{\a\cdot|\l|} = \exp\left(\sum\limits_{k=1}^d{\a_k | \l_k|}\right),
\end{equation}
Observe that if $\a\in{\mathbb R}^d_+$, i.e. $\a_k>0$, $k=1,\ldots, d$, then ${\textbf{h}}_{-\a} = \hat f_\a$, where
$f_\a\in L_1({\mathbb R}^d)$ is the function introduced in \eqref{znet}, and for all $x\in D({\textbf{h}}_\a)$
\begin {equation}\label{fheq}
{\textbf{h}}_{-\a}\diamond({\textbf{h}}_\a\diamond x) =f_{\a}({\textbf{h}}_\a\diamond x) = {\textbf{h}}_\a\diamond(f_\a x) = x.
\end{equation}
Observe also that if
$\a=(0,\ldots,0,\a_k,0,\ldots,0)$, the following relation is valid:
\begin {equation}\label{releh}
{\textbf{e}}_\alpha +\textbf{e}_{-\a} = {\textbf{h}}_\a+{\textbf{h}}_{-\a},
\end{equation}
and, hence, using ${\textbf{h}}_{-\a} \in \widehat{L_1}({\mathbb R}^d)$,
\begin {equation}\label{dreleh1}
D({\textbf{e}}_\alpha )\cap D(\textbf{e}_{-\a})\subseteq D({\textbf{h}}_\a),\ \a\in{\mathbb R}^d_+.
\end{equation}
Another family of functions we shall use is given by $e_z(\l) = e^{iz\cdot \l}$, $z\in {\mathbb C}^d$. Observe that $e_{i\a} = \textbf{e}_{-\a}$, $\a\in{\mathbb R}^d$, and, hence, Propositions \ref{spro} and \ref{inmeas} imply
\begin {equation}\label{ee}
D(e_z) = D(\textbf{e}_{-\a}), \check\mathcal{T}(e_z) = \check\mathcal{T}(\textbf{e}_{-\a})\mathcal{T}(t),\ z = t+i\a,\ t,\a\in{\mathbb R}^d.
\end{equation}
\subsection{\texorpdfstring{Generators of $L_1({\mathbb R})$-modules}{Generators of L1({\mathbb R})-modules}}\label{gener}
Consider $L_1({\mathbb R})$-module $(\mathcal{X},\mathcal{T})$, satisfying $\mathcal{X} =\mathcal{X}_c$, that is the module structure is associated with a
strongly continuous isometric representation $\mathcal{T}: {\mathbb R} \to B(\mathcal{X})$. In this section we consider the properties of the (infinitesimal) generator $iA$ of the one-parameter group $\mathcal{T}$ \cite{EN06}.
The operator $A$ is called the generator of the module $\mathcal{X}$.
\begin {thm}\label{spmap
Assume $\L(\mathcal{X})$ is compact and nonempty. Then the generator $A\in B(\mathcal{X})$ and
we have the spectral mapping formula $\sigma(A) = \L(\mathcal{X})$. Moreover, the function $\mathcal{T}:{\mathbb R}\to B(\mathcal{X})$ admits a holomorphic extension to an entire function $\mathcal{T}(z) = e^{izA} = \mathcal{T}(t)e^{-\a A} = e^{-\a A}\mathcal{T}(t) = \check\mathcal{T}(e_z)$, where $e_z(\l) = e^{iz\l}$, $z = t+i\a\in{\mathbb C}$, and
\[\|\mathcal{T}(t+i\a)\| =\|\mathcal{T}(i\a)\| \le \max_{\l\in\L(\mathcal{X})}e^{-\a \l}, \ t,\a\in{\mathbb R}.\]
\end {thm}
\begin {proof}
Most of the assertions of this theorem are well known or nearly obvious (see, e.g., \cite{B79, B04}, \cite[Theorems 3.7, 3.8]{BK05} and references therein). In this paper, we shall prove only the norm estimate. We may assume that $\a \in {\mathbb R}_+$ and $a = |\inf \L(\mathcal{X})| \ge |\sup \L(\mathcal{X})|>0$ without loss of generality (otherwise we can consider the representation $\widetilde{\mathcal{T}}(t) = \mathcal{T}(-t)$). Let us estimate the norm
$\|\mathcal{T}(i\a)\|$.
Consider
the $4a$-periodic function $\eta$ defined on the interval $[-3a,a]$ via
\begin {equation}\label{curvehat}
\eta(\l) = \left\{
\begin{array}{cl}
e^{-\a\l}, & \l\in[-a,a],\\
e^{\a (2a+\l)}, & \l\in[-3a,-a).
\end{array}
\right.
\end{equation}
Computing the Fourier coefficients $c_n$, $n\in{\mathbb Z}$, of the periodic even function $\eta(\cdot - a)$ we get
\begin {equation}\label{fcoef}
c_n = 2\int_0^{2a} e^{\a(a-\l)}\cos \frac{\pi n \l}{2a}d\l=\frac{8\a a^2 e^{\a a}}{4\a^2a^2 +\pi^2n^2}(1 - e^{-2\a a}\cos\pi n)>0.
\end{equation}
Hence, in view of Proposition \ref{inmeas}, $\|\check\mathcal{T}(\eta)\| \le
e^{\a a}$.
Since $\eta \equiv e_{i\a}$
on $[-a,a]\supseteq\L(A)$, Lemma \ref{sssprop} ensures $\|\mathcal{T}(i\a)\| =\|\check\mathcal{T}(e_{i\a})\| = \|\check\mathcal{T}(\eta)\|$ and the theorem is proved.
\end {proof}
\begin {rem}
It is not hard to show that the inequality in the norm estimate is, in fact, an equality. We do not need this fact in this paper and, therefore, will refrain from proving it.
\end {rem}
The following corollary is immediate. There we do not assume that $\L(\mathcal{X})$ is compact.
\begin {cor}\label{Bineq}
Assume $x\in(\mathcal{X},\mathcal{T})$ and $\L(x)$ is compact and nonempty. Then the function $x_\mathcal{T}$: ${\mathbb R}\to \mathcal{X}$ given by $x_\mathcal{T}(t) = \mathcal{T}(t)x$ admits a holomorphic extension to an entire function and
\[\|x_\mathcal{T}(i\a)\| = \|e^{-\a A}x\| = \|\check\mathcal{T}(e_{i\a})x\| \le \|x\|\max_{\l\in\L(\mathcal{X})}e^{-\a \l},\ \a\in{\mathbb R}.\]
\end {cor}
\begin {proof}
Lemma \ref{sprop} implies that $\L([x]) = \L(x)$, where $[x]\subseteq\mathcal{X}$ is the submodule generated by $x$, i.e. the smallest
submodule of $\mathcal{X}$ containing $x$. It remains to apply Theorem \ref{spmap} to $[x]$.
\end {proof}
\begin {rem}
The above result leads to Bernstein-type inequalities that go back at least to \cite{B79} (see also \cite{BK05, GK10}).
Note that if $x\in\mathcal{X}_{Comp}$ then $x\in D(A)$ and $\|Ax\| \le \rho(x)\|x\|$, where $\rho(x) = \max\limits_{\l\in\L(x)} |\l|$. Hence, if $\L(\mathcal{X})$ is compact, then $A$ is a bounded operator and
$\|A\| = \rho(A)$, where $\rho(A)$ is the spectral radius of $A$, i.e. $\rho(A) = \max\{|\l|, \l\in\sigma(A)\}$.
\end {rem}
\begin {rem}\label{multig}
In case $\mathbb{G}={\mathbb R}^d$ we will consider \emph{multigenerators} of $L_1({\mathbb R}^d)$-modules (see Subsection \ref{sclass}). These are $d$-tuples comprised of the associated generators of $L_1({\mathbb R})$-modules from the Remark \ref{tenp}. We also note, in passing, that for the purposes of this paper (proving Wiener-type results) it is not necessary to consider representations $\mathcal{T}$ that are not strongly continuous (in which case the generator may not be densely defined). In view of Theorem \ref{CoHe}, all one would need to do if $\mathcal{T}$ is not strongly continuous to begin with, is to restrict attention to the submodule $\mathcal{X}_c$. As we shall se below, the results on inverse closedness will then be obtained without any loss of generality (compare with \cite{GK10}).
\end {rem}
\section{Different classes of spectral decay}\label{Cla}
In this section we introduce classes of elements with different spectral decay.
Unless stated otherwise, in this section we assume that $\mathbb G ={\mathbb R}^d$ and $\nu\equiv 1$.
\subsection{Exponential decay, one-sided exponential decay and causal classes}\label{exclass}
A major role in this paper is played by classes of elements with exponential spectral decay.
\begin {defn}\label{expdecay}
We say that $x\in(\mathcal{X},\mathcal{T})$ has \emph{exponential spectral decay of type} $\a\in{\mathbb R}^d_+$ if $x\in D({\textbf{h}}_\a)$. Similarly, $x$ has \emph{one-sided exponential spectral decay of type} $\a\in{\mathbb R}^d$ if $x\in D({\textbf{e}}_\alpha )$.
\end {defn}
In the following lemmas we provide useful characterizations of the newly defined classes. In the proof we use the classes of causal (and anticausal) elements which we define as in \cite{BK05}.
\begin {defn}
The submodules $\mathcal C_+(\mathcal{X})$ and $\mathcal C_-(\mathcal{X})$ of \emph{causal} and \emph{anticausal} elements in $\mathcal{X}$, respectively, are defined via
\[\mathcal C_\pm(\mathcal{X}) = \{x\in\mathcal{X}:\, \L(x)\subseteq \overline{{\mathbb R}^d_\pm}\}.\]
\end {defn}
We begin the study of the classes with a characterization of $D({\textbf{e}}_\alpha )$ in case $d=1$. In what follows we
consider a family of functions
$\phi_N \in L_1({\mathbb R}^d)$, $N\in{\mathbb N}$, such that $\norm{\phi_N}=1$,
$\hat\phi_N$ is continuous, $\text{ supp\!}\hat\phi_N\subset [-N,N]^d$, and $\phi_{N,n}(s) = \phi_N(s)e^{ i Nn\cdot s}$ so that $\hat\phi_{N,n}(\l) = \hat\phi_N(\l - Nn)$, $n\in{\mathbb Z}^d$.
In this paper we routinely use specific
functions $\phi_N$, $N\in{\mathbb N}$, defined via
\begin {equation}\label{triangle}
\hat\phi_N(\l) = \hat\phi(\frac\l N),\quad
\hat\phi(\l) = \hat\phi_1(\l)=(1-|\l|)\chi_{[-1,1]}(\l),
\end{equation}
where $\chi_S$ is, as usually, a characteristic function of the set $S$. It should be understood, however, that other functions with the specified properties may be used just as well.
\begin {rem}\label{Schwrep}
We can replace the function $\phi$ in \eqref{triangle} with a function $\varphi$ that is in the Schwartz class $\mathcal S$. As long as we have $\text{ supp\!} \hat\varphi\subseteq [-1,1]$ and
\[\sum_{n\in{\mathbb Z}} \hat \varphi (\l - n) \equiv 1,\]
all of the results in this paper remain valid.
\end {rem}
\begin {lem}\label{1side}
Let $x\in\mathcal{X}_c$, $N\in{\mathbb N}$, and $\phi_{N,n}$, $n\in{\mathbb Z}$, be defined by \eqref{triangle}.
Then the following are equivalent:
\begin{description}
\item[(i)] $x\in \bigcup_{\a > 0} D(\emph{\textbf{e}}_{\pm\a})$;
\item[(ii)] there are $M> 0$ and $\gamma\in [0,1)$ such that $\|\phi_{N, \pm n} x\|\le M\gamma^{Nn}$ for all $n \in {\mathbb N}$ (equivalently, for all $n\in{\mathbb Z}$ with $n\ge k$, $k\in{\mathbb Z}$);
\item[(iii)] the function $x_\mathcal{T}: {\mathbb R}\to\mathcal{X}$ given by
$x_\mathcal{T}(t) =\mathcal{T}(t)x$, $t\in{\mathbb R}$, admits a holomorphic extension to the strip
\[{\mathbb C}_{\mp\d} = \{z\in{\mathbb C}: 0 < \mp {\Im}m z < \d\}\]
for some $\d \in (0,\infty]$, which is continuous and uniformly bounded in $\overline{{\mathbb C}_{\mp\d}}$.
\end{description}
\end {lem}
\begin {proof}
(i)$\implies$(ii). Let us assume that $x\in D({\textbf{e}}_\alpha )$ for some $\a > 0$ and proceed to estimate $\|\phi_{1,n}x\|$ for $n\in{\mathbb N}$. Using Lemma \ref{ssprop} and Proposition \ref{spro}, we get
\[\|\phi_{N,n}x\| = \|f_\a({\textbf{e}}_\alpha \diamond(\phi_{N,n}x))\| = \|(f_{\a}*\phi_{N,n})({\textbf{e}}_\alpha \diamond x)\| \le \|f_{N,n}\|\|{\textbf{e}}_\alpha \diamond x\|,\]
where $f_{N,n}\in L_1({\mathbb R}^d)$ with $\hat f_{N,n} (\l)= \frac{e^{\alpha N(1-n)}}{{\textbf{h}}_{\alpha}(N(\l-n+1))}\in \widehat{L_1}({\mathbb R}^d)$, $n\in{\mathbb N}$. Since $f_{N,n}(s)e^{ i N(n-1)s}$ is positive for all $x\in{\mathbb R}$, $\|f_{N,n}\| = \hat f_n(n-1) = e^{\alpha N(1-n)}$ and we deduce that
$\|\phi_{N,n}x\| \le e^{\alpha N(1-n)}\|{\textbf{e}}_\alpha \diamond x\|$ for $n \in {\mathbb N}$. Similar argument works for
$x \in D(\textbf{e}_{-\a})$ and we obtain
\begin {equation}\label{eexpest}
\|\phi_{N,\pm n}x\| \le e^{\alpha N(1-n)}\|\textbf{e}_{\pm\a}\diamond x\|,\ n\in {\mathbb N}.
\end{equation}
(ii)$\implies$(iii). Assume now that $\|\phi_{N, -n} x\|\le M\gamma^{Nn}$, $n\ge -1$ (the other case is handled similarly). Then, we have $x = x_+ + x_-$, where
$x_- = \sum\limits_{n\ge -1} \phi_{N,-n}x$ with the series converging absolutely and $x_+= x - x_-$. Let us prove that $x_+ \in \mathcal C_+$. Indeed, let $\l\in [-Nk, -N(k-1))$, $k\in{\mathbb N}$. Consider a function $f\in L_1({\mathbb R})$ with $\hat f(\l)\neq 0$ and $\text{ supp\!}\hat f\subset (-N(k+1), -N(k-1))$. Then $$fx_+ = fx - f\left( \sum\limits_{n\ge -1} \phi_{N,-n}x\right) = fx - \left(f * \sum\limits_{n= k-2}^{k+2} \phi_{N,-n}\right)x = 0.$$ Therefore, $\l\notin\L(x_+)$. By \cite[Lemma 8.2]{BK05} the function $(x_+)_\mathcal{T}$: $t\mapsto \mathcal{T}(t)x_+$ admits the desired extension to the halfplane ${\mathbb C}_{\infty}$. Next, we observe that each vector $\phi_{N,n}x$, $n\in{\mathbb Z}$, has compact Beurling spectrum and, hence, Corollary \ref{Bineq} applies. In this way we obtain holomorphic extensions $(\phi_{N,-n}x)_\mathcal{T}$, $n\ge -1$, that satisfy
\[\|(\phi_{N,-n}x)_\mathcal{T}(t+i\a)\| \le e^{|\a| N(1-n)}\|\phi_{N,-n}x\| \le Me^{|\a| N} (\gamma e^{-|\a|})^{Nn},\ t,\a\in{\mathbb R}.\]
Hence, the series
\[(x_-)_\mathcal{T}(z) = \sum_{n\ge -1} (\phi_{N,-n}x)_\mathcal{T}(z)\]
converges absolutely and uniformly in the strip ${\mathbb C}_{\d}$, where $0 < \d < -\ln \gamma$. Therefore,
$x_\mathcal{T} (z) = (x_-)_\mathcal{T}(z) +(x_+)_\mathcal{T}(z)$ is a well-defined bounded function which is holomorphic in ${\mathbb C}_\d$ and
continuous in $\overline{{\mathbb C}_\d}$.
(iii)$\implies$(i). Assume that $x_\mathcal{T}$ is well defined, holomorphic in ${\mathbb C}_{-\d}$ for some $\d > 0$ as well as continuous and bounded in $\overline{{\mathbb C}_{-\d}}$. Let $f_k$ be a b.a.i. in $L_1({\mathbb R})$ with $\text{ supp\!} \hat f_k$ compact, $k\in{\mathbb N}$.
The functions $z\mapsto e_z\diamond (f_k x)$ and $z\mapsto f_k(x_\mathcal{T}(z))$ are both holomorphic in ${\mathbb C}_{-\d}$ and continuous and bounded in $\overline{{\mathbb C}_{-\d}}$. Since they coincide on ${\mathbb R}$, they coincide in ${\mathbb C}_{-\d}$ as well. Hence, the limit
\[\lim_{k\to\infty} e_z\diamond (f_k x) = \lim_{k\to\infty} f_k (x_\mathcal{T}(z)) = x_\mathcal{T}(z) \]
exists and, therefore, $x\in D(e_z) = D(\textbf{e}_{-\a})$ for all $z = t+i\a \in {\mathbb C}_{-\d}$ by \eqref{ee}.
\end {proof}
Observe that the estimates in the first part of the above proof remain valid when ${\textbf{e}}_\alpha $ is replaced
by ${\textbf{h}}_{\alpha}$, that is
\begin {equation}\label{expest}
\|\phi_{N,n}x\| \le e^{\alpha N(1-|n|)}\|{\textbf{h}}_{\alpha}\diamond x\|,\ |n|\ge1.
\end{equation}
In addition, since
$\phi_N\equiv \phi_{N,0}\in L_1({\mathbb R})$ and $\|\phi_N\| = 1$, we have that
\[\|\phi_Nx\|=\|\phi_{N,0}x\|\le\|x\|\le\|{\textbf{h}}_\a\diamond x\|, N\in{\mathbb N}, \a>0,\]
due to \eqref{fheq}.
The above estimates indicate that the terminology in Definition \ref{expdecay} is appropriate. Moreover, together with \eqref{dreleh1} and Remark \ref{tenp} they imply
\begin {equation}\label{dreleh}
D({\textbf{e}}_\alpha )\cap D(\textbf{e}_{-\a})= D({\textbf{h}}_\a),\ \a\in{\mathbb R}^d_+.
\end{equation}
The following lemma is almost immediate now.
\begin {lem}\label{2side}
Let $x\in\mathcal{X}_c$, $N\in{\mathbb N}$, and $\phi_{N,n}$, $n\in{\mathbb Z}$, be defined by \eqref{triangle}.
Then the following are equivalent:
\begin{description}
\item[(i)] $x\in \bigcup_{\a > 0} D(\emph{\textbf{h}}_{\a})$;
\item[(ii)] there are $M\ge 0$ and $\gamma\in [0,1)$ such that $\|\phi_{N, n} x\|\le M\gamma^{N|n|}$ for all $n \in {\mathbb Z}$;
\item[(iii)] the function $x_\mathcal{T}: {\mathbb R}\to\mathcal{X}$ given by
$x_\mathcal{T}(t) =\mathcal{T}(t)x$, $t\in{\mathbb R}$, admits a holomorphic extension to the strip
\[{\mathbb C}_{-\d,\d} = \{z\in{\mathbb C}: -\d < {\Im}m z < \d\}\]
for some $\d \in (0,\infty]$, which is continuous and uniformly bounded in $\overline{{\mathbb C}_{-\d,\d}}$.
\end{description}
\end {lem}
\begin {proof}
Assume $x\in D({\textbf{h}}_{\alpha})$. In view of \eqref{dreleh}, Lemma \ref{1side} provides us with a holomorphic extension of $x_\mathcal{T}$ to ${\mathbb C}_{\pm\d}$ for some $\d >0$. The only thing that remains to prove is that this extension is holomorphic in a neighborhood of ${\mathbb R}\subset {\mathbb C}$. Observe that the family of functions $x_\a \equiv x_\mathcal{T}(\cdot + i\a): {\mathbb R}\to\mathcal{X}$ is uniformly bounded and equicontinuous for $\a\in[-\d,\d]$. Hence, the extension $x_\mathcal{T}$ is holomorphic in ${\mathbb C}_{-\d,\d}$ as a uniform limit of a sequence of entire functions $f_n x_\mathcal{T}$, where $(f_n)$ is a b.a.i.~ in $L_1({\mathbb R})$ with $\text{ supp\!} \hat f_n$ compact.
\end {proof}
Remark \ref{tenp} allows us to extend Lemmas \ref{1side} and \ref{2side} to the multidimensional case.
Using the functions from \eqref{triangle}, we define $\phi_{N,a}^d\in L_1({\mathbb R}^d)$, $a\in{\mathbb R}^d$, $N\in{\mathbb N}$, via
\begin {equation}\label{triangled}
\hat\phi_{N,a}^d (\l) = \hat\phi_{N}^d(\l-Na), \ \mbox{ where }
\hat \phi_{N}^d (\l_1,\ldots,\l_d)= \prod_{k=1}^d \hat\phi_{N}(\l_k).
\end{equation}
We obtain the following characterizations of the classes with exponential spectral decay.
\begin {prop}\label{1expestprop}
Let $x\in\mathcal{X}_c$,
$N\in{\mathbb N}$, and $\phi_{N,n}^d$, $n\in{\mathbb Z}^d$, be defined by \eqref{triangle} and \eqref{triangled}. Then the following are equivalent:
\begin{description}
\item[(i)] $x\in \bigcup_{\a \in{\mathbb R}^d_\pm} D(\emph{\textbf{e}}_{\a})$;
\item[(ii)] there are $M\ge 0$ and $\gamma\in [0,1)$ such that $\|\phi_{N,\pm n}^d x\|\le M\gamma^{N|n|}$, $n\in{\mathbb Z}^d_+$;
\item[(iii)] the function $x_\mathcal{T}: {\mathbb R}\to\mathcal{X}$ given by
$x_\mathcal{T}(t) =\mathcal{T}(t)x$, $t\in{\mathbb R}$, admits a holomorphic extension to the strip
\[{\mathbb C}^d_{\mp\d} = \{z = (z_1, z_2, \ldots, z_d)\in{\mathbb C}^d: 0 < \mp {\Im}m z_k < \d_k, k=1,2,\ldots, d\}\]
for some $\d = (\d_1, \d_2,\ldots, \d_d) \in {\mathbb C}_+^d$, which is continuous and uniformly bounded in $\overline{{\mathbb C}^d_{\mp\d}}$.
\end{description}
\end {prop}
\begin {prop}\label{expestprop}
Let $x\in\mathcal{X}_c$,
$N\in{\mathbb N}$, and $\phi_{N,n}^d$, $n\in{\mathbb Z}^d$, be defined by \eqref{triangle} and \eqref{triangled}.
Then the following are equivalent:
\begin{description}
\item[(i)] $x\in \bigcup_{\a \in{\mathbb R}^d_+} D(\emph{\textbf{h}}_{\a})$;
\item[(ii)] there are $M\ge 0$ and $\gamma\in [0,1)$ such that $\|\phi_{N,\pm n}^d x\|\le M\gamma^{N|n|}$;
\item[(iii)] the function $x_\mathcal{T}: {\mathbb R}\to\mathcal{X}$ given by
$x_\mathcal{T}(t) =\mathcal{T}(t)x$, $t\in{\mathbb R}$, admits a holomorphic extension to the strip
\[{\mathbb C}^d_{-\d,\d} = \{z = (z_1, z_2, \ldots, z_d)\in{\mathbb C}^d: -\d_k < {\Im}m z_k < \d_k, k=1,2,\ldots, d\}\]
for some $\d = (\d_1, \d_2,\ldots, \d_d) \in {\mathbb C}_+^d$, which is continuous and uniformly bounded in $\overline{{\mathbb C}_{\mp\d}}$.
\end{description}
\end {prop}
We will often use the following more precise estimates that were obtained in the course of the proof:
\begin {equation}\label{expestd}
\|\phi_{N,n}^dx\|\le e^{NM_n(\a)}\|{\textbf{h}}_{\alpha}\diamond x\|, \ x\in D({\textbf{h}}_{\alpha}), \ n\in{\mathbb Z}^d,
\end{equation}
\begin {equation}\label{eexpestd}
\|\phi_{N,\pm n}^dx\|\le e^{NM_n(\a)}\|\textbf{e}_{\pm\a}\diamond x\|, \ x\in D(\textbf{e}_{\pm\a}),\
n\in{\mathbb Z}^d_+,
\end{equation}
where
\begin {equation}\label{mn}
M_n (\a)= \sum_{k:\ n_k\ne 0} \a_k(1-|n_k|), \ n=(n_1,\ldots,n_d)\in{\mathbb Z}^d.
\end{equation}
Observe also that for any
$\a\in{\mathbb R}^d_+$ we have $\mathcal{X}_{Comp} \subseteq D({\textbf{e}}_\alpha )$ and
$\mathcal{X}_c\cap D({\textbf{e}}_{\beta}) \subseteq X_c\cap D({\textbf{e}}_{\a})$ whenever $\beta\in{\mathbb R}^d_\pm$ and $\a-\beta \in{\mathbb R}^d_\pm$. We also mention explicitly that extending \cite[Lemma 8.2]{BK05} via Remark \ref{tenp}, we see that for every
$\mathcal{T}$-continuous vector $x\in\mathcal C_+(\mathcal{X})$ the function $x_\mathcal{T}$ has a bounded holomorphic extension to the halfspace ${\mathbb C}_{\infty}^d$ given by
\begin {equation}\label{holex}
x_\mathcal{T}(z) = {\textbf{h}}_{\a}\diamond (\mathcal{T}(t)x)= \textbf{e}_{-\a}\diamond (\mathcal{T}(t)x) = e^{iz\cdot\l}\diamond x,\ z = t+i\a\in\overline{{\mathbb C}^d_{\infty}},
\end{equation}
and a similar formula holds for $x\in \mathcal C_-(\mathcal{X})\cap\mathcal{X}_c$.
\subsection{Wiener Class}\label{wiencl}
The analog of the classical Wiener class $\mathcal{W}\subseteq \mathcal{X}$ in our setting is defined as follows:
\begin {equation}
\mathcal{W} =\mathcal{W}(\mathcal{X}) = \mathcal{W}(\mathcal{X},\mathcal{T}):= \left\{x\in\mathcal{X}:\ \|x\|_\mathcal{W} =\int_{{\mathbb R}^d}\|\phi^d_{1,a} x\|da< \infty \right\},
\end{equation}
where the functions $\phi^d_{1,a}$ are given by \eqref{triangled}. Since the representation $M$ defined by \eqref{mod1} on $L_1({\mathbb R}^d)$ is strongly continuous, the function $a\mapsto \phi^d_{1,a} x: {\mathbb R}^d\to\mathcal{X}$ is continuous and, therefore, the finiteness of the integral is a reasonable way to define the corresponding type of spectral decay. Presently, we shall see that
$\mathcal{W}$ is a Banach space; for our purposes, however, it is more convenient to use equivalent
series norms rather than the integral norm in the definition. We define these norms via
\begin {equation}\label{wiensumel}
\|x\|_{1,N} = 5^d \sum_{n\in{\mathbb Z}^d} \|\phi^d_{N,n} x\|,
\ N\in{\mathbb N}.
\end{equation}
The seemingly strange choice of the constant $5^d$ will become natural when we study the Wiener Class of operators.
\begin {prop}\label{preq}
The set $\mathcal{W}$ is a Banach space with respect to any of the equivalent norms in \eqref{wiensumel}.
Moreover,
\begin {equation}\label{normeq}
\frac1{(2N+1)^d} \|x\|_{1,1} \le \|x\|_{1,N} \le 3^d\|x\|_{1,1},\ \mbox{and}
\end{equation}
\begin {equation}\label{normeq1}
\|x\|_{\mathcal{W}} \le \|x\|_{1,1} \le 20^d\|x\|_{\mathcal{W}},\ x\in\mathcal{W}.
\end{equation}
\end {prop}
\begin {proof}
We begin by showing that $x\in\mathcal{W}$ if and only if $\|x\|_{1,1}<\infty$, obtaining \eqref{normeq1} in the process. The crucial observation here is (for $d=1$)
that \[\phi_{1,a}x =\phi_{1,a}((\phi_{1,n-1}+\phi_{1,n}+\phi_{1,n+1}+\phi_{1,n+2})x),\] where
$n = \lfloor a \rfloor$ is the largest integer less than or equal to $a\in{\mathbb R}$. A similar equality for $a\in{\mathbb R}^d$ immediately implies
\[ \int_{{\mathbb R}^d}\|\phi_{1,a}^d x\|da \le 4^d \sum_{n\in {\mathbb Z}^d} \|\phi_{1,n}^dx\| \le \|x\|_{1,1}.\]
Similarly, since
\[\|\phi_{1,n}^dx\| = \int_{n+[0,1]^d} \left\|\phi_{1,n}^d\left(\left(\sum_{k\in S_n}\phi_{1,a-k}^d\right)x\right)\right\|da,\]
where each $S_n$, $n\in{\mathbb Z}^d$, is the set of cardinality at most $4^d$ such that the Fourier transform of the sum is equal to 1 on $\text{ supp\!} \hat\phi_{1,n}^d$, we have
\[\|x\|_{1,1} \le 20^d \int_{{\mathbb R}^d}\|\phi^d_{1,a} x\|da.\]
Next, we prove the inequalities \eqref{normeq}. We use Lemma \ref{ssprop} and Remark \ref{tenp} in the usual way to obtain them.
The second of the estimates follows from
\begin{equation*
\begin {split}
\sum_{n\in{\mathbb Z}^d} \|\phi^d_{N,n}x\| &= \sum_{n\in{\mathbb Z}^d} \left\|\phi^d_{N,n}\left(\sum_{k\in{\mathbb Z}^d} \phi^d_{1,k}x\right)\right\| \\
&\le\sum_{k\in{\mathbb Z}^d} \sum_{n\in S_k}\|\phi^d_{N,n}\| \left\|\phi^d_{1,k}x\right\| \le 3^d \sum_{k\in{\mathbb Z}^d} \left\|\phi^d_{1,k}x\right\|,
\end{split}
\end{equation*}
where $S_k\subset {\mathbb Z}^d$ is a set of $d$-tuples $n\in{\mathbb Z}^d$ such that the intersection of the supports
of $\hat\phi^d_{N,n}$ and $\hat\phi^d_{1,k}$ has non-empty interior (all these sets have cardinality at most $3^d$).
Similarly, the first inequality in \eqref{normeq} follows from
\begin{equation*
\begin {split}
\sum_{n\in{\mathbb Z}^d} \|\phi^d_{1,n}x\| &= \sum_{n\in{\mathbb Z}^d} \left\|\phi^d_{1,n}\left(\sum_{k\in{\mathbb Z}^d} \phi^d_{N,k}x\right)\right\| \\
&\le \sum_{k\in{\mathbb Z}^d} \sum_{n\in S^k}\|\phi^d_{1,n}\| \left\|\phi^d_{N,k}A\right\| = (2N + 1)^d\sum_{k\in{\mathbb Z}^d} \left\|\phi^d_{N,k}x\right\|,
\end{split}
\end{equation*}
where $S^k\subset {\mathbb Z}^d$ is a set of numbers $n\in{\mathbb Z}^d$ such that the intersection of the supports
of $\hat\phi^d_{1,n}$ and $\hat\phi^d_{N,k}$ has non-empty interior (all these sets have cardinality
$(2N+1)^d$).
The rest of the assertions in the proposition are now trivial.\end {proof}
\begin {rem}\label{wc}
Observe that in view of Theorem \ref{CoHe} the definition of Wiener class implies $\mathcal{W}(\mathcal{X})\subseteq \mathcal{X}_c$. This is why we decided to limit many of the auxiliary statements appearing in this
paper to $\mathcal{X}_c$ even though in most cases they can be stated more generally.
\end {rem}
Let us present a few simple examples of the class $\mathcal{W}$.
\begin {exmp}\label{ex1}
Let $\mathcal{X} = L^p({\mathbb R}^d)$, $1\le p \le\infty$, be the space of equivalence classes of $p$-integrable complex-valued functions, and $T$ and $M$ be the translation and modulation representations defined as in \eqref{trans1} and \eqref{mod1}. Let also $U$ be a representation based on a resolution of the identity $\mathcal{P} = \{P_k\}_{k\in{\mathbb Z}^d}$, as in \eqref{exres}. Then
\[
\begin {split}
\mathcal{W}(\mathcal{X}, M) & = \left\{f\in\mathcal{X}:\ \sum_{n\in{\mathbb Z}^d} \left(\int_{{\mathbb R}^d}\left|\hat\phi^d_{1,n}(t)x(t)\right|^pdt\right)^{\frac1p}<\infty\right\} \\
& = \left\{f\in\mathcal{X}:\ \sum_{n\in{\mathbb Z}^d} \left(\int_{[-1,1]^d}\left|x(t-n)\right|^pdt\right)^{\frac1p}<\infty\right\},
\end{split}\]
$1\le p< \infty$, with the obvious change for
$p=\infty$,
is the Wiener amalgam space $L^{p,1}({\mathbb R}^d)$ \cite{FS85, H75, W32}. For $p=2$ we also have
$\mathcal{W}(L^2({\mathbb R}^d), T) = \widehat L^{2,1}({\mathbb R}^d) = \{x\in L^2({\mathbb R}^d):\ \hat x\in L^{2,1}({\mathbb R}^d)\}$.
In case of the representation $U$, we get $\mathcal{W}(\mathcal{X}, U) = \{x\in\mathcal{X}: \sum_{k\in{\mathbb Z}^d} \|P_kx\|<\infty\}$. In particular, if $(P_k x)(t) = \chi_{k+[0,1]^d}(t)x(t)$, we get $\mathcal{W}(\mathcal{X}, U) = \mathcal{W}(\mathcal{X}, M) = L^{p,1}({\mathbb R}^d)$.
Alternatively, if $(P_k x)(t) = (\chi_{k+[0,1]^d}\hat x)^\vee(t)$, $x\in L^2({\mathbb R}^d)$, we get
$\mathcal{W}(L^2({\mathbb R}^d), U) = \mathcal{W}(L^2({\mathbb R}^d), T)$. This shows that it may not be advantageous to consider
representations other than $U$ when $\mathcal{X} = L^p$. As a consequence, when we get to presenting
examples in Section \ref{examp} we concentrate primarily on Banach spaces other than $L^p$. Indeed,
in case of more general Banach spaces, desired resolutions of the identity may not exist and the use of different representations becomes extremely beneficial.
\end {exmp}
\subsection{Beurling Class} This class of elements is designed to encompass the classical
Beurling algebra of functions
\[A^*({\mathbb T}) = \left\{\sum_{n=-\infty}^\infty a(n)e^{in\xi}:\ \sum_{k=0}^\infty \sup_{|n|\ge k}|a(n)|<\infty\right\}\]
and was inspired by the work of Q.~Sun in \cite{S10ca}, see also \cite{BLT97}.
Again, we use the functions $\phi^d_{1,n}$, $n\in{\mathbb Z}^d$, defined by \eqref{triangled} and let
${\mathcal B} = \{x\in \mathcal{X}:\ \|x\|_{{\mathcal B}}<\infty\}$, where
\begin {equation}\label{bnorm}
\|x\|_{{\mathcal B}}=
\sum_{k\in{\mathbb Z}^d}^\infty \max_{|n|_\infty\ge |k|_\infty} \|\phi^d_{1,n}x\|,
\end{equation}
where $|k|_\infty = \max\{k_1,k_2,\ldots, k_d\}$, $k = (k_1,k_2,\ldots, k_d)$.
It is easily seen that \eqref{bnorm} defines a Banach space norm on ${\mathcal B}$ and for any $\a > 0$
we have
\begin {equation}
D({\textbf{h}}_{\alpha}) \subseteq {\mathcal B}\subseteq \mathcal{W}.
\end{equation}
\subsection{Sobolev-type classes}\label{sclass}
It is natural to consider weighted extensions of the classes introduced in this section. In general, we shall pursue such extensions in a sequel to this paper. Here we limit ourselves to Sobolev-type classes defined as follows (see Remark \ref{winsob}).
As usually, we consider a Banach space $\mathcal{X}$ equipped with a non-degenerate $L_1({\mathbb R}^d)$-module structure associated with a strongly continuous isometric representation $\mathcal{T}:{\mathbb R}^d\to B(\mathcal{X})$. We let $A = (A_1,\ldots, A_d)$ be the \emph{multigenerator} of the module $(\mathcal{X},\mathcal{T})$, i.e. each operator $iA_k$ is the infinitesimal generator of the one-parameter group $\mathcal{T}^{(k)}:{\mathbb R}\to B(\mathcal{X})$ from Remark \ref{tenp}, $k = 1,\ldots, d$. In the following definitions we use the standard multi-index notation, i.e. $\a = (\a_1,\ldots \a_d)\in {\mathbb Z}^d_+$,
$|\a| =|\a|_1 = \a_1+\ldots +\a_d$, and $A^\a = A_1^{\a_1}\ldots A_d^{\a_d}$.
\begin {defn}
For each $m\in{\mathbb N}$ the \emph{Sobolev-Wiener class} $\mathcal{W}^{(m)}$ is given by
\[\mathcal{W}^{(m)} = \{x\in\mathcal{X}:\, x\in D(A^\a) \mbox{ and } A^\a x\in\mathcal{W} \mbox{ for all } |\a|\le m, \a\in{\mathbb Z}^d_+\}.\]
\end {defn}
\begin {defn}
For each $m\in{\mathbb N}$ the \emph{Sobolev-Beurling class} ${\mathcal B}^{(m)}$ is given by
\[{\mathcal B}^{(m)} = \{x\in\mathcal{X}:\, x\in D(A^\a) \mbox{ and } A^\a x\in{\mathcal B} \mbox{ for all } |\a|\le m, \a\in{\mathbb Z}^d_+\}.\]
\end {defn}
\begin {rem}\label{winsob}
It can be easily shown that for $\a = (m,\ldots, m)\in {\mathbb N}^d$ we have $D(A^{\a}) = D_\mathcal{T}(\varphi_{m})$, where
$\varphi_{m}(\l) = (\l_1\ldots \l_d)^{m}$, $\l = (\l_1,\ldots, \l_d)\in {\mathbb R}^d$. Hence, $A\in \mathcal{W}^{(m)}$ if and only if
\[\int_{{\mathbb R}^d}\|\phi^d_{1,a}A\|(1+|a|)^{m}da < \infty.\]
We refer to \cite{GK10, Kl11} for similar and other classes utilizing $D(A^\a)$.
\end {rem}
\begin {rem}
It is not hard to see that in the case of Hilbert spaces or when $\mathcal{X} = C_{ub}({\mathbb R}^d,{\mathcal H})$ is a space of bounded uniformly continuous functions with values in a Hilbert space ${\mathcal H}$, smoothness of the function $x_\mathcal{T}$ (the property $x\in D(A^\a)$) is closely related to the spectral decay of $x$ (see, e.g., \cite{GK10}). In particular, for $x\in C_{ub}({\mathbb R},{\mathcal H})=\mathcal{X}$ we have that $x\in D(A)$ implies $x\in \mathcal{W}(\mathcal{X}, T)$, where $T$ is as in \eqref{trans1}. In general Banach spaces, however, this is not the case. Consider the following example. Let $\mathcal{X} = C_{ub}({\mathbb R}, C_{ub}({\mathbb R}))$ and $\{\a_n\}_{n\in{\mathbb Z}}$ be a null sequence ($\lim a_n = 0$) such that the series
$\sum\limits_{n\in{\mathbb Z}} \frac{\a_n}n$ diverges. Define $\varphi_n = \a_n\hat\phi_{\frac12, n}$ where $\phi_{\frac12,n}$ is as in \eqref{triangled} and let $x\in\mathcal{X}$ be the periodic function given by
\[x(t) = \sum_{n\in{\mathbb Z}} \frac1n\varphi_ne^{int},\ \varphi_n\in C_{ub}({\mathbb R}),\]
The function is well defined because the series converges unconditionally in $\mathcal{X}$. Moreover, the same is true for
\[x^\prime(t) = \sum_{n\in{\mathbb Z}} i\varphi_ne^{int},\]
so that $x\in D(A)$, $A =-i\frac{d}{dt}$. On the other hand, it is clear that $x\notin \mathcal{W}(\mathcal{X},T)$. Hence, in general, the relation between smoothness of $x_\mathcal{T}$ and the spectral decay of $x$ is subtler.
Let us prove that $x\in D(A^2)$ implies $x\in \mathcal{W}(\mathcal{X},\mathcal{T})$ for general $\mathcal{X}$ and $\mathcal{T}: {\mathbb R}\to B(\mathcal{X})$. Let
$y\in D(A^2)$. Then there exists $x\in\mathcal{X}$ such that $y = (A^2+I)^{-1}x = fx$, where $\hat f(\l) = (\l^2+1)^{-1}$. Hence, $\phi_{1,n}y = (f*\phi_{1,n})x$ satisfies $\|\phi_{1,n}y\| \le \|f*\phi_{1,n}\|\|x\| \le \frac{c}{n^2}$ for some $c\in{\mathbb R}$ and, therefore, $y\in\mathcal{W}(\mathcal{X},\mathcal{T})$. Clearly, a small modification of the above argument shows
$D(A^{1+\varepsilon})\subseteq \mathcal{W}(\mathcal{X},\mathcal{T})$ for any $\varepsilon > 0$.
\end {rem}
\begin {rem}\label{periodize}
We defined various classes of vectors with spectral decay assuming $\mathbb G = {\mathbb R}^d$. Such a restriction, however, is less stringent than one may think. Indeed, if $\mathbb G = {\mathbb T}^{d_1}\times{\mathbb R}^{d_2}$ one can pass to a representation
on ${\mathbb R}^{d_1+d_2}$ which would be periodic in the first $d_1$ variables and obtain the same kind of results just as well.
\end {rem}
\section{Module structures and memory of linear operators}\label{Clop}
In view of our interest in Wiener-type results, we consider the classes defined in the previous section for Banach modules $L(\mathcal{X}, \mathcal{Y})$ and $L(\mathcal{Y}, \mathcal{X})$ of bounded linear operators between
complex Banach spaces $\mathcal{X}$ and $\mathcal{Y}$. To avoid confusion we shall use
$\mathcal{T}_{\mathcal{X}\mathcal{Y}}: \mathbb{G}\to B(L(\mathcal{X},\mathcal{Y}))$ to denote the representation defining the module structure on
$L(\mathcal{X},\mathcal{Y})$ and $\mathcal{T}_{\mathcal{Y}\mathcal{X}}: \mathbb{G}\to B(L(\mathcal{Y},\mathcal{X}))$ -- for the representation on $L(\mathcal{Y},\mathcal{X})$. Since we are interested only in operators with some kind of memory decay we will only deal with operators in
$(L(\mathcal{X},\mathcal{Y}))_c$ and $(L(\mathcal{Y},\mathcal{X}))_c$. This allows us to use formula \eqref{scdef} and avoid tedious complications that we had to deal with in \cite{BK05}.
We shall also always assume that the
representations $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$ and $\mathcal{T}_{\mathcal{Y}\mathcal{X}}$ are \emph{coupled}, i.e.
the operators $A\in L(\mathcal{X},\mathcal{Y})$, $B\in L(\mathcal{Y},\mathcal{X})$ are inverses of each other if and only if
$\mathcal{T}_{\mathcal{X}\mathcal{Y}}(g)A$ and $\mathcal{T}_{\mathcal{Y}\mathcal{X}}(g)B$ have the same property for every $g\in\mathbb{G}$.
Moreover, given three (not necessarily distinct) complex Banach spaces $\mathcal{X}$, $\mathcal{Y}$, $\mathcal{Z}$ and
Banach modules $L(\mathcal{X},\mathcal{Y})$, $L(\mathcal{Y},\mathcal{Z})$, and $L(\mathcal{X},\mathcal{Z})$ with structures associated with $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$,
$\mathcal{T}_{\mathcal{Y}\mathcal{Z}}$ and $\mathcal{T}_{\mathcal{X}\mathcal{Z}}$, respectively, we assume that
\begin {equation}\label{couple}
\mathcal{T}_{\mathcal{X}\mathcal{Z}}(g)(BA) = (\mathcal{T}_{\mathcal{Y}\mathcal{Z}}(g)B)(\mathcal{T}_{\mathcal{X}\mathcal{Y}}(g)A),\ g\in\mathbb{G},
\end{equation}
for all $A\in L(\mathcal{X},\mathcal{Y})$, $B\in L(\mathcal{Y},\mathcal{Z})$. In particular, if $\mathcal{X} = \mathcal{Y}$, we assume that $\mathcal{T}_{\mathcal{X}\X}(g)$, $g\in\mathbb G $, are Banach algebra automorphisms of $B(\mathcal{X})$.
If the spaces $\mathcal{X}$, $\mathcal{Y}$, and $\mathcal{Z}$ are themselves Banach modules, we shall use $\mathcal{T}_\mathcal{X}$, $\mathcal{T}_\mathcal{Y}$, and $\mathcal{T}_\mathcal{Z}$ to denote the corresponding representations.
However, in what follows, we omit the indices of the representations if the choice is unambiguous.
\begin {exmp}\label{typop}
The most typical example of module structures on $L(\mathcal{X}, \mathcal{Y})$ and $L(\mathcal{Y},\mathcal{X})$ satisfying the above assumptions arises when $\mathcal{X}$ and $\mathcal{Y}$ are equipped with module structures associated with $\mathcal{T}_\mathcal{X}$ and $\mathcal{T}_\mathcal{Y}$, respectively. In this case, we let $\mathcal{T}_{\mathcal{X}\mathcal{Y}}(g)A = \mathcal{T}_\mathcal{Y}(g)A\mathcal{T}_\mathcal{X}(-g)$ and
$\mathcal{T}_{\mathcal{Y}\mathcal{X}}(g)B = \mathcal{T}_\mathcal{X}(g)B\mathcal{T}_\mathcal{Y}(-g)$, $g\in\mathbb{G}$, $A\in L(\mathcal{X},\mathcal{Y})$, $B\in L(\mathcal{Y},\mathcal{X})$. Clearly, these are coupled representations and \eqref{couple} is satisfied in case we have three spaces and define the representations in this way. We also observe that when the representations $\mathcal{T}_\mathcal{X}$ and $\mathcal{T}_\mathcal{Y}$ are defined via resolutions of the identity as in \eqref{exres} and $A_{mn}$ are operator blocks
of an operator $A\in L(\mathcal{X},\mathcal{Y})$ as described in the introduction, then $k\notin \L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})$ if and only if $A_{mn} = 0$ for all $m,n \in {\mathbb Z}^d$ such that $m-n=k$. In other words, the Beurling spectrum of an operator $A$ consists of the numbers of the non-zero diagonals of its matrix.
\end {exmp}
The module structures not conforming to the previous example are rarely considered in the literature. It is, however, an unnecessary and, at times, harmful restriction to consider only such kind of representations. In particular, within the framework of the above example one cannot consider more general Banach algebras than $B(\mathcal{X})$ (see Remark \ref{ba}). The following structure, although very similar, also lies beyond the framework.
\begin {exmp}
For simplicity, we present this example in case $\mathbb{G} = {\mathbb R}^{2d}$ and leave an obvious extension
to more general LCA-groups to the reader. Consider the translation and modulation representations $T$ and $M$ defined by \eqref{trans1} and \eqref{mod1} on two
appropriate Banach spaces $\mathcal{X}$ and $\mathcal{Y}$ of functions on ${\mathbb R}^d$. In this case, the representations
satisfying our assumptions can be defined as
\begin {equation}\label{Weyl}
\mathcal{T}_{\mathcal{X}\mathcal{Y}}(t,s)A = M(t)T(s)AT(-s)M(-t), \ s,t\in{\mathbb R}^d, \ A\in L(\mathcal{X},\mathcal{Y}),\end{equation}
and similarly for $\mathcal{T}_{\mathcal{Y}\mathcal{X}}$.
\end {exmp}
The following lemma is crucial for understanding the properties of the memory, i.e. Beurling spectra, of linear operators. Although a close analog of this result appears, for example, in \cite{BK05, BR75} we feel compelled to present a proof here.
\begin {lem}\label{spprodf}
Let $A\in L(\mathcal{X},\mathcal{Y})$, $B\in L(\mathcal{Y},\mathcal{Z})$ and $(L(\mathcal{X},\mathcal{Y}), \mathcal{T}_{\mathcal{X}\mathcal{Y}})$, $(L(\mathcal{Y},\mathcal{Z}), \mathcal{T}_{\mathcal{Y}\mathcal{Z}})$, and
$(L(\mathcal{X},\mathcal{Z}), \mathcal{T}_{\mathcal{X}\mathcal{Z}})$ be non-degenerate Banach modules over the algebras $L_{\nu_1}(\mathbb{G})$, $L_{\nu_2}(\mathbb{G})$, and $L_{\nu_3}(\mathbb{G})$, respectively, where $\nu_i$, $i=1,2,3$, are non-quazi-analytic weights. Assume also that \eqref{couple} is satisfied and that $A\in \overline{\Omega(A)}$ and
$B\in \overline{\Omega(B)}$ (see Definition \ref{modorbit}).
Then
\[\L(BA,\mathcal{T}_{\mathcal{X}\mathcal{Z}})\subseteq\overline{\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})+\L(B,\mathcal{T}_{\mathcal{Y}\mathcal{Z}})}.\]
\end {lem}
\begin {proof}
Assume that $\gamma\notin \overline{\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})+\L(B,\mathcal{T}_{\mathcal{Y}\mathcal{Z}})}=:\Delta$, and let
$\mathcal{U}$ and $\mathcal{V}$ be two neighborhoods of $\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})$
and $\L(B,\mathcal{T}_{\mathcal{Y}\mathcal{Z}})$, respectively, such that $\gamma\notin\mathcal U+\mathcal V$.
Let also
$\phi\in L_{\nu_1}(\mathbb{G})$, $\psi\in L_{\nu_2}(\mathbb{G})$, and $f\in L_{\nu_3}(\mathbb{G})$ be such that $\text{ supp\!} \hat\phi\in \mathcal U$, $\text{ supp\!}\hat\psi\in\mathcal V$, $\hat f(\gamma) \ne 0$, and $\text{ supp\!} \hat f \cap \overline{\text{ supp\!} \hat\phi +\text{ supp\!}\hat\psi}=\emptyset$.
Such a function
$f$ exists because the weight $\nu_3$ is non-quazianalytic \cite{D56, LMF73}.
In view of Lemma \ref{sprop}(vi) and \eqref{goodf} it is enough to prove that
$\L(B_1A_1,\mathcal{T}_{\mathcal{X}\mathcal{Z}})\subseteq\Delta$,
where $A_1 = \phi A$ and $B_1 = \psi B$.
Then, omitting the indices of the representations, we get
\[
\begin {split}
& \mathcal{T}(f)(B_1A_1) = \int_\mathbb{G} f(g)\mathcal{T}(-g)(B_1A_1)dg \\ &= \int_\mathbb{G} f(g)(\mathcal{T}(-g)(\psi B))(\mathcal{T}(-g)(\phi A))dg \\
& = \int_\mathbb{G}\int_\mathbb{G}\int_\mathbb{G} f(g)\phi(g_1)\psi(g_2)(\mathcal{T}(-g_2-g)B)(\mathcal{T}(-g_1-g)A)dg_1dg_2dg \\
& = \int_\mathbb{G}\int_\mathbb{G}\int_\mathbb{G} f(g)\phi(g_1-g)\psi(g_2-g)(\mathcal{T}(-g_2)B)(\mathcal{T}(-g_1)A)dgdg_1dg_2 \\
& = \int_\mathbb{G}\int_\mathbb{G} F(g_1,g_2)(\mathcal{T}(-g_2)B)(\mathcal{T}(-g_1)A)dg_1dg_2.
\end{split}
\]
Observe that by our assumptions $\hat F(\gamma_1,\gamma_2) = \hat f(\gamma_1+\gamma_2)\hat\phi(\gamma_1)\hat\psi(\gamma_2) = 0$, and, hence, $\mathcal{T}(f)(B_1A_1) = 0$. Therefore, $\gamma\notin \Delta$ and the lemma is proved.
\end {proof}
Lemma \ref{goodx} allows us to formulate the following special case of the above result.
\begin {lem}\label{spprod}
Let $A\in L(\mathcal{X},\mathcal{Y})$, $B\in L(\mathcal{Y},\mathcal{Z})$ and $(L(\mathcal{X},\mathcal{Y}), \mathcal{T}_{\mathcal{X}\mathcal{Y}})$, $(L(\mathcal{Y},\mathcal{Z}), \mathcal{T}_{\mathcal{Y}\mathcal{Z}})$, and
$(L(\mathcal{X},\mathcal{Z}), \mathcal{T}_{\mathcal{X}\mathcal{Z}})$ be non-degenerate Banach modules over the algebras $L_{\nu_1}(\mathbb{G})$, $L_{\nu_2}(\mathbb{G})$, and $L_{\nu_3}(\mathbb{G})$, respectively, where $\nu_i$, $i=1,2,3$, are non-quazi-analytic weights. Assume also that the above three algebras contain b.a.i., \eqref{couple} is satisfied, and that
$A$ and $B$ are $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$- and $\mathcal{T}_{\mathcal{Y}\mathcal{Z}}$-continuous, respectively. Then
\[\L(BA,\mathcal{T}_{\mathcal{X}\mathcal{Z}})\subseteq\overline{\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})+\L(B,\mathcal{T}_{\mathcal{Y}\mathcal{Z}})}.\]
\end {lem}
The following corollary appears, for example, in \cite{BK05, BR75}.
\begin {cor}\label{spprodv}
Let $A \in (L(\mathcal{X},\mathcal{Y}),\mathcal{T}_{\mathcal{X}\mathcal{Y}})_c$, where $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$ is defined as in Example \ref{typop} and $x\in(\mathcal{X},\mathcal{T}_\mathcal{X})$. Then
\[\L(Ax,\mathcal{T}_{\mathcal{Y}})\subseteq\overline{\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})+\L(x,\mathcal{T}_{\mathcal{X}})}.\]
\end {cor}
To help the reader, we note that the above is a generalization of the fact that a product of a one-diagonal matrix with the $m$-th non-zero diagonal and a one-diagonal matrix with the
$n$-th non-zero diagonal is either $0$ or another one-diagonal matrix with the $(m+n)$-th non-zero diagonal.
\begin {rem}\label{ba}
Analogous results hold if $A$, $B$ belong to a Banach algebra $\mathfrak B$, which has a non-degenerate $L_\nu(\mathbb{G})$-Banach module with respect to a strongly continuous representation $\mathcal{T}$ satisfying an obvious analog of \eqref{couple}, i.e. when $\{\mathcal{T}(g),\ g\in\mathbb{G}\}$ is a group of automorphisms of the algebra $\mathfrak B$. In fact, most of the results about linear operators proved in this paper remain valid in the case of Banach algebras with a group of automorphisms and the proofs apply nearly verbatim. We shall discuss several examples in the last section of the paper.
\end {rem}
We conclude this section with the following obvious, but useful, observation.
\begin {prop}\label{cinv}
Let $A \in (L(\mathcal{X},\mathcal{Y}),\mathcal{T}_{\mathcal{X}\mathcal{Y}})_c$, where $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$ is defined as in Example \ref{typop}. Then
$A\mathcal{X}_c\subseteq\mathcal{X}_c$. If, moreover, $A$ is invertible then $A^{-1}\in(L(\mathcal{Y},\mathcal{X}),\mathcal{T}_{\mathcal{Y}\mathcal{X}})_c$.
\end {prop}
\section{Memory decay for inverses of operators with compact Beurling spectrum}\label{Wcomp}
We begin with the simplest case of $\mathbb G = {\mathbb R}$ and $\nu \equiv 1$. We consider an
invertible operator $A\in L(\mathcal{X},\mathcal{Y})$ with $\L(A) =\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})$ compact. We let $B = A^{-1}\in L(\mathcal{Y},\mathcal{X})$
and study the memory decay of $B$ with respect to the representation $\mathcal{T}_{\mathcal{Y}\mathcal{X}}$. Because
$\nu \equiv 1$ the representations $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$ and $\mathcal{T}_{\mathcal{Y}\mathcal{X}}$ are assumed to be isometric.
\begin {thm}\label{wcompact}
Assume $A \in L(\mathcal{X},\mathcal{Y})$ is invertible, $\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})\subset[-a,a]$, and $B = A^{-1}$. Then there is $\alpha > 0$ such that $B$ has exponential spectral decay of type $\alpha$, that is
$B\in D(\textbf{\emph{h}}_\a, \mathcal{T}_{\mathcal{Y}\mathcal{X}})$. Moreover, if $\phi_N \in L_1({\mathbb R})$, $N\in{\mathbb N}$, are
such that $\norm{\phi_N}=1$,
$\hat\phi_N$ is continuous, $\text{ supp\!}\hat\phi_N\subset [-N,N]$, and $\phi_{N,n}(x) = \phi_N(x)e^{ i Nnx}$, then
\begin {equation}\label{infest}
\|\phi_{N,n} B\|\le \inf_{\a\in(0,\,
a^{-1}\ln(1+\ae^{-1}(A))}\frac{e^{\alpha N(1-|n|))}
\|B\|}{1-(e^{\a a}-1)\mbox{\rm \ae}(A)}, \ |n|>1,
\end{equation}
where $\mbox{\rm \ae}(A) = \|A\|\|B\|$ is the condition number of $A$.
\end {thm}
\begin {proof}
Observe that since $\L(A)\subset [-a,a]$ Corollary \ref{Bineq} implies that for every $z\in {\mathbb C}$ we have $A\in D(e_z)$, where $e_z(\l) = e^{iz\l}$, and $A_\mathcal{T}: z\mapsto e_z\diamond A$ is an entire function.
Since invertibility is stable under small perturbations and the representation $\mathcal{T}$ is isometric, we get that $A_\mathcal{T}(z)$ is invertible for $z\in{\mathbb C}_{-\d,\d}$ for some $\d > 0$. Clearly, then $(e_z\diamond A)^{-1}$ is a holomorphic extension of $B_\mathcal{T}$ and Lemma \ref{2side} implies $B\in D({\textbf{h}}_{\alpha})$ for $0<\a<\d$. Obtaining the estimate \eqref{infest}, however, requires more work.
For $z = t + i\a$,
$\a>0$,
let us estimate the norm $\|e_{z}\diamond A - \mathcal{T}_{\mathcal{X}\mathcal{Y}}(t)A\| = \|\textbf{e}_{-\a}\diamond A - A\|$. Consider
the $4a$-periodic function $h(\l) = \eta(\l)-1$, where $\eta$ is defined by \eqref{curvehat}. We have shown in \eqref{fcoef} that
the Fourier coefficients $c_n$, $n\in{\mathbb Z}\backslash\{0\}$, of the periodic even function $\eta(\cdot - a)$ are positive. Clearly, they coincide with non-zero Fourier coefficients of $h(\cdot - a)$. The $0$-th coefficient is also positive since
$c_0 = \frac4\a(\sinh \a a-\a a)>0$. Hence, by Proposition \ref{inmeas}, $\|\check\mathcal{T}_{\mathcal{X}\mathcal{Y}}(h)\| \le
e^{\a a}-1$.
Moreover, $h \equiv \textbf{e}_{-\a} -1\equiv e_{i\a}-1$
on $[-a,a]\supseteq\L(A)$ and, hence, by Lemma \ref{sssprop}
\[\|\textbf{e}_{-\a}\diamond A - A\| = \|\check\mathcal{T}_{\mathcal{X}\mathcal{Y}}(h) A\|
\le\|\check\mathcal{T}_{\mathcal{X}\mathcal{Y}}(h)\|\|A\| = (e^{\a a}-1)\|A\|.\]
Using the same argument for the representation $\widetilde{\mathcal{T}}_{\mathcal{X}\mathcal{Y}}(t) = \mathcal{T}_{\mathcal{X}\mathcal{Y}}(-t)$,
we get
\[\|\check\mathcal{T}(\textbf{e}_{\a}) A - A\| = \|\textbf{e}_{\a}\diamond A - A\| \le (e^{\a a}-1)\|A\|,\ \a>0.\]
Therefore, if $|\a| < a^{-1}\ln(1+\mbox{\rm \ae}^{-1}(A))$, we have that
\[(\textbf{e}_{\a} \diamond A)^{-1} = B(I+(\textbf{e}_{\a} \diamond A-A)B)^{-1} = B\sum_{n=0}^\infty (-1)^n ((\textbf{e}_{\a} \diamond A-A)B)^n,\]
$B_\mathcal{T}(z): = (e_z \diamond A)^{-1} = \mathcal{T}_{\mathcal{Y}\mathcal{X}}(t)(\textbf{e}_{\mp\a} \diamond A)^{-1}$, $z=t\pm i\a$,
$\a > 0$, satisfies
\[\|B_\mathcal{T}(z)\| \le \frac{\|B\|}{1-(e^{\a a}-1)\mbox{\rm \ae}(A)}, \]
and the estimate \eqref{infest} follows from \eqref{expest}.
\end {proof}
Next, we extend the above result to the case when $\mathbb G = {\mathbb R}^d$.
\begin {thm}
Assume $A \in L(\mathcal{X},\mathcal{Y})$ is invertible and $B = A^{-1}$.
Assume also that $a = (a_1,\ldots, a_d)\in{\mathbb R}^d_+$ and $\L(A,\mathcal{T}_{\mathcal{X}\mathcal{Y}})\subset[-a_1,a_1]\times\ldots\times[-a_d,a_d]$. Then there is $\alpha = (\a_1,\ldots,\a_d) \in{\mathbb R}^d_+$ such that the memory of $B$ has exponential decay of type $\alpha$, that is
\emph{$B\in D(\textbf{{h}}_\a, \mathcal{T}_{\mathcal{Y}\mathcal{X}})$}. Moreover, if $\phi^d_{N,n} \in L_1({\mathbb R}^d)$, $N\in{\mathbb N}$, $n\in{\mathbb Z}^d$, are defined via \eqref{triangled}, then
\begin {equation}\label{infestd}
\|\phi_{N,n} B\|\le \inf_{ \a
}\frac{e^{ NM_n(\a)}
\|B\|}{1-\left(e^{\a\cdot a}-1\right)\mbox{\rm \ae}(A)}, \ n\in {\mathbb Z}^d,
\end{equation}
where
$M_n(\a) = \sum\limits_{k:\ n_k\ne 0} \a_k(1-|n_k|)$ and the infimum is taken over all $\a\in{\mathbb R}^d$ such that the denominator is positive.
\end {thm}
\begin {proof}
The proof of the one-dimensional case can be applied almost verbatim if the periodic function $\eta$ is replaced with
$\eta^d(\l) = \prod_{k=1}^d \eta(\l_k)$, $\l = (\l_1,\ldots, \l_d)\in{\mathbb R}^d$.
Remark \ref{tenp} and Lemmas \ref{ssprop} and \ref{sssprop}, however, cannot be used in this case. To overcome this minor obstacle one uses Lemma \ref{sprop} to prove the estimates with $a+\varepsilon$, $\varepsilon\in{\mathbb R}^d_+$, in place of $a$. The desired estimates then follow since $\varepsilon\in{\mathbb R}^d_+$ is arbitrary.
We omit the remaining details for the sake of brevity.
\end {proof}
\section{Memory decay of inverses to causal operators and operators with exponential memory decay}\label{Wcaus}
In this section we extend the above results to operators with exponential spectral decay as well as to causal operators and operators with one-sided exponential spectral decay. As before the proof is based on the holomorphic extensions of the function $t\mapsto \mathcal{T}(t)A$. We begin, however, with establishing the algebraic properties of the classes of operators considered in this section.
\begin {lem}\label{causalg}
Assume $A\in\mathcal C(L(\mathcal{Y},\mathcal{Z}))$ and $B\in\mathcal C(L(\mathcal{X},\mathcal{Y}))$. Then $AB\in\mathcal C(L(\mathcal{X},\mathcal{Z}))$.
\end {lem}
\begin {proof}
Follows from Lemma \ref{spprod}.
\end {proof}
\begin {lem}\label{prodexp1}
Let $\a\in{\mathbb R}^d$. Assume $A\in L(\mathcal{Y},\mathcal{Z})_c\cap D(\emph{\textbf{e}}_{\a}, \mathcal{T}_{\mathcal{Y}\mathcal{Z}})$ and
$B\in L(\mathcal{X},\mathcal{Y})_c\cap D(\emph{\textbf{e}}_{\a}, \mathcal{T}_{\mathcal{X}\mathcal{Y}})$. Then $AB\in L(\mathcal{X},\mathcal{Z})_c\cap D(
\emph{\textbf{e}}_{\a}, \mathcal{T}_{\mathcal{X}\mathcal{Z}})$ and
\begin {equation}\label{hom1side}
\emph{\textbf{e}}_{\a}\diamond(AB) = (\emph{\textbf{e}}_{\a}\diamond A)( \emph{\textbf{e}}_{\a}\diamond B).
\end{equation}
\end {lem}
\begin {proof}
The proof follows by uniqueness of the holomorphic extension provided by Proposition \ref{1expestprop}.
\end {proof}
\begin {thm}\label{1sidexpwin}
Let $\a\in{\mathbb R}^d_\pm$ and $A\in L(\mathcal{X},\mathcal{Y})_c\cap D(\emph{\textbf{e}}_{\a})$. Assume
$B = A^{-1} \in L(\mathcal{Y},\mathcal{X})$. Then there is $\beta\in {\mathbb R}^d_\pm$ such that
$B\in L(\mathcal{Y},\mathcal{X})_c\cap D(\emph{\textbf{e}}_{\beta})$.
\end {thm}
\begin {proof}
All we need to do is repeat the first part of the proof of Theorem \ref{wcompact}. Indeed, if $A_\mathcal{T}$ is
the holomorphic extension provided by Proposition \ref{1expestprop}, it is continuous in the uniform operator topology in
$\overline{{\mathbb C}_{\mp\d}}$ for some $\d\in{\mathbb R}^d_+$. Hence, because the representation $\mathcal{T}$ is bounded and invertibility is
stable under small perturbations, there exists $\beta \in {\mathbb R}^d_+$ such that $A_\mathcal{T}(z)$ is invertible for $z\in\overline{{\mathbb C}_{\mp\beta}}$. It remains only to apply Proposition \ref{1expestprop} once again.
\end {proof}
The following two results are immediate consequences of Theorem \ref{1sidexpwin}. A version of the first of them was originally announced in \cite{BK06}.
\begin {thm}\label{causwin}
Assume $A\in L(\mathcal{X},\mathcal{Y})_c\cap \mathcal C_\pm(L(\mathcal{X},\mathcal{Y}))$ and $B = A^{-1}\in L(\mathcal{Y},\mathcal{X})$. Then there is $\a\in {\mathbb R}^d_+$ such that
$B\in L(\mathcal{Y},\mathcal{X})_c\cap D(\emph{\textbf{e}}_{\mp\a})$.
\end {thm}
\begin {thm}\label{invexp}
Let $\a\in{\mathbb R}^d_+$ and $A\in L(\mathcal{X},\mathcal{Y})_c\cap D(\emph{\textbf{h}}_{\a})$. Assume
$B = A^{-1} \in L(\mathcal{Y},\mathcal{X})$. Then there is $\beta\in {\mathbb R}^d_+$ such that
$B\in L(\mathcal{Y},\mathcal{X})_c\cap D(\emph{\textbf{h}}_{\beta})$.
\end {thm}
\begin {rem}
Guided by the fact that the inverse of a triangular matrix is also triangular, one might think that the inverse to a causal operator should always be causal. The bilateral shift operator shows, however, that the inverse to a causal operator may, in fact, be anticausal. Theorem \ref{causwin}, on the other hand, ensures that the anticausal part of the inverse always has exponential memory decay. We refer to \cite{BK05} for different criteria of causal invertibility.
\end {rem}
\begin {rem}
Given an invertible operator $A\in D({\textbf{e}}_\alpha )$ we cannot obtain estimates for the memory decay
of $A^{-1}$ without assuming more than $A\in L(\mathcal{X},\mathcal{Y})_c$. It is possible, however, to use the approach of Theorem \ref{wcompact} to obtain such estimates if we can quantify the convergence of
$\mathcal{T}(t)A\to A$ as $t\to 0$. In particular, we can get the estimates if $A$ is in one of the H\"older-Zygmund classes studied in \cite{GK10}.
\end {rem}
\section{Extension of the classical Wiener's Lemma}\label{Wclass}
We begin by establishing the algebraic property of the Wiener Class $\mathcal{W} = \mathcal{W}(L(\mathcal{X},\mathcal{Y}))$ of operators in $L(\mathcal{X},\mathcal{Y})$. Recall from Section \ref{wiencl} that the norms of operators $A\in\mathcal{W}$ are given by
\begin {equation}\label{wiensum}
\|A\|_{1,N} = 5^d \sum_{n\in{\mathbb Z}^d} \|\phi^d_{N,n} A\|= 5^d \sum_{n\in{\mathbb Z}^d} \|\mathcal{T}_{\mathcal{X}\mathcal{Y}}(\phi^d_{N,n}) A\|,
\ N\in{\mathbb N},
\end{equation}
where we used the functions $\phi_N^d$ defined by
\eqref{triangled} and their translates $\phi^d_{N,n}$.
\begin {lem}\label{algprop}
If
$B\in\mathcal{W}(L(\mathcal{X},\mathcal{Y}))$ and $A\in\mathcal{W}(L(\mathcal{Y},\mathcal{Z}))$ then $AB\in\mathcal{W}(L(\mathcal{X},\mathcal{Z}))$ and
\[\|AB\|_{1,N}\le\|A\|_{1,N}\|B\|_{1,N},\]
in particular, $\mathcal{W}(B(\mathcal{X}))$ is a Banach algebra.
\end {lem}
\begin {proof}
We deduce the desired property from
\begin{equation*
\begin{split}
\|AB\|_{1,N} & = 5^d \sum_{n\in{\mathbb Z}^d} \|\phi^d_{N,n} (AB)\| \\ &=
5^d \sum_{n\in{\mathbb Z}^d} \left\|\phi^d_{N,n} \left(\left(\sum_{m\in{\mathbb Z}^d} \phi^d_{N,m}A\right)
\left(\sum_{k\in{\mathbb Z}^d} \phi^d_{N,k}B\right)\right)\right\| \\ &
\le 5^{2d} \sum_{m\in{\mathbb Z}^d} \|\phi^d_{N,m} A\|\sum_{k\in{\mathbb Z}^d} \|\phi^d_{N,k} B\| = \|A\|_{1,N}\|B\|_{1,N},
\end{split}
\end{equation*}
where the inequality is true because for every choice of $m,n\in{\mathbb Z}^d$ there are at most $5^d$ different numbers $k\in{\mathbb Z}^d$ such that the set $\text{ supp\!} \hat\phi^d_{N,k}\cap \L((\phi^d_{N,m}A)( \phi^d_{N,n}B),\mathcal{T}_{\mathcal{X}\mathcal{Z}})$ has non-empty interior. We used Lemmae \ref{ssprop}, \ref{spprod} and Remark \ref{tenp}.
\end {proof}
\begin {thm}\label{maint}
Let $A\in \mathcal{W}(L(\mathcal{X},\mathcal{Y}))$ be invertible with $B = A^{-1} \in L(\mathcal{Y},\mathcal{X})$. Then $B\in \mathcal{W}(L(\mathcal{Y},\mathcal{X}))$ and
\begin {equation}\label{wienest}
\|B\|_{1,1} \le (2N+1)^d\|B\|_{1, N} \le\frac{\|B\|}{\epsilon_A}\left(2\psi_A\left(\frac{\epsilon_A}{\|B\|}\right)+1\right)^d,
\end{equation}
where $N = \psi_A(\frac{\epsilon_A}{\|B\|})$, $\epsilon_A = 5^{-d}\left(16\left(\frac{2\d_A - 1}{\d_A-1}\right)^d-12\right)^{-1}$,
$\d_A = \left(\frac{4\ae(A)+3}{4\ae(A)+2}\right)^{\frac1{3d}}$, $\mbox{\rm \ae}(A) = \|A\|\|A^{-1}\|$, and $\psi_A: {\mathbb R}_+\to {\mathbb N}$ is defined by
\[
\psi_A\left(t\right) = \min \left\{K\in{\mathbb N}:\ \norm{A - \sum_{|k|_\infty \le 2} \phi^d_{K,k}A}_{1,K} \le t\right\},
\]
where $\phi^d_{K,k}$ is given by \eqref{triangled}.
\end {thm}
\begin {proof}
We begin by fixing some $N\in{\mathbb N}$, which is to be determined later, and using the functions
$\phi_{N,n}^d$ defined via \eqref{triangled} to represent $A = C+D$, $C = \sum\limits_{|n|_\infty \le 2} \phi^d_{N,n}A$,
in such a way that
$\|B\|\|D\|_{1,N} \le\frac12$ and, hence, also $\|B\|\|D\| \le\frac12$. The first inequality is true for some $N\in{\mathbb N}$ because
$A\in \mathcal{W}(L(\mathcal{X},\mathcal{Y}))$. In this case, $\|C\|\le \|A\|+\frac1{2\|B\|}$, $C$ is invertible, and
$L = C^{-1}=(A-D)^{-1} = B(I-DB)^{-1} = B\sum_{n=0}^\infty (DB)^n$ satisfies
\begin {equation}
\|L\|\le \frac{\|B\|}{1-\|B\|\|D\|} \le 2 \|B\|.
\end{equation}
Using \eqref{infestd} for the operators $C$ and $L = C^{-1}$ with $a= (3N,\ldots,3N)\in{\mathbb R}^d$,
we get
\begin {equation}
\begin {split}
\|\phi^d_{N,n}L\| &\le \inf_{\a} \frac{e^{\a NM_n}\|L\|}{1-(e^{3\a d N}-1)\mbox{\rm \ae}(C)} \\ &\le
\inf_{\a} \frac{e^{\a NM_n}\|B\|}{1-\|B\|\|D\|-(e^{3\a dN}-1)(\mbox{\rm \ae}(A)+\frac12)},
\\ &\le \inf_{\a} \frac{e^{\a NM_n}\|B\|}{\frac12-(e^{3\a dN}-1)(\mbox{\rm \ae}(A)+\frac12)},
\end{split}
\end{equation}
where $M_n = \sum\limits_{k=1}^d 1-|n_k|$, $n=(n_1,\ldots, n_d)\in{\mathbb Z}^d$, $\ n_k\ne 0$, and the infima are taken over all $\a\in{\mathbb R}_+$ such that the corresponding denominator is positive.
Next, we choose $\a$ such that
\[\frac12-(e^{3\a dN}-1)(\mbox{\rm \ae}(A)+\frac12) = \frac14.\]
Then $e^{\a N} = \left(\frac{4\ae(A)+3}{4\ae(A)+2}\right)^{\frac1{3d}}=:\d_A > 1$ and, therefore,
\begin {equation}
\|\phi^d_{N,n}L\| \le 4\d_A^{M_n}\|B\|, \ n\neq 0, \mbox{ and } \|\phi^d_{N,0} L\|\le\|L\|\le 2\|B\|.
\end{equation}
Summing over $n\in{\mathbb Z}^d$ and multiplying by $5^d$, we get
\begin {equation}
\|L\|_{1,N} \le 5^d\left(8\left(\frac{2\d_A - 1}{\d_A-1}\right)^d-6\right)\|B\| = \frac1{2\epsilon_A}\|B\|.
\end{equation}
Finally, we specify some $N\in{\mathbb N}$ such that
\[
\|L\|_{1,N}\|D\|_{1,N} \le \frac1{2\epsilon_A}\|B\|\|D\|_{1,N} \le\frac12,
\]
that is, we let
$N = \psi_A\left( \frac{\epsilon_A}{\|B\|}\right)$. Then the series in $B = (C + D)^{-1} = L(I+DL)^{-1} = L\sum_{n=0}^\infty (-1)^{n}(DL)^n$ converges in $\mathcal{W}(B(\mathcal{Y}))$, and,
using
\[\|B\|_{1,N} \le\frac{\|L\|_{1,N}}{1-\|L\|_{1,N}\|D\|_{1,N} }\le \frac{\|B\|}{\epsilon_A},\]
we obtain the estimate \eqref{wienest}.
\end {proof}
\begin {prop}\label{invarw}
Assume $A\in\mathcal{W}(L(\mathcal{X},\mathcal{Y}),\mathcal{T}_{\mathcal{X}\mathcal{Y}})$, where the representation $\mathcal{T}_{\mathcal{X}\mathcal{Y}}: {\mathbb R}^d\to B(L(\mathcal{X},\mathcal{Y}))$ is defined as in Example \ref{typop} by $\mathcal{T}_{\mathcal{X}\mathcal{Y}}(t)A = \mathcal{T}_\mathcal{Y}(t)A\mathcal{T}_\mathcal{X}(-t)$. Then
\[A\mathcal{W}(\mathcal{X},\mathcal{T}_\mathcal{X}) \subseteq \mathcal{W}(\mathcal{Y},\mathcal{T}_\mathcal{Y}).\]
Moreover, if $A$ is invertible then it is also an isomorphism between $\mathcal{W}(\mathcal{X},\mathcal{T}_\mathcal{X})$ and $\mathcal{W}(\mathcal{Y},\mathcal{T}_\mathcal{Y})$.
\end {prop}
\begin {proof}
Let $x\in \mathcal{W}(\mathcal{X},\mathcal{T}_\mathcal{X})$ and $y = Ax$. As in Lemma \ref{algprop}, we get
\begin{equation*
\begin {split}
\sum_{k\in{\mathbb Z}^d} \|\phi^d_{N,k}y \| &=
\sum_{k\in{\mathbb Z}^d} \left\|\phi^d_{N,k} \left(\left(\sum_{m\in{\mathbb Z}^d} \phi^d_{N,m}A\right)
\left(\sum_{n\in{\mathbb Z}^d} \phi^d_{N,n}x\right)\right)\right\| \\
& \le 5^{d} \sum_{m\in{\mathbb Z}^d} \|\phi^d_{N,m} A\|\sum_{n\in{\mathbb Z}^d} \|\phi^d_{N,n} x\| = \|A\|_{1,N}\|x\|_{1,N},
\end{split}
\end{equation*}
where we used Corollary \ref{spprodv}.
\end {proof}
\begin {rem}
Various corollaries of the above proposition play a crucial role in the study of canonical duals of localized frames (see, e.g. \cite{BCHL06I, F09, G04, KO08} and references therein).
\end {rem}
\section{Inverse Closedness of the Beurling Class}\label{Wbeur}
The Beurling Class ${\mathcal B} = {\mathcal B}(\mathcal{X},\mathcal{Y})$ of operators in $L(\mathcal{X},\mathcal{Y})$ has the norm
\begin {equation}\label{bnormop}
\|A\|_{{\mathcal B}}= \sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty} \|\phi^d_{1,n}A\| =
\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty}\|\mathcal{T}_{\mathcal{X}\mathcal{Y}}(\phi^d_{1,n})A\|,
\end{equation}
where $|k|_\infty = \max\{|k_1|, |k_2|,\ldots, |k_d|\}$, $k = (k_1, k_2,\ldots, k_d)\in{\mathbb R}^d$.
We prove the inverse closedness of this class by adapting a similar argument in \cite{S10ca} to our general setting and applying the technique of the previous section.
We begin with the following
very strong algebraic property.
\begin {lem}
If
$B\in{\mathcal B}(\mathcal{X},\mathcal{Y})$ and $A\in{\mathcal B}(\mathcal{Y},\mathcal{Z})$ then $AB\in{\mathcal B}(\mathcal{X},\mathcal{Z})$ and
\begin {equation}\label{brand}
\|AB\|_{{\mathcal B}}\le 5^d\|A\|_{1,1}\|B\|_{{\mathcal B}}+2^d\|A\|_{{\mathcal B}}\|B\|_{1,1},
\end{equation}
in particular, ${\mathcal B}(\mathcal{X})={\mathcal B}(\mathcal{X},\mathcal{X})$ is a Banach algebra with respect to an equivalent norm.
\end {lem}
\begin {proof}
Using the absolute convergence of the series below, we have
\begin{equation*
\begin{split}
\|\phi_{1,n}(AB)\| & =\left\|\phi_{1,n}\left(\left(\sum_{i\in{\mathbb Z}^d} \phi_{1,i}A\right)\left(\sum_{j\in{\mathbb Z}^d}
\phi_{1,j}B\right)\right)\right\| \\
& = \left\|\sum_{i\in{\mathbb Z}^d}\sum_{j\in{\mathbb Z}^d}\phi_{1,n}\left(\left(\phi_{1,i}A\right)\left(
\phi_{1,j-i}B\right)\right)\right\| \\
\le \left\|\sum_{|i|_\infty \ge\frac12|n|_\infty}\sum_{j\in{\mathbb Z}^d}\phi_{1,n}\left(\left(\phi_{1,i}A\right)\left(
\phi_{1,j-i}B\right)\right)\right\| \\ &+ \left\|\sum_{|i|_\infty <\frac12|n|_\infty}\sum_{j\in{\mathbb Z}^d}\phi_{1,n}\left(\left(\phi_{1,i}A\right)\left(\phi_{1,j-i}B\right)\right)\right\|
\end{split}
\end{equation*}
From Lemma \ref{spprod} we deduce that when $|j-n|_\infty >2$ the terms in both series vanish. Hence,
\begin{equation*
\begin {split}
\|\phi_{1,n}(AB)\| &\le \sum_{|i|_\infty \ge\frac12|n|_\infty}\sum_{|j-n|_\infty \le 2}\|\phi_{1,i}A\| \| \phi_{1,j-i}B\| \\
& +\sum_{|j-n|_\infty \le 2}\sum_{|j-i|_\infty <\frac12|n|_\infty}\|\phi_{1,j-i}A\| \|\phi_{1,i}B\|
\\
& \le \sum_{|i|_\infty \ge\frac12|n|_\infty}\sum_{|j-n|_\infty \le 2}\|\phi_{1,i}A\| \| \phi_{1,j-i}B\| \\
& +\sum_{|i|_\infty \ge\frac12|n|_\infty-1}\sum_{|j-n|_\infty \le 2}\|\phi_{1,j-i}A\| \|\phi_{1,i}B\|.
\end{split}
\end{equation*}
Therefore,
\begin{equation*
\begin {split}
\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty} & \|\phi_{1,n}(AB)\| \\ &\le
\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty} \sum_{|i|_\infty \ge\frac12|n|_\infty}\sum_{|j-n|_\infty \le 2}\|\phi_{1,i}A\| \| \phi_{1,j-i}B\| \\
& +\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty}\sum_{|i|_\infty \ge\frac12|n|_\infty-1}\sum_{|j-n|_\infty \le 2}\|\phi_{1,j-i}A\| \|\phi_{1,i}B\| \\
&\le \sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge \frac { |k|_\infty}2}\|\phi_{1,n}A\|\sum_{i\in{\mathbb Z}^d}5^d\|\phi_{1,i}B\| \\
&+\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge \frac { |k|_\infty}2-1}\|\phi_{1,n}B\|\sum_{i\in{\mathbb Z}^d}5^d\|\phi_{1,i}A\|,
\end{split}
\end{equation*}
and the desired inequality follows. Since $\|A\|_\mathcal{W}\le 5^d\|A\|_{\mathcal B}$, the inequality implies that ${\mathcal B}(\mathcal{X})$ is indeed a Banach algebra.
\end {proof}
Next we use the standard Brandenburg trick \cite{B75} to show that the spectral radii in the Wiener and
Beurling algebras coincide, i.e., for all $A\in {\mathcal B}(\mathcal{X})$
\begin {equation}\label{radii}
\rho_{\mathcal B}(A) = \lim_{n\to\infty}\|A^n\|^{1/n}_{{\mathcal B}} = \lim_{n\to\infty}\|A^n\|^{1/n}_{\mathcal{W}} = \rho_\mathcal{W}(A).
\end{equation}
Indeed, since $\|A\|_\mathcal{W}\le 5^d\|A\|_{\mathcal B}$, we have $\rho_\mathcal{W}(A)\le\rho_{\mathcal B}(A)$. On the other hand,
from \eqref{brand} we get $\|A^{2n}\|_{\mathcal B} \le (5^d+2^d)\|A^n\|_{\mathcal B}\|A^n\|_{1,1}$. Taking the $n$-th root and
passing to the limit we get $\rho_{\mathcal B}(A^2) = (\rho_{\mathcal B}(A))^2 \le \rho_{\mathcal B}(A)\rho_\mathcal{W}(A)$ which implies
\eqref{radii}.
At this point, Hulanicki's lemma \cite{H72} is typically used to finish the argument about inverse closedness \cite{G10, GK10, S10}. In our setting, however, this result is not applicable since we may be entirely outside of the realm of $*$-algebras. We overcome this obstacle by looking more closely into the proof of the result in the previous section.
\begin {thm}\label{absB}
Assume that $A\in{\mathcal B}(\mathcal{X},\mathcal{Y})$ and $B\in L(\mathcal{Y},\mathcal{X})$ is its inverse operator. Then $B\in{\mathcal B}(\mathcal{Y},\mathcal{X})$.
\end {thm}
\begin {proof}
As in the proof of Theorem \ref{maint}, let $A = C + D$ where $\L(C)$ is compact and
\[B = C^{-1}(I+DC^{-1})^{-1} = C^{-1}\sum_{k=0}^\infty (-DC^{-1})^n,\]
where the Neumann series converges in $\mathcal{W}(\mathcal{Y})$ because $\|DC^{-1}\|_\mathcal{W} < 1$.
Theorem \ref{wcompact} implies $C^{-1}\in D({\textbf{h}}_{\alpha})\subseteq {\mathcal B}$ and we can use \eqref{radii} to conclude that $\rho_{\mathcal B}(DC^{-1}) < 1$ and, hence, the Neumann series converges in ${\mathcal B}(\mathcal{Y})$.
\end {proof}
The proof of the following result is analogous to that of proposition \ref{invarw} and is left to the reader.
\begin {prop}\label{invarb}
Assume $A\in{\mathcal B}(L(\mathcal{X},\mathcal{Y}),\mathcal{T}_{\mathcal{X}\mathcal{Y}})$, where the representation $\mathcal{T}_{\mathcal{X}\mathcal{Y}}: {\mathbb R}^d\to B(L(\mathcal{X},\mathcal{Y}))$ is defined as in Example \ref{typop} by $\mathcal{T}_{\mathcal{X}\mathcal{Y}}(t)A = \mathcal{T}_\mathcal{Y}(t)A\mathcal{T}_\mathcal{X}(-t)$. Then
\[A{\mathcal B}(\mathcal{X},\mathcal{T}_\mathcal{X}) \subseteq {\mathcal B}(\mathcal{Y},\mathcal{T}_\mathcal{Y}).\]
Moreover, if $A$ is invertible then it is also an isomorphism between ${\mathcal B}(\mathcal{X},\mathcal{T}_\mathcal{X})$ and ${\mathcal B}(\mathcal{Y},\mathcal{T}_\mathcal{Y})$.
\end {prop}
\section{Sobolev-type Algebras}\label{sobol}
Here we extend the previous results to the Sobolev-type classes $\mathcal{W}^{(m)}$ and ${\mathcal B}^{(m)}$.
Unlike the previous two sections, rather than deal with operators in $L(\mathcal{X},\mathcal{Y})$ we choose to work with a unital Banach algebra $\mathfrak B$ such that the representation $\mathcal{T}$ defines a group of isometric automorphisms of $\mathfrak B$. It should be clear how to modify the statements to get the results for $L(\mathcal{X},\mathcal{Y})$. The definitions of $\mathcal{W}^{(m)}$ and ${\mathcal B}^{(m)}$ were
given in Subsection \ref{sclass}; in what follows we are using the same notation.
In pursuing our extension, we follow the ideas in \cite{BR75, GK10}. We note, however, that since
our Wiener and Beurling classes are different from those considered in \cite{GK10}, the Sobolev-type classes are also different, and the results in this section do not follow immediately from \cite{GK10}.
Observe that since each $\mathcal{T}^{(k)}(t)$ defined in Remark \ref{tenp} is an automorphism of $\mathfrak B$, the definition of the infinitesimal generator $A_k$ implies
\begin {equation}\label{deriv}
A_k(xy) = (A_k x)y + x (A_k y), \ x,y\in D(A_k),
\end{equation}
i.e. each $A_k$, $k = 1,\ldots, d$, is a derivation \cite{BR75} on $\mathfrak B$.
These equalities ensure that both $\mathcal{W}^{(m)}$ and ${\mathcal B}^{(m)}$ are Banach algebras with the norm
\[\|x\|_{\mathcal{F}^{(m)}} = \sum_{|\a|\le m} \|A^\a x\|_\mathcal{F},\ x\in\mathcal{F},\]
where $\mathcal{F}$ is either $\mathcal{W}$ or ${\mathcal B}$. We refer to these algebras as Sobolev-Wiener or Sobolev-Beurling algebras, respectively.
\begin {thm}\label{sobolt}
Let $\mathcal{F}^{(m)}$, $m\in{\mathbb N}$, be either a Sobolev-Wiener or a Sobolev-Beurling algebra. Assume that
$x \in\mathcal{F}^{(m)}$ is invertible, i.e. $x^{-1}\in \mathfrak B$. Then $x^{-1}\in \mathcal{F}^{(m)}$, i.e. $\mathcal{F}^{(m)}$ is an inverse-closed subalgebra of $\mathfrak B$.
\end {thm}
\begin {proof}
Observe that $A_k e = 0$ for the unit $e\in\mathfrak B$ and $D(A_k)$ is inverse closed since the function $t\mapsto\mathcal{T}^{(k)}(t)x^{-1} = (\mathcal{T}^{(k)}(t)x)^{-1}$ is differentiable when $x\in D(A_k)$ is invertible in $\mathfrak B$
(see also \cite{BR75, GK10}). Therefore, using equalities \eqref{deriv} we get
$0 = A_k(xx^{-1}) = (A_k x)x^{-1} + x(A_k x^{-1})$ and, hence,
\[A_k x^{-1} = -x^{-1}(A_k x) x^{-1}.\]
The result is now an immediate consequence of Theorems \ref{maint}, and \ref{absB}.
\end {proof}
\begin {prop}\label{invarsvb}
Assume $C\in\mathcal{F}^{(m)}(L(\mathcal{X},\mathcal{Y}),\mathcal{T}_{\mathcal{X}\mathcal{Y}})$, where the representation $\mathcal{T}_{\mathcal{X}\mathcal{Y}}: {\mathbb R}^d\to B(L(\mathcal{X},\mathcal{Y}))$ is defined as in Example \ref{typop} by $\mathcal{T}_{\mathcal{X}\mathcal{Y}}(t)C = \mathcal{T}_\mathcal{Y}(t)a\mathcal{T}_\mathcal{X}(-t)$. Then
\[C\mathcal{F}^{(m)}(\mathcal{X},\mathcal{T}_\mathcal{X}) \subseteq \mathcal{F}^{(m)}(\mathcal{Y},\mathcal{T}_\mathcal{Y}).\]
Moreover, if $C$ is invertible then it is also an isomorphism between $\mathcal{F}^{(m)}(\mathcal{X},\mathcal{T}_\mathcal{X})$ and $\mathcal{F}^{(m)}(\mathcal{Y},\mathcal{T}_\mathcal{Y})$.
\end {prop}
\begin {proof}
Let $A^{\mathcal{X}}$, $A^{\mathcal{Y}}$ and $A^{\mathcal{X}\mathcal{Y}}$ be the multigenerators of the representations $\mathcal{T}_\mathcal{X}$, $T_\mathcal{Y}$ and $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$, respectively. Then, using the relation between the representations, we obtain
\begin {equation}
A_k^{\mathcal{Y}}(Cx) = (A_k^{\mathcal{X}\mathcal{Y}} C)x + C (A_k^\mathcal{X} x), \ x\in D(A^\mathcal{X}_k),
\end{equation}
similarly to \eqref{deriv}. It remains to apply Propositions \ref{invarw} and \ref{invarb}.
\end {proof}
\section{Examples}\label{examp}
\subsection{Matrices and resolutions of the identity}\label{matrex}
We begin by revisiting the example we presented in the introduction and proceed to illustrate how general, in fact, it is or can be made.
Let us consider operators in $L(\mathcal{X},\mathcal{Y})$ where $\mathcal{X}$ and $\mathcal{Y}$ are Banach modules with the structures associated with the representations $\mathcal{T}_\mathcal{X}$ and $\mathcal{T}_\mathcal{Y}$, respectively. Assume that the representations $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$ and $\mathcal{T}_{\mathcal{Y}\mathcal{X}}$ are defined as in the example \ref{typop}.
Let $\sigma_k$, $k\in{\mathbb Z}^d$, be a partition of ${\mathbb R}^d$ such that
\begin{itemize}
\item The spectral submodules $\mathcal{X}(\sigma_k,\mathcal{T}_\mathcal{X})$, $\mathcal{Y}(\sigma_k,\mathcal{T}_\mathcal{Y})$, $k\in{\mathbb Z}^d$, are complementable Banach subspaces, i.e.~for each $k\in{\mathbb Z}^d$ there exist two projections $P_k\in B(\mathcal{X})$ and $Q_k\in B(\mathcal{Y})$ onto $\mathcal{X}(\sigma_k, \mathcal{T})$ and $\mathcal{Y}(\sigma_k, \mathcal{T})$ respectively;
\item Projections $P_k$, $Q_k$, $k\in{\mathbb Z}^d$, form disjunctive resolutions of the identity ${\mathcal P}$ and ${\mathcal Q}$ in $\mathcal{X}$ and $\mathcal{Y}$ respectively;
\item ${\mathcal P}$ and ${\mathcal Q}$ satisfy \eqref{crp}.
\end{itemize}
Observe that the above conditions are typically satisfied in Hilbert spaces as well as in different kinds
of $L^p$ spaces.
Let us now define the representations $U_\mathcal{X}$, $U_\mathcal{Y}$ by \eqref{exres} and $U_{\mathcal{X}\mathcal{Y}}$ by \eqref{typop1}. Now we can define various classes $\mathcal{F}(L(\mathcal{X},\mathcal{Y}), U_{\mathcal{X}\mathcal{Y}})$ in terms of the matrix
$(Q_kAP_j)_{k,j\in{\mathbb Z}^d}$ of an operator $A\in L(\mathcal{X},\mathcal{Y})$ as it was done in \cite{B97Izv}. We leave it to
the reader to obtain the specific formulas. We only notice that if
$\sigma_k = k+[0,1)^d$, $k\in{\mathbb Z}^d$, computations similar to those in the proof of Proposition \ref{preq}
show that $\mathcal{F}(L(\mathcal{X},\mathcal{Y}),U_{\mathcal{X}\mathcal{Y}}) = \mathcal{F}(L(\mathcal{X},\mathcal{Y}),\mathcal{T}_{\mathcal{X}\mathcal{Y}})$.
Alternatively, let $\mathfrak B$ be a (complex) unital Banach algebra and $\mathcal{T}: {\mathbb R}^d\to B(\mathfrak B)$ be a group
of bounded (isometric) automorphisms of $\mathfrak B$. Let $\sigma_k$, $k\in{\mathbb Z}^d$, be a partition of ${\mathbb R}^d$ such that
\begin{itemize}
\item The spectral submodules $\mathfrak B(\sigma_k,\mathcal{T})$, $k\in{\mathbb Z}^d$, are complementable Banach subspaces, i.e. for each $k\in{\mathbb Z}^d$ there exists a projection $P_k\in B(\mathfrak B)$ onto $\mathfrak B(\sigma_k, \mathcal{T})$;
\item Projections $P_k$, $k\in{\mathbb Z}^d$, form a disjunctive resolution of the identity ${\mathcal P}$;
\item ${\mathcal P}$ satisfies \eqref{crp}.
\end{itemize}
Given the above conditions we can define a bounded $2\pi$-periodic representation $U_\mathfrak B: {\mathbb R}^d\to B(\mathfrak B)$ similar to \eqref{exres}:
\[U_\mathfrak B(t)x = \sum_{k\in {\mathbb Z}^d} e^{int}P_kx, \ t\in{\mathbb R}^d, x\in\mathfrak B.\]
Various classes $\mathcal{F}(\mathfrak B, U_\mathfrak B)$ are then easily defined in terms of $\|P_kx\|$, $k\in {\mathbb Z}^d$, as e.g.~ in Example \ref{ex1}. Again,
if $\sigma_k = k+[0,1)^d$, $k\in{\mathbb Z}^d$, computations similar to those in the proof of Proposition \ref{preq}
show that $\mathcal{F}(\mathfrak B,U_\mathfrak B) = \mathcal{F}(\mathfrak B,\mathcal{T})$.
Clearly, it is unreasonable to expect the conditions on the spectral submodules outlined in this example to hold outside of the realm of Hilbert and $L^p$ spaces. The following examples
are designed to show that the techniques of this paper still produce meaningful results when previously used methods fail.
\subsection{Continuous operator-valued functions}\label{1examp}
Let $\mathfrak B$ be a (complex) unital Banach algebra. In this example we consider the Banach algebra $C_b = C_b({\mathbb R}^d, \mathfrak B)$ of all bounded
continuous $\mathfrak B$-valued functions with the norm $\|x\| = \sup_{s\in{\mathbb R}^d} \|x(s)\|$, $x\in C_b$.
The algebra $C_b$ is equipped with a non-degenerate $L_1({\mathbb R}^d)$-module structure associated with the translation representation $\mathcal{T} = T: {\mathbb R}^d\to B(C_b)$ similar to \eqref{trans1}:
\[(T(t)x)(s) = x(t+s),\ t,s\in{\mathbb R}^d,\ x\in C_b.\]
In particular, for $f\in L_1({\mathbb R}^d)$ the module structure is given by the convolution
\[(T(f)x)(s) = (f*x)(s) = \int_{{\mathbb R}^d} f(t)x(s-t)dt, \ x\in C_b.\]
The Beurling spectrum $\L(x,T)$ in this case coincides with $\text{ supp\!} \hat x$, where the Fourier transform is understood in the sense of tempered distributions. The class of $T$-continuous functions in this case is
the algebra $C_{ub}=C_{ub}({\mathbb R}^d,\mathfrak B)$ of all bounded uniformly continuous $\mathfrak B$-valued functions.
Consider the functions $\phi_{1,a}^d$ defined by \eqref{triangled}. The following subclasses of $C_{ub}$ (see Remark \ref{wc}) correspond to the classes considered in the previous sections. The Wiener class is
\[\mathcal{W} = \mathcal{W}(C_b)= \mathcal{W}(C_b, T) = \{x\in C_b:\ \int_{{\mathbb R}^d} \|\phi_{1,a}^d*x\| da <\infty\},\]
the class of functions with exponential spectral decay is
\[
\begin {split}
\mathcal{W}_{exp}
= \bigcup_{\a\in{\mathbb R}^d_+}D({\textbf{h}}_{\alpha}) = \ & \{x\in C_b: \|\phi_{1,n}^d*x\| \le M e^{-\varepsilon|n|}, n\in{\mathbb Z},\\ &
\mbox{ for some positive } M=M(x) \mbox{ and } \varepsilon = \varepsilon(x)\},
\end{split}
\]
the classes of functions with one-sided exponential spectral decay are
\[
\begin {split}
\mathcal{W}_{exp}^\pm
= \bigcup_{\a\in{\mathbb R}^d_\pm}D({\textbf{e}}_\alpha ) = \ & \{x\in C_b: \|\phi_{1,n}^d*x\| \le M e^{-\varepsilon|n|}, \pm n \in {\mathbb N},\\ &
\mbox{ for some positive } M=M(x) \mbox{ and } \varepsilon = \varepsilon(x)\},
\end{split}
\]
and the Beurling class is
\[{\mathcal B} = {\mathcal B}(C_b) = \{x\in C_b:\ \sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty} \|\phi^d_{1,n} * x\| < \infty\}.\]
A straightforward computation shows that the Sobolev-type algebras in this case are
\[
\begin {split}
\mathcal{W}^{(m)} & = \mathcal{W}^{(m)}(C_b) = \{x\in C^m_b:\ \sum_{|\a|\le m}\int_{{\mathbb R}^d} \|\phi_{1,a}^d*x^{(\a)}\| da <\infty\} \\
& = \{x\in C_b:\ \int_{{\mathbb R}^d} \|\phi_{1,a}^d*x\|(1+|a|)^m da <\infty\},
\end{split}
\]
and
\[{\mathcal B}^{(m)} = {\mathcal B}^{(m)}(C_b) = \{x\in C^m_b:\ \sum_{|\a|\le m}\sum_{k\in{\mathbb Z}^d}^\infty \max_{|n|_\infty\ge |k|_\infty} \|\phi^d_{1,n} * x^{(\a)}\| < \infty\},\]
where $C_b^m$ is the space of functions with continuous bounded derivatives up to order $m\in{\mathbb N}$ and
$x^{(\a)}$ is the (higher order) partial derivative corresponding to the index $\a\in{\mathbb R}^d$.
In the following theorem we denote by $\mathcal{F}(C_b)$ one of the above classes.
\begin {thm}\label{cb}
Assume that $x \in \mathcal{F}(C_b)$ is such that $x(s)$ is invertible in $\mathfrak B$ for all $s\in {\mathbb R}^d$ and the function $y(\cdot) = (x(\cdot))^{-1}$ is bounded. Then $y\in\mathcal{F}(C_b)$.
\end {thm}
\begin {proof}
The result follows immediately from Theorems \ref{1sidexpwin}, \ref{invexp}, \ref{maint}, \ref{absB}, \ref{sobolt} and Remark \ref{ba}.
\end {proof}
\begin {rem}
In case $\mathfrak B = {\mathbb C}$ and $d=1$, the classes $\mathcal{W}_1$ and $\mathcal{W}_{exp}$ were considered in \cite{B97Sib}. The following two families of classes indexed by $q>1$ were also shown to be inverse closed there:
\[\mathcal{W}_q = \{x\in C_b({\mathbb R}):\ \int_{{\mathbb R}^d} \|\phi_{1,a}^d*x\|(1+|a|)^q da <\infty\} \mbox{ and }\]
\[\mathcal{W}_{q,0} = \{x\in C_b({\mathbb R}):\ \lim_{|a|\to\infty}\|\phi_{1,a}^d*x\| |a|^q =0\}.\]
\end {rem}
\begin {rem}
Observe that if $x\in C_b({\mathbb R})$ is a periodic function, $x(t) \simeq \sum_{n\in{\mathbb Z}} c_ne^{int}$, then $x\in \mathcal{W}$ if and only if it has summable Fourier coefficients. Indeed, in this case, we have
$c_ne^{int} = (\phi_{1,n}*x)(t)$, $n\in{\mathbb Z}$. Hence, the classical Wiener's lemma \cite{W32} follows from Theorem \ref{cb}.
\end {rem}
\begin {rem}
In general, if $x\in C_b({\mathbb R}^d)$, then $|(\phi_{1,a}^d*x)(t)|$ coincides with the absolute value of the Short-time Fourier
transform (STFT) $|(V_{\phi}x)(t,a)|$ of the function $x$ with respect to the window $\phi$ \cite{G01}. Hence, $\mathcal{W}(C_b)$ coincides with the modulation space $M^{\infty,1}({\mathbb R}^d)$ and the above theorem implies that if a function $x\in M^{\infty,1}({\mathbb R}^d)$ is bounded away from $0$ then
$1/x \in M^{\infty,1}({\mathbb R}^d)$ \cite{S95}. In view of the described relation with the STFT, one may call
the map $x \mapsto (\mathcal{V}_\phi x)(a) =\mathcal{T}(\phi_{1,a}x)$ the \emph{abstract STFT} of the element $x$ with respect to a window $\phi$ and a representation $\mathcal{T}$. It is also natural to ask if the methods in this paper can be used to recover the results in \cite{S94, S95, G06}, that is to show that
$M^{\infty,1}({\mathbb R}^d)$, also known as the Sj\"ostrand's class, is an inverse closed subalgebra when the product is defined via twisted convolution as opposed to the ordinary convolution. We believe, that the answer is positive but non-trivial and, hence, deserves a separate paper.
\end {rem}
\begin {rem}
We have presented Theorem \ref{cb} as a special case of Theorems
\ref{1sidexpwin}, \ref{invexp}, \ref{maint}, \ref{absB}, and \ref{sobolt}.
In fact, this case is generic when the algebras are considered.
Indeed, let $\mathfrak B$ be a unital Banach algebra with a strongly continuous group of automorphisms
$\mathcal{T}:{\mathbb R}^d\to B(\mathfrak B)$.
We can regard $\mathfrak B$ as an inverse closed subalgebra of $C_b({\mathbb R}^d,\mathfrak B)$ by identifying $x\in\mathfrak B$ with $x_\mathcal{T} \in C_b$,
$x_\mathcal{T}(t) = \mathcal{T}(t)x$. Moreover, we then have $T(s)x_\mathcal{T}(t) = x_\mathcal{T}(t+s) = \mathcal{T}(t+s)x$ and, hence,
\[(T(f)x_\mathcal{T})(t) = \int_{{\mathbb R}^d} f(s)\mathcal{T}(t-s)xds = \mathcal{T}(t)(fx) = (fx)_\mathcal{T}(t),\ f\in L_1({\mathbb R}^d).\]
Therefore, $\L(x,\mathcal{T}) = \L(x_\mathcal{T},T)$ and $x\in \mathcal{F}(\mathfrak B)$ if and only if $x_\mathcal{T}\in\mathcal{F}(C_b)$, where $\mathcal{F}$ denotes one of the classes in Theorem \ref{cb}.
\end {rem}
\subsection{Integrable operator-valued functions} Keeping some of the notation of the previous example, let $L_1 = L_1({\mathbb R}^d, \mathfrak B)$ be the algebra of Bochner integrable $\mathfrak B$-valued functions with respect to convolution. Adjoining a formal unit $e$ to this algebra we get a unital Banach algebra denoted by $\widetilde{L_1} = \widetilde{L_1}({\mathbb R}^d, \mathfrak B)$. The algebra of Fourier transforms of elements in
$\widetilde{L_1}({\mathbb R}^d, \mathfrak B)$ is then identified with a (not closed) subalgebra $\mathcal{F}(\widetilde{L_1}) \subset C_b$, with
$\hat e \equiv 1_{\mathfrak B}$ and $\hat\Phi$ vanishing at infinity for $\Phi\in L_1$. Moreover,
since $(\widehat{f*\hat\Phi})(s) = 2\pi \hat f(s)\Phi(-s)$, $s\in{\mathbb R}^d$, $f\in L_1({\mathbb R}^d)$, we have
\[\|\hat\Phi\|_{1,1} = 5^d \sum_{n\in{\mathbb Z}^d} \|\phi^d_{1,n} \hat\Phi\|
\le 5^d \int_{{\mathbb R}^d}\int_{[-1,1]^d}\|\Phi(s-a)\|dsda <\infty\]
and, hence, $\mathcal{F}(\widetilde{L_1})\subset \mathcal{W}(C_b)$.
The following result is a version of the celebrated Bochner-Phillips theorem \cite{BP42} for the
algebra $\widetilde{L_1}$.
\begin {thm}
Consider $A = \a e+\Phi \in \widetilde{L_1}$. Then $A$ is invertible in $\widetilde{L_1}$
if and only if $\hat A(\l) = \a1_\mathfrak B + \hat\Phi(\l)$ are invertible in $\mathfrak B$ for all $\l\in{\mathbb R}^d$.
\end {thm}
\begin {proof}
Since $\hat\Phi$ vanishes at infinity, Theorem \ref{cb} applies for $\hat A\in\mathcal{W}(C_b)$ and, therefore, $\hat A^{-1} \in \mathcal{W}(C_b)$. We need to prove that $\hat A^{-1}\in
\mathcal{F}(\widetilde L_1)$.
Consider a b.a.i.~ $(\varphi_n)$, $n\in{\mathbb N}$, in $L_1({\mathbb R}^d)$ such that $\text{ supp\!}\hat\varphi_n$ is compact and $\varphi_n$ are differentiable infinitely many times, $n\in{\mathbb N}$. For example,
let $\varphi\in L_1({\mathbb R}^d)$ be some infinitely many times differentiable function with $\text{ supp\!}\hat\varphi$ compact and
$\hat\varphi(0)=1$, and let $\varphi_n(s) = n^d\varphi(ns)$, $s\in{\mathbb R}^d$, $n\in{\mathbb N}$.
Then $\Phi_n = \varphi_n*\Phi\in L^1({\mathbb R}^d,\mathfrak B)$ is a sequence of entire functions that converge to $\Phi$ in $L_1$ and, moreover, the sequence $\hat A_n (\cdot) = \a 1_\mathfrak B + \hat\Phi_n(\cdot)$,
$n\in{\mathbb N}$, converges to $\hat A$ in $\mathcal{W}(C_b)$.
Hence, there exists $N\in{\mathbb N}$ such that for all $n\ge N$ elements
$\hat A_n$ are invertible in $\mathcal{W}(C_b)$ and $\sup_{n\ge N} \|\hat A_n^{-1}\|_{1,1}< \infty$. Since
$\text{ supp\!}\hat\Phi_n$ is compact and
\[(\hat A_n(\l))^{-1} = \frac1\a 1_\mathfrak B- \frac1\a\hat\Phi_n(\l)(\hat A_n(\l))^{-1}, \l\in{\mathbb R}^d, n\ge N,\]
we may conclude that functions $\hat F_n := \frac1\a\hat\Phi_n\hat A_n^{-1}$, $n\ge N$, have compact support and therefore are Fourier transforms of some functions $F_n\in L_1$, $n\ge N$.
Moreover, by construction, the sequence $(F_n)$, $n\ge N$, is Cauchy and, therefore, converges
to some $F\in L_1$. It remains to observe that $(\hat A(\l))^{-1} = \frac1\a 1_\mathfrak B-\hat F(\l)$, $\l\in{\mathbb R}^d$,
and, hence, $A^{-1} = \frac1\a e - F \in \widetilde{L_1}$.
The converse statement is trivial and the theorem is proved.
\end {proof}
\subsection{Convolution Operators and Multipliers}
The following example is related to the previous two but provides a view from a different vantage point.
Consider an operator valued function $\sigma\in L^\infty({\mathbb R}^d, B({\mathcal H}))$, where ${\mathcal H}$ is a (complex) Hilbert space. Let $A_\sigma$ be an operator defined by
$A_\sigma x= \mathfrak F(\sigma\mathfrak F^{-1}x)$, $x\in L^2({\mathbb R}^d,{\mathcal H}) = \mathcal{X}$, where $\mathfrak F$ is the Fourier transform operator, so that $\|\mathfrak F\|\|\mathfrak F^{-1}\|=1$. Clearly, $A_\sigma\in B(L^2({\mathbb R}^d,{\mathcal H}))$ and $\|A_\sigma\| = \|\sigma\|_\infty$ -- the norm of the function $\sigma$ in $L^\infty$. Observe also that if $x_0 \in C_b({\mathbb R}^d,{\mathcal H})$ is infinitely differentiable and has compact support then $A_\sigma x_0 = \hat\sigma*x_0$ is the convolution with the Fourier transform of $\sigma$ in the sense of tempered
distributions.
Consider the modulation representation $\mathcal{T} = M$: ${\mathbb R}^d\to B(\mathcal{X})$ defined by \eqref{mod1} and the corresponding representation $\mathcal{T}_{\mathcal{X}\X} = \widetilde{M}$: ${\mathbb R}^d\to B(B(\mathcal{X}))$ defined as in Example \ref{typop}, that is
\[\widetilde{M}(t) A= M(t)AM(-t), \ t\in{\mathbb R}^d, A\in B(\mathcal{X}). \]
Applying the above formula to the operator $A_\sigma$ we get $\widetilde{M}(t)A_\sigma x =
\mathfrak F((T(t)\sigma \mathfrak F^{-1} x)$, $x\in\mathcal{X}$, where $T$ is the translation representation defined by \eqref{trans1}. Hence, given $f\in L_1({\mathbb R}^d)$ we get $\widetilde{M}(f)A_\sigma x =
A_{f*\sigma} x$ and, therefore,
\[\|\widetilde{M}(f)A_\sigma\| = \|A_{f*\sigma}\| = \|f*\sigma\|_\infty.\]
The above equalities imply that the Wiener class $\mathcal{W}(\mathcal{X})$ contains the operator $A_\sigma$
if and only if $\sigma\in\mathcal{W}(C_b)$. It also follows that if $\sigma\in\mathcal{W}(C_b)$ is invertible in $L^\infty({\mathbb R}^d, B({\mathcal H}))$ then $A_\sigma$ is invertible and $A^{-1}_\sigma = A_{\sigma^{-1}}\in \mathcal{W}(\mathcal{X})$.
If $\mathcal{X} = L^1({\mathbb R}^d, {\mathcal H})$, then the operator $A_\sigma$ may be defined directly as the convolution
$A_\sigma x = \hat\sigma *x$, $x\in\mathcal{X}$, provided that $\hat\sigma\in L^1({\mathbb R}^d, B({\mathcal H}))$. In this case, $\|A_\sigma\|_{\mathcal{W}} = \|\hat\sigma\|_1$. This shows that if one considers convolution operators on $L^p$, Wiener class
depends non-trivially on $p\in[1,\infty]$. Recall for comparison that in case of matrices the most typical Wiener class consists of matrices with summable diagonals, i.e., it is independent of the choice of $p$ when the matrix acts on $\ell^p$. The more general approach we employed in this paper provides more flexibility and allows us to capture a greater variety of classes of memory decay.
\subsection{Integral Operators.}\label{intops} In this example we consider integral operators
in $B(C_b({\mathbb R}^d, {\mathbb C}^m))=B(\mathcal{X})$. We shall impose module structure on $B(\mathcal{X})$ as in Example \ref{typop}; we will write $\mathcal{T}_{\mathcal{X}\X} = \widetilde M$ if $\mathcal{T}_\mathcal{X} = M$ and
$\mathcal{T}_{\mathcal{X}\X} = \widetilde T$ if $\mathcal{T}_\mathcal{X} = T$.
We begin by describing the properties of the class of operators in question. The following two propositions arise as a special case of \cite[Theorem VI.7.1]{DS88I}.
\begin {prop}\label{repint}
Assume that $A\in B(C_b)$ satisfies
\begin {equation}\label{baicom}
\lim_\a\|M(f_\a)A - AM(f_\a)\| = 0,
\end{equation}
where $(f_a)$ is a b.a.i.~ in $L_1({\mathbb R}^d)$.
Then there exists a family $\mu_t\in Matr_{m}(M_1({\mathbb R}^d))$ of $m\times m$ matrices of finite Borel measures on ${\mathbb R}^d$ such that
\begin {equation}\label{measrep}
(Ax)(t) = \int_{{\mathbb R}^d} \mu_t(ds)x(s), x\in C_b, t\in{\mathbb R}^d. \end{equation}
\end {prop}
\begin {proof}
Condition \eqref{baicom} ensures not only that $AC_0\subseteq C_0$, where $C_0$ is the space of (${\mathbb C}^m$-valued) functions vanishing at infinity, but also that $A$ is completely determined by its
action on $C_0$. Hence, \eqref{measrep} follows from
\cite[Theorem VI.7.1]{DS88I}.
\end {proof}
\begin {prop}\label{repint1}
Assume that $A\in B(C_b)$ satisfies
\eqref{baicom}. Let $\mu_t$, $t\in{\mathbb R}^d$, be the measure matrices in the representation \eqref{measrep} of $A$. Then the following are equivalent:
\begin{description}
\item[(i)] the function $t\mapsto \mu_t: {\mathbb R}^d\to Matr_{m}(M_1({\mathbb R}^d))$ is continuous;
\item[(ii)] $M(\psi)A$ is a compact operator for any $\psi\in L_1({\mathbb R}^d)$;
\item[(iii)] $AM(\psi)$ is a compact operator for any $\psi\in L_1({\mathbb R}^d)$.
\end{description}
\end {prop}
\begin {proof}
(i) $\Leftrightarrow$ (ii) follows immediately from \cite[Theorem VI.7.1]{DS88I} in view of Proposition
\ref{repint}.
(ii) $\Rightarrow$ (iii). If $M(f_\a)A$,
where $(f_\a)$ is a b.a.i.,
are
compact operators then, for any $\psi\in L_1({\mathbb R}^d)$,
\[AM(\psi) = \lim_\a AM(\psi*f_\a)= \lim_\a AM(\psi)M(f_\a) = \lim_\a M(f_\a)AM(\psi)\]
is a uniform limit of compact operators and, therefore, is compact.
(iii) $\Rightarrow$ (ii) follows by a similar argument.
\end {proof}
\begin {prop}
Assume that $A\in B(C_b)$ satisfies
\eqref{baicom}. Then $A$ has a representation
\begin {equation}\label{intop1}
(Ax)(t) = \int_{{\mathbb R}^d} G(t,s)x(s)ds, x\in C_b,\end{equation}
with $G(t,\cdot)\in Matr_{m}(L_1({\mathbb R}^d))$ for all $t\in {\mathbb R}^d$, if and only if
\[\lim_\a AM(h_a) = 0\]
(in the uniform operator topology) for any $\gamma$-net $(h_\a)$, $\gamma\in{\mathbb R}^d$.
\end {prop}
\begin {proof}
Let $\mu_t$, $t\in{\mathbb R}^d$, be the measure matrices in the representation \eqref{measrep} of $A$.
All these measures are absolutely continuous if and only if
for every $\epsilon > 0$ there exists $\d > 0$ such that
$\sup \|Ax\| <\epsilon$ where the supremum is taken over all $x\in C_b$ such that $diam(\text{ supp\!} x) < \d$.
The last condition is equivalent to the one we stated by Definition \ref{gnet} of $\gamma$-nets and the properties of the Beurling spectrum in Lemma \ref{sprop}.
\end {proof}
\begin {defn}
We define a class of operators $\mathfrak I\subset B(C_b)$ by requiring that each $A\in \mathfrak I$ satisfy the following conditions:
\begin{description}
\item[(i)] $A\in (B(\mathcal{X}),\widetilde M)_c$;
\item[(ii)] $AM(h_\a)$ converges to $0$ in the uniform operator topology for any $\gamma$-net $(h_\a)$, $\gamma\in{\mathbb R}^d$;
\item[(iii)] $M(\psi)A$ is a compact operator for each $\psi\in L_1({\mathbb R}^d))$.
\end{description}
\end {defn}
Observe that in view of Proposition \ref{repint1} we can replace (iii) with the equivalent requirement:
\begin{description}
\item[(iii')] $AM(\psi)$ is a compact operator for each $\psi\in L_1({\mathbb R}^d))$.
\end{description}
We note that condition \eqref{baicom} is weaker than $A\in (B(C_b),\widetilde M)_c$ as we have shown in \cite[Section 5]{BK05}.
Hence, combining the propositions we stated in this subsection we get the following characterization of the class $\mathfrak I$.
\begin {prop}
We have $A\in\mathfrak I$ if and only if $A$ has an integral representation \eqref{intop1}, where the kernel $G$ is such that the function $t\mapsto G(t,\cdot)\!: {\mathbb R}^d\to Matr_{m}(L_1({\mathbb R}^d))$ is continuous and bounded. Moreover, for such an $A$ we have
\[\|A\| = \sup_{t\in{\mathbb R}^d} \int_{{\mathbb R}^d} \|G(t,s)\|ds <\infty,\]
if we choose to use $|\cdot|_\infty$ as a norm on ${\mathbb C}^m$.
\end {prop}
Clearly, $\mathfrak I$ is a closed subalgebra of $B(\mathcal{X})$.
Moreover, the following lemma follows immediately from the conditions (i), (ii), (iii), and (iii').
\begin {lem}\label{rep1}
The set $\mathfrak I$ is a left ideal in $(B(\mathcal{X}),\widetilde M)_c$.
\end {lem}
The following key proposition is now immediate.
\begin {prop}\label{rep2}
Assume $A\in\mathfrak I$ and $I+B = (I+A)^{-1} \in B(C_b)$. Then $B\in\mathfrak I$.
\end {prop}
\begin {proof}
Since $I = (I+A)(I+B) = (I+B)(I+A)$, we get $B = - A - AB = -A - BA$ and $AB = BA$. Since $B$ is obviously in $(B(\mathcal{X}),\widetilde M)_c$, Lemma \ref{rep1} applies, and we get $BA\in\mathfrak I$. Finally, since $\mathfrak I$ is a vector space, we get $B\in\mathfrak I$.
\end {proof}
Let us now obtain the definitions of various classes of memory decay of operators in $\mathfrak I$ with respect to the representation $\widetilde M$ in terms of the kernel $G$. After a straightforward computation, we have
\[\mathcal{W} = \mathcal{W}(\mathfrak I, \widetilde M) = \left\{A\in\mathfrak I\!: \int_{{\mathbb R}^d}\sup_{t\in{\mathbb R}^d}\left(
\int_{|t-s-a|\le 1}\left\|G(t,s)\right\|ds\right)da < \infty\right\};\]
\[\begin {split}
\mathcal{W}_{exp} =\! & \bigcup_{\a\in{\mathbb R}^d_+}\! D({\textbf{h}}_{\alpha}) = \Big\{A\in\mathfrak I\!: \sup_{t\in{\mathbb R}^d}
\int_{|t-s-a|\le 1}\left\|G(t,s)\right\|ds
\le M e^{-\varepsilon|a|}, \\ &
\mbox{ for all } a\in {\mathbb R} \mbox{ and some positive } M=M(x) \mbox{ and } \varepsilon = \varepsilon(x)
\Big\};
\end{split}
\]
\[\begin {split}
\mathcal{W}_{exp}^\pm =\! & \bigcup_{\a\in{\mathbb R}^d_\pm}\! D({\textbf{e}}_\alpha ) = \Big\{A\in\mathfrak I\!: \sup_{t\in{\mathbb R}^d}
\int_{|t-s-a|\le 1}\left\|G(t,s)\right\|ds
\le M e^{-\varepsilon|a|}, \\ &
\mbox{ for all } a\in {\mathbb R}_\pm \mbox{ and some positive } M=M(x) \mbox{ and } \varepsilon = \varepsilon(x)
\Big\};
\end{split}
\]
\[{\mathcal B} = {\mathcal B}(\mathfrak I, \widetilde M) =
\left\{A\in\mathfrak I\!: \sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty}\sup_{t\in{\mathbb R}^d}
\int_{|t-s-n|\le 1}\left\|G(t,s)\right\|ds < \infty\right\}.\]
The Sobolev-type classes are, for $m\in{\mathbb N}$,
\[\begin {split}
\mathcal{W}^{(m)} =\ & \mathcal{W}(\mathfrak I, \widetilde M) =
\Big\{A\in\mathfrak I\!: \\
&
\int_{{\mathbb R}^d}\sup_{t\in{\mathbb R}^d}\left(
\int_{|t-s-a|\le 1}\left\|G(t,s)\right\|(1+|t-s|)^mds\right)da < \infty\Big\};
\end{split}
\]
\[\begin {split}
{\mathcal B}^{(m)} =\ & {\mathcal B}^{(m)}(\mathfrak I, \widetilde M) =
\Big\{A\in\mathfrak I\!: \\
&
\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty}
\sup_{t\in{\mathbb R}^d}
\int_{|t-s-n|\le 1}\left\|G(t,s)\right\|(1+|t-s|)^mds < \infty\Big\}.
\end{split}
\]
Next, we obtain the definitions of various classes of memory decay of operators in $\mathfrak I$ with respect to the representation $\widetilde T$. As usually, we employ
functions $\phi_{1,a}^d$ defined by \eqref{triangled}. We have
\[\begin {split}
\mathcal{W} =\ & \mathcal{W}(\mathfrak I, \widetilde T) = \Big\{A\in\mathfrak I\!:\\
& \int_{{\mathbb R}^d}\sup_{t\in{\mathbb R}^d}
\int_{{\mathbb R}^d}\left\|\int_{{\mathbb R}^d} \phi_{1,a}^d(u)G(t-u,s-u)du\right\|dsda < \infty\Big\};
\end{split}\]
\[\begin {split}
\mathcal{W}_{exp} =\ &\mathcal{W}_{exp}(\mathfrak I, \widetilde T)= \bigcup_{\a\in{\mathbb R}^d_+}\! D({\textbf{h}}_{\alpha}) \\
=\ & \Big\{A\in\mathfrak I\!: \sup_{t\in{\mathbb R}^d}
\int_{{\mathbb R}^d}\left\|\int_{{\mathbb R}^d} \phi_{1,a}^d(u)G(t-u,s-u)du\right\| ds
\le M e^{-\varepsilon|a|}, \\ &
\mbox{ for all } a\in {\mathbb R} \mbox{ and some positive } M=M(x) \mbox{ and } \varepsilon = \varepsilon(x)
\Big\};
\end{split}
\]
\[\begin {split}
\mathcal{W}_{exp}^\pm =\ &\mathcal{W}_{exp}^\pm(\mathfrak I, \widetilde T)= \bigcup_{\a\in{\mathbb R}^d_\pm}\! D({\textbf{e}}_\alpha ) \\
=\ & \Big\{A\in\mathfrak I\!: \sup_{t\in{\mathbb R}^d}
\int_{{\mathbb R}^d}\left\|\int_{{\mathbb R}^d} \phi_{1,a}^d(u)G(t-u,s-u)du\right\| ds
\le M e^{-\varepsilon|a|}, \\ &
\mbox{ for all } a\in {\mathbb R}_\pm \mbox{ and some positive } M=M(x) \mbox{ and } \varepsilon = \varepsilon(x)
\Big\};
\end{split}
\]
\[\begin {split}
{\mathcal B} =\ & {\mathcal B}(\mathfrak I, \widetilde T) = \Big\{A\in\mathfrak I\!: \\
&
\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty}\sup_{t\in{\mathbb R}^d}
\int_{{\mathbb R}^d}\left\|\int_{{\mathbb R}^d} \phi_{1,n}^d(u)G(t-u,s-u)du\right\|ds < \infty\Big\};
\end{split}
\]
\[\begin {split}
\mathcal{W}^{(m)} =\ & \mathcal{W}^{(m)}(\mathfrak I, \widetilde T) =
\Big\{A\in\mathfrak I\!: \\
&
\int_{{\mathbb R}^d}\sup_{t\in{\mathbb R}^d}
\int_{{\mathbb R}^d}\left\|\int_{{\mathbb R}^d} \phi_{1,a}^d(u)\frac{d^m}{du^m}G(t-u,s-u)du\right\|dsda < \infty\Big\};
\end{split}
\]
\[\begin {split}
& {\mathcal B}^{(m)} = {\mathcal B}^{(m)}(\mathfrak I, \widetilde T) =
\Big\{A\in\mathfrak I\!: \\
\sum_{k\in{\mathbb Z}^d} \max_{|n|_\infty\ge |k|_\infty}
\sup_{t\in{\mathbb R}^d}
\int_{{\mathbb R}^d}\left\|\int_{{\mathbb R}^d} \phi_{1,n}^d(u)\frac{d^m}{du^m}G(t-u,s-u)du\right\|dsda < \infty\Big\}.
\end{split}
\]
In what follows we denote by $\mathcal{F}(\mathfrak I) = \mathcal{F}(\mathfrak I, \mathcal{T}_{\mathcal{X}\X})$ one of the above classes (with respect to
either $\mathcal{T}_{\mathcal{X}\X} =\widetilde M$ or $\mathcal{T}_{\mathcal{X}\X} =\widetilde T$).
\begin {thm}\label{intop}
Assume that $A \in \mathcal{F}(\mathfrak I)$ is such that $I+A$ is invertible in $B(C_b)$. Then $B = (I+A)^{-1} - I\in \mathcal{F}(\mathfrak I)$.
\end {thm}
\begin {proof}
The result follows immediately from Theorems \ref{1sidexpwin}, \ref{invexp}, \ref{maint}, \ref{absB}, \ref{sobolt}, and Proposition \ref{rep2}.
\end {proof}
\begin {rem}
We can easily define the classes $\mathcal{F}(\mathfrak I, W)$ where $W = \mathcal{T}_{\mathcal{X}\X}$ is the Weyl representation
defined as in \eqref{Weyl}. This way, however, we will not get anything significantly different from
Theorem \ref{intop} because $\mathcal{F}(\mathfrak I, W) = \mathcal{F}(\mathfrak I, \widetilde M) \cap
\mathcal{F}(\mathfrak I, \widetilde T)$.
\end {rem}
\begin {exmp}
An operator $A\in \mathfrak I$ is called $2\pi$-\emph{periodic} if $\widetilde{T}(2\pi)A=A$.
For such operators we have
\[G(t+2\pi, s+2\pi) = G(t,s),\ t,s\in{\mathbb R},\]
and the function $A_{\widetilde T}$ given by $t\mapsto \widetilde{T}(t)A: {\mathbb R}\to \mathfrak I$ is $2\pi$-periodic in continuous in the uniform operator topology. Let us consider the Fourier series
\cite{BK10, B97Sib, B97Izv} of this function:
\[A_{\widetilde T}(\tau)\simeq \sum_{n\in{\mathbb Z}} A_ne^{in\tau},\ \tau\in{\mathbb R},\]
where the Fourier coefficients are given by the standard formula
\[A_n = \frac1{2\pi}\int_0^{2\pi} A_{\widetilde T}(\tau) e^{in\tau}d\tau \in\mathfrak I, \ n\in{\mathbb Z}.\]
Observe that $\widetilde T(t)A_n = e^{int}A_n$, that is the Fourier coefficients are eigenvectors of the representation $\widetilde T$. Hence, $\widetilde T(t)[A_nM(-n)] = A_nM(-n)$, $t\in{\mathbb R}$, $n\in{\mathbb Z}$, and, therefore,
\[(A_nx)(t) = \int_{\mathbb R} G_n(t-s)e^{ins}x(s)ds, t\in{\mathbb R}, n\in{\mathbb Z}.\]
Observe now that for $f\in L_1({\mathbb R})$ we have $\widetilde M(f)A \simeq \sum_{n\in{\mathbb Z}} \hat f(n) A_n$.
Therefore, a $2\pi$-periodic operator $A\in \mathfrak I$ belongs to $\mathcal{W}(\mathfrak I,\widetilde M)$ if and only if it has summable Fourier coefficients, i.e.
\[\sum_{n\in{\mathbb Z}} \|A_n\| = \sum_{n\in{\mathbb Z}} \|G_n\|_{L_1({\mathbb R})} < \infty.\]
\end {exmp}
\begin {rem}
We refer to \cite{FS10, K99, K01, SS09, S08} for related results involving different classes of integral operators.
\end {rem}
\subsection{Integro-differential operators with delay} Here we consider operators $A = \frac d{dt} +
L+C$, where
\[(Lx)(t) = \sum_{k=1}^\infty (L_kT(h_k) x)(t) = \sum_{k=1}^\infty \ell_k(t)x(t+h_k),\]
$ x\in C_b({\mathbb R},{\mathbb C}^m)$, $k\in{\mathbb N}$, $h_k \in {\mathbb R}$, $\ell_k\in C_b({\mathbb R},B({\mathbb C}^m))$, $\sum\limits_{k\in{\mathbb N}} L_kT(h_k)$ converges in the uniform operator topology, and
\[(Cx)(t) = \int_{\mathbb R} G(t,s)x(s)ds,\]
is an integral operator from Subsection \ref{intops}, i.e., $C\in \mathfrak I$. We will consider
$A$ as a bounded operator from the space $\mathcal{X} = C^1_b({\mathbb R}, {\mathbb C}^m)$ of bounded continuously differentiable functions into $\mathcal{Y} = C_b({\mathbb R}, {\mathbb C}^m)$ as well as an unbounded operator
$A: \mathcal{X} =D(A)\subset \mathcal{Y}\to\mathcal{Y}$.
\begin {lem}
Assume that $B = A^{-1} \in B(\mathcal{Y})$. Then $B\in\mathfrak I$.
\end {lem}
\begin {proof}
Let $D=\frac d{dt} + I$: $D(A)\subset\mathcal{Y}\to\mathcal{Y}$. Clearly, $D^{-1} \in B(\mathcal{Y})$ and
\begin {equation}\label{din}
(D^{-1}x)(t) = \int_\infty^t e^{s-t}x(s)ds,
\end{equation}
so that $D^{-1}\in\mathfrak I$. Hence, we have
\begin {equation}\label{nibd}
B = (L+C+D - I)^{-1} = D^{-1}(I +(L+C-I)D^{-1})^{-1} \in \mathfrak I\end{equation}
by Lemma \ref{rep1} and Proposition \ref{rep2} which are both applicable because convergence of $\sum\limits_{k\in{\mathbb N}} L_kT(h_k)$ in the uniform operator topology implies $L\in (B(\mathcal{Y}), \widetilde M)_c$.
\end {proof}
Firstly, we consider the module structure on $L(\mathcal{X},\mathcal{Y})$ associated with the representation $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$ defined as in the example \ref{typop} where $\mathcal{T}_\mathcal{X}$ and $\mathcal{T}_\mathcal{Y}$ are the
translation representations given by \eqref{trans1}. With a slight abuse of notation we will use the symbol $\widetilde T$ to denote either $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$, $\mathcal{T}_{\mathcal{Y}\mathcal{X}}$, $\mathcal{T}_{\mathcal{X}\X}$ or $\mathcal{T}_{\mathcal{Y}\Y}$. It will be clear from the context which of the representations may be used in each instance.
Let $\mathcal{F}$ denote one of the classes of operators with memory decay considered in this paper.
The following
conditions are sufficient to ensure $A\in \mathcal{F} = \mathcal{F}(L(\mathcal{X},\mathcal{Y}),\widetilde T)$:
\begin{description}
\item[(A)] $L_k \in \mathcal{F}(C_b, T)$ for each $k\in{\mathbb N}$, and $\sum\limits_{k\in{\mathbb N}}\|L_k\|_{\mathcal{F}} < \infty$;
\item[(B)] $C\in\mathcal{F}(\mathfrak I, \widetilde T)$.
\end{description}
\begin {thm}
Assume that an operator $A = \frac d{dt} + L+C$ satisfies conditions \emph{(A)} and \emph{(B)} and
$B = A^{-1} \in B(\mathcal{Y})$. Then $B \in \mathcal{F}(\mathfrak I,\widetilde T)$. Moreover, if we regard $B\in L(\mathcal{Y},\mathcal{X})$ then
$B\in \mathcal{F}(L(\mathcal{Y},\mathcal{X}),\widetilde T)$.
\end {thm}
\begin {proof}
Observe that \eqref{din} implies that $D^{-1}\in \mathcal{F}$ for any $\mathcal{F}$. Hence, the
result follows from \eqref{nibd} and Theorem
\ref{intop}.
\end {proof}
Secondly, we consider the module structure on $L(\mathcal{X},\mathcal{Y})$ associated with the representation $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$ defined as in the example \ref{typop} where $\mathcal{T}_\mathcal{X}$ and $\mathcal{T}_\mathcal{Y}$ are the
translation representations given by \eqref{mod1}. We will use the symbol $\widetilde M$ to denote either $\mathcal{T}_{\mathcal{X}\mathcal{Y}}$
or $\mathcal{T}_{\mathcal{Y}\Y}$ with the understanding that the context will determine which of the representations may be used in each instance.
It is not hard to determine when
$L+C \in \mathcal{F} = \mathcal{F}(B(\mathcal{Y}),\widetilde M)$.
In all cases we would need $C\in\mathcal{F}(\mathfrak I, \widetilde M)$. Conditions on the operator $L$ vary:
\[ L\in \mathcal{W}(B(\mathcal{Y}),\widetilde M) \mbox{ if and only if }
\sum\limits_{n\in{\mathbb Z}}\left\|\sum\limits_{k\in{\mathbb Z}} \hat\phi^d_{1,n}(h_k)\ell_k\right\| <\infty;\]
\[L\in \mathcal{W}_{exp}(B(\mathcal{Y}),\widetilde M) \mbox{ if and only if }
\left\|\sum\limits_{k\in{\mathbb Z}} \hat\phi^d_{1,n}(h_k)\ell_k\right\| \le M\gamma^{|n|}, n\in{\mathbb N},\]
for some $M > 0$, and $\gamma\in(0,1)$,
and so on; we leave the remaining classes to the reader.
\begin {thm}
Assume that an operator $A = \frac d{dt} + L+C$ satisfies $L\in \mathcal{F}(B(\mathcal{Y}),\widetilde M)$ and
$C\in\mathcal{F}(\mathfrak I, \widetilde M)$. If $B = A^{-1}\in B(\mathcal{Y})$, then $B \in \mathcal{F}(\mathfrak I,\widetilde M)$.
\end {thm}
\begin {proof}
The
result follows from \eqref{din}, \eqref{nibd} and Theorem
\ref{intop}.
\end {proof}
\begin {rem}
Observe that unlike $\mathcal{T}_\mathcal{Y}$, $\mathcal{T}_\mathcal{X}$ is not a bounded representation in this case (see \eqref{mod1}). Hence, we cannot simply quote the general results of this paper to obtain the memory decay for $B = A^{-1}\in L(\mathcal{Y},\mathcal{X})$. First, one would have to modify the
general definitions of the classes $\mathcal{F}$ considered in this paper.
For an operator given by \eqref{intop1}
description of the classes $\mathcal{F}(L(\mathcal{Y},\mathcal{X}),\widetilde T)$ and $\mathcal{F}(L(\mathcal{Y},\mathcal{X}),\widetilde M)$ is
obtained using the formulas for $\mathcal{F}(\mathfrak I)$ with the kernel $G$ replaced by
$DG = G+\frac{\partial G}{\partial t}$. For more general operators more work would be needed.
Secondly, one would have to prove an analog of Lemma
\ref{algprop} and similar results for other classes of memory decay. We will pursue these and other extensions in a sequel to this paper.
\end {rem}
\medskip
\centerline{\bf Acknowledgements}
\medskip
We are grateful to K. Gr\"ochenig and Q.~Sun for inspiration provided by their papers and useful comments that helped us improve the presentation.
\bibliographystyle{siam}
|
\section{Introduction}
There are currently 538 officially confirmed extrasolar planets known to astronomy\footnote{As of 2011 March 15. A full list of known exoplanets can be found at http://exoplanet.eu}, a substantial fraction of which belong to the `hot Jupiters' category of Jovian-type planets that orbit within $0.1$\,AU of their host star. Improved observational precision has, in recent years, led to the discovery of extreme examples of this class of planet with orbital semi-major axes of the order of $0.02$\,AU or less(see, for example, \citet{sasselov2003}, \citet{hellier2009}, \citet{hebb2010}). The Jovian masses and small orbital separations of these planets imply that they interact tidally with their host stars.
Tidal interactions between stars and planets were first studied in the context of the Solar system \citep{goldreich1966}, and have also been extensively studied as they apply to binary systems \citep{hut1980}, but it is only comparatively recently that such studies have been extended to consider exoplanetary systems. Tidal interactions lead to long-term changes in the orbital parameters of a planet, specifically the eccentricity and the semi-major axis (see e.g. \citet{mardling2002}; \citet{jackson2008}; \citet{barker2009}; \citet{jackson2009}). The end result of this process is the spiral-in of the planet towards its host until it reaches the Roche limit, where it undergoes mass transfer through the L1 point \citep{gu2003}, becoming disrupted in the process. Previous studies of tidal evolution have, for the most part, focused on the effect of tidal interactions on these two orbital parameters whilst neglecting the evolution of other parameters that are involved.
\citet{dobbsdixon2004} demonstrated that the effects of tidal interactions extend to the rotation rates of the two bodies, but until recently little work had been done to investigate this aspect of tidal theory as applied to exoplanetary systems. There is, however, mounting interest in this aspect of exoplanetary tides, with several studies investigating the large scale effect of the presence of exoplanets on stellar rotation \citep{pont2009,alves2010,lanza2010}, and some work investigating tidal interactions specifically \citep{leconte2010}. It is well known that the rotation rate of an isolated star declines with age through the action of magnetic braking \citep{weberdavis1967}, which leads asymptotically to a power-law decay of rotation rate with the inverse square root of stellar age \citep{skumanich1972}. \citet{barnes2007} presented a method for determining the ages of stars using their rotation periods and colours, based on a Skumanich-type magnetic braking law and calibrated using the solar rotation period, as an alternative to stellar model fitting. What appears to be less widely recognized, or at least acknowledged, is that the torques that act on both star and planet as a result of tidal interactions can affect this natural rotational evolution, and may be sufficiently strong to overwhelm it, at least for short periods of time.
In this paper we simulate the evolution of hot Jupiter systems to investigate the effects on the orbital eccentricity, orbital separation, planetary rotation rate and stellar rotation rate of tidal interactions between hot Jupiters and their host stars. In Section\,\ref{sec:maths} we set out the mathematical basis for our models, whilst in Section\,\ref{sec:compute} we describe the computational methods that were used. Section\,\ref{sec:wasp18} and Section\,\ref{sec:wasp19} contain discussion of the application of our simulations to the specific cases of the WASP-18 and WASP-19 systems respectively. It should be noted that these systems were not selected at random, but were chosen owing to pre-existing disagreement between age estimates calculated using different methods. These transiting hot Jupiters also have two of the shortest known orbital periods, and are therefore particularly susceptible to the influence of tidal interactions.
\section{Tidal and wind evolution}
\label{sec:maths}
Following \citet{eggleton1998}, \citet{mardling2002} and \citet{dobbsdixon2004} tidal energy is dissipated within a star and planet whose spin axes are aligned with the orbital axis at rates defined by the tidal quality factors $Q'_s=3Q_s/2k_*$ and $Q'_p=3Q_p/2k_p$, where $k_*$ and $k_p$ are the tidal Love numbers of the two bodies. The eccentricity of the orbit evolves at a rate
\begin{eqnarray}
\frac{\dot{e}}{e}&=&\frac{81}{2}\frac{n}{Q'_p}\frac{M_s}{M_p}\left(\frac{R_p}{a}\right)^5
\left[-f1(e)+\frac{11}{18}\frac{\Omega_p}{n}f_2(e)\right]+\nonumber\\
&&+\frac{81}{2}\frac{n}{Q'_s}\frac{M_p}{M_s}\left(\frac{R_s}{a}\right)^5
\left[-f1(e)+\frac{11}{18}\frac{\Omega_s}{n}f_2(e)\right]
\label{eq:dlnedt}
\end{eqnarray}
where
\begin{equation}
f_1(e)=\left(1+\frac{15}{4}e^2+\frac{15}{8}e^4+\frac{5}{64}e^6\right)/(1-e^2)^{13/2}
\label{eq:f1}
\end{equation}
and
\begin{equation}
f_2(e)=\left(1+\frac{3}{2}e^2+\frac{1}{8}e^4\right)/(1-e^2)^5.
\label{eq:f2}
\end{equation}
The star and planet have masses $M_s$ and $M_p$ and radii $R_s$ and $R_p$ respectively, and rotation rates $\Omega_s$ and $\Omega_p$. The orbital frequency is defined by Kepler's 3rd law, $n^2=G(M_s+M_p)/a^3$.
If the system is fully aligned, then the total angular momentum of the orbit and the axial rotation of the star and the planet, perpendicular to the orbit, is
\begin{equation}
L_{\rm tot} = M_p M_s\sqrt{\frac{Ga(1-e^2)}{M_p+M_s}}+\alpha_sM_sR_s^2\Omega_s+\alpha_pM_pR_p^2\Omega_p.
\label{eq:ltot}
\end{equation}
The stellar and planetary moments of inertia are determined by their effective squared radii of gyration, $\alpha_s$ and $\alpha_p$.
Angular momentum is carried away from the system via a magnetically-channeled, thermally-driven stellar wind, at a rate described by a standard Weber-Davis model
\begin{equation}
\dot{L}_{\rm wind} = -I_s\kappa\Omega_s{\rm Min}(\Omega_s,\tilde{\Omega})^2
\label{eq:ldotwind}
\end{equation}
where the stellar moment of inertia $I_s=\alpha_sM_sR_s^2$. The physical scaling of the braking rate is determined by the constant of proportionality $\kappa$, and $\tilde{\Omega}$ is the `saturation' rotation rate above which the stellar magnetic field strength is assumed to become independent of the stellar rotation rate.
We use the expressions of \citet{dobbsdixon2004} to describe the tidal spin evolution of the planet to the tidal torque:
\begin{equation}
\dot{\Omega}_p =
\frac{9}{2}
\left(\frac{n^2}{\alpha_pQ'_p}\right)
\left(\frac{M_s}{M_p}\right)
\left(\frac{R_p}{a}\right)^3
\left[f_3(e)-f_4(e)\frac{\Omega_p}{n}\right]
\label{eq:dompdt}
\end{equation}
and the star under the influence of both tidal and wind torques:
\begin{eqnarray}
\dot{\Omega}_s &= &
\frac{9}{2}
\left(\frac{n^2}{\alpha_sQ'_s}\right)
\left(\frac{M_p}{M_s}\right)^2
\left(\frac{R_s}{a}\right)^3
\left[f_3(e)-f_4(e)\frac{\Omega_s}{n}\right]
\nonumber\\
&&
-\kappa\Omega_s{\rm Min}(\Omega_s,\tilde{\Omega})^2.
\label{eq:domsdt}
\end{eqnarray}
The polynomials $f_3(e)$ and $f_4(e)$ have the form
\begin{equation}
f_3(e)=\left(1+\frac{15}{2}e^2+\frac{45}{8}e^4+\frac{5}{16}e^6\right)/(1-e^2)^6
\label{eq:f3}
\end{equation}
and
\begin{equation}
f_4(e)=\left(1+3e^2+\frac{3}{8}e^4\right)/(1-e^2)^{9/2}.
\label{eq:f4}
\end{equation}
The total angular momentum of the system evolves as
\begin{eqnarray}
\dot{L}_{\rm wind} &=& M_p M_s\sqrt{\frac{G}{M_p+M_s}}
\left[
\frac{\dot{a}(1-e^2)-2ae\dot{e}}{2\sqrt{a(1-e^2)}}
\right]+\nonumber\\
&&+\alpha_sM_sR_s^2\dot{\Omega}_s+\alpha_pM_pR_p^2\dot{\Omega}_p.
\label{eq:ldottotwind}
\end{eqnarray}
Dividing the angular momentum loss rate by the orbital angular momentum
\begin{equation}
L_{\rm orb}=M_p M_s\sqrt{\frac{Ga(1-e^2)}{M_p+M_s}}
\label{eq{lorb}}
\end{equation}
and rearranging, we obtain the expression for the evolution of the orbital semi-major axis:
\begin{equation}
\frac{\dot{a}}{a} = 2
\left[
\frac{\dot{e}}{e}\frac{e^2}{1-e^2}-\frac{I_s\dot{\Omega}_s}{L_{\rm orb}}-\frac{I_p\dot{\Omega}_p}{L_{\rm orb}}
-\frac{I_s\kappa\Omega_s{\rm Min}(\Omega_s,\tilde{\Omega})^2}{L_{\rm orb}}
\right].
\label{eq:dlnadt}
\end{equation}
By integrating this equation we obtain an estimate of the time remaining to spiral-in
for a planet orbiting a slowly-rotating star,
\begin{equation}
t_{\rm remain}=\frac{2 Q'_s}{117n}\frac{M_s}{M_p}\left(\frac{a}{R_s}\right)^5,
\label{eq:tremain}
\end{equation}
giving a result almost identical to the estimate of the same quantity derived from a slightly different formulation by \citet{levrard2008}.
The quantity $a/R_s$ is derived directly from the transit duration for a transiting planet, enabling $t_{\rm remain}$ to be estimated from directly-observed quantities for a given value of $Q'_s$.
\citet{leconte2010} recently suggested that parametrizing the tidal evolution equations in this manner, with the stellar and planetary tidal quality factors constant in time, is not equivalent to the more traditional approach of considering either a constant phase lag or a constant time lag of the tidal bulge. They further suggested that such equations, if truncated to O($e^2$), produce both qualitatively and quantitatively incorrect evolutionary histories for systems with $e>0.2$, and promote the use of equations derived from \citet{hut1981}. Despite using a constant tidal quality factor model the tidal equations that we use in this work are not truncated in this fashion, and thus do not suffer from the problems that \citeauthor{leconte2010} describe.
\section{Computational method}
\label{sec:compute}
An estimate of the value of the scaling constant for magnetic braking, $\kappa$, is required to implement (\ref{eq:ldotwind}). Following the standard Weber-Davis model, the rate of change of the rotation rate of an isolated star owing to magnetic braking can be described by
\begin{equation}
\dot{\Omega}_s = -\kappa\Omega_s^3.
\label{eq:omegadot}
\end{equation}
Integrating this under the assumption that $\Omega_2<<\Omega_1$ and $t_2>>t_1$ gives
\begin{equation}
\kappa = \frac{\Omega_{2}^{-2}}{2 t_2},
\label{eq:kappa}
\end{equation}which leads to the standard power law for magnetic braking,
\begin{equation}
P \propto t^{\alpha},
\label{eq:Sku}
\end{equation} with $\alpha=0.5$ \citep{skumanich1972}. This neglect of the initial conditions is justified by the observed strong convergence of spin rates to a narrow period-colour relation in the Hyades \citep{radick1987}, and in other clusters of similar ages. We set $\Omega_2=\Omega_{s,Hyades}$, calculated by scaling the $P_{rot}$-(J-K) colour relationship found for the Coma-Berenices cluster by \citet{cameron2009} and using J and K magnitudes taken from the SIMBAD on-line data archive, to obtain an estimate for $\kappa$ at the age of the Hyades. This value is assumed to be constant throughout the evolution of the system. For simplicity, in all cases the planet's rotation is initially assumed to be tidally locked such that $\Omega_{p,0}=n$, but $\Omega_p$ is permitted to evolve independently thereafter.
More recent work has found that the braking law exponent, $\alpha$, diverges slightly from the ideal \citeauthor{skumanich1972} value in some environments (see e.g. \citet{cameron2009}). If we generalise equation\,\ref{eq:omegadot} to raise $\Omega$ to the power $\beta$, then we derive
\begin{equation}
\kappa = \frac{\Omega^{-\frac{1}{\alpha}}}{t/\alpha},
\label{eq:truekappa}
\end{equation}a universal expression for $\kappa$ with $\beta=\frac{1}{\alpha}+1$. We calibrated our implementation of the magnetic braking power law using the stars listed in table\,$4$ of \citet{cameron2009}, finding that an exponent of $\alpha=0.495\pm0.002$ gives good agreement between the measured rotation periods and the evolved stellar rotation rates calculated using our method. We found that there was little to differentiate the evolutionary tracks produced using values of $\alpha$ within this range, and therefore adopted the central value of $\alpha=0.495$.
Starting from a given set of initial conditions $(t_0, e_0, \Omega_{s,0}, \Omega_{p,0}, a_0)$, we integrate the four equations (\ref{eq:dlnedt}), (\ref{eq:dompdt}), (\ref{eq:domsdt}), and (\ref{eq:dlnadt}) using a fourth-order Runge-Kutta scheme, adapted from algorithms in \citet{press1992}. The unit of time is 1\,Gyr, and the integration is allowed to run for the approximate main-sequence lifetime, $t_{\rm MS}\simeq 7(M_s/M_{\odot})^{-3}$\,Gyr for a star of mass $M_s$, or until the planet reaches the Roche limit, as defined by \citet{eggleton1983}. To calculate the stellar radius of gyration the metallicity and mass of the star are used to select an appropriate table from \citet{claret2004,claret2005,claret2006,claret2007}; at each timestep we interpolate through this data using the current time to derive a value for $\alpha_s$. This is not an entirely accurate method, as the series of papers by \citeauthor{claret2004} provides only tables for discrete values of metallicity and stellar mass that in many cases do not coincide with the metallicity of the systems being studied. In these cases we therefore use the table that most closely corresponds to our data.
We use the observed values of $e$, $\Omega_s$ and $a$, and their uncertainties, to evaluate the goodness-of-fit statistic at each timestep according to
\begin{equation}
\chi^2(t) = \frac{(e_{\rm obs}-e(t))^2}{\sigma^2_e}+\frac{(\Omega_{\rm s,obs}-\Omega_{\rm s}(t))^2}{\sigma^2_{\Omega_s}}+\frac{(a_{\rm obs}-a(t))^2}{\sigma^2_a}.
\end{equation}
To this we add a Bayesian prior on the stellar age $t$ to obtain the statistic
\begin{equation}
C=\chi^2(t) +\frac{(t_{\rm obs}-t)^2}{\sigma^2_t},
\label{eq:C}
\end{equation}
where $t_{\rm obs}$ and $\sigma_t$ are the stellar age estimated from the star's density and effective temperature \citep{sozzetti2007}, and its associated uncertainty. The age of the system is taken to be the time of the step at which C is a minimum.
This forward integration method was built into two separate computational schemes: a grid search that carried out the integration for each node in a four-dimensional grid in $a_0$-$e_0$-$\log(Q'_s)$-$\log(Q'_p)$ parameter space, and a Markov-chain Monte-Carlo (MCMC) optimisation scheme to determine the posterior probability distributions of $\log(Q'_s)$, $\log(Q'_p)$, $e_0$ and $a_0$, and to refine the estimates of the initial parameters returned by the grid search.
The best-fitting evolutionary history investigated by the grid search is taken to be the solution that returns the absolute minimum value of the C statistic, whilst under the MCMC scheme the best-fitting set of parameters are taken to be the median values of the respective posterior probability distributions, with $1-\sigma$ errors derived from the values that delineate the central $68.3$\,percent of the distribution. We choose this approach over the absolute minimum, as the latter strongly depends on the precise sampling of the parameter space by any given Markov chain.
It was assumed that the orbit of the planet would monotonically shrink throughout the integration, thus placing a lower limit on the initial semi-major axis of $a_{0,min}=a_{obs}$. The upper limit on the same parameter was set at $a_{0,max}=0.1$\,AU, a value which encompasses $95$\,percent of the current distribution for transiting planets. The ranges of the tidal quality factors were set to $4.0\leq \log(Q'_p)\leq10.0$ and $5.0\leq \log(Q'_s)\leq10.0$ such that the commonly accepted ranges of values ($10^5 - 10^6$ for $Q'_s$, $10^7 - 10^8$ for $Q'_p$ \citep{baraffe2010}) were encompassed together with an additional section of parameter space.
The initial eccentricity distribution for the ensemble of exoplanets is the subject of much discussion. A value of $e_{0,max}=0.2$ encompasses $93$\,percent of the present distribution and limits our parameter space to the region of validity for our tidal equations, as defined by \citet{leconte2010}. However the migration method by which hot Jupiters find themselves at such short orbital periods is still uncertain. One particular mechanism, Kozai scattering \citep{kozai1962}, can pump the eccentricity to large values, and it has been suggested that short period planets are captured at periastron from such highly eccentric orbits. If this is the case then a much higher upper limit on $e_0$ could be justified.
We must thus take into account the initial times at which we will be starting our integrations. \citet{meibom2005} carried out a survey of tidal circularisation in binary stars. They found that open clusters show a characteristic orbital period at which binaries with the most common initial eccentricity circularise (they define a circularised orbit as one with $e=0.01$). This period was found to vary with the age of the cluster. However it is questionable whether exoplanets, owing to the much lower secondary mass, will tend to circularise at the same orbital period. In fact \citet{hansen2010} applied the formalism of \citet{meibom2005} to exoplanetary systems, finding that circularisation periods were generally much shorter than for stellar binaries. We therefore set $e_{0,max}=0.8$, a value encompassing 99\,percent of the current distribution for transiting exoplanets, to allow for the possibility of eccentric orbits produced by Kozai oscillations.
It is important to note that stellar and planetary obliquities greater than $0.0$ are not considered by the integration methods used herein owing to the algebraic formulation that underlies our computational methods. We therefore assume that the orbital and spin angular momenta vectors are aligned, or anti-aligned.
\subsection{Markov-chain Monte-Carlo simulation}
\label{sec:MCMC}
Under our MCMC optimisation scheme, the coordinates of the origin for the Markov chain are set to the centre of the parameter space. An exception is made if this lies within $3-\sigma_C$ of the coordinates of the best-fitting initial conditions, in which case the origin coordinates are set to lie outwith the $3-\sigma_C$ contour in parameter space. The burn-in phase was judged to be complete when the minimum value of the test statistic from the current integration exceeded the median of all previous minimum test statistic values for the first time \citep{knutson2008}.
\section{The WASP-18 system}
\label{sec:wasp18}
The hot Jupiter WASP-18\,b orbits the star $\mbox{HD}10069$, an F6 star with an effective temperature $T_{eff}=6400\pm100$\,K and a V magnitude of $9.3$, situated approximately $100$\,pc from the Earth. A fit to stellar isochrones using the model of \citet{girardi2000} gives a stellar age of $0.630^{+0.950}_{-0.530}$\,Gyr, which can be further constrained using the observed lithium abundance. \citet{hellier2009} (hereafter \defcitealias{hellier2009}{H09}\citetalias{hellier2009}) therefore assign the system an age of $0.5-1.5$\,Gyr, a value that the current work attempts to improve upon. WASP-18\,b was the first planet \emph{confirmed} to have a period of less than a day, orbiting its host in just $0.94$\,days. Recent measurements of the Rossiter-McLaughlin effect by \citet{triaud2010} have shown that the system is well aligned; the assumption of spin-orbit alignment required for our analysis is therefore justified. In the absence of a measured rotation period for WASP-18, we assume that the inclination of the stellar axis is $90^o$ such that $\sin\,I=1.0$ and $v_{rot}=v\sin\,I$.
\begin{table}
\caption{System parameters for WASP-18, taken from \citetalias{hellier2009} with the exception of the rotation period, which is derived from the observed $v\sin\,I$ under the assumption that the inclination of the stellar axis is $90^o$, the J-K colour, which was calculated using data taken from the SIMBAD on-line database, and $\lambda$ \citep{triaud2010}.}
\label{tab:1}
\begin{tabular}{lcc}
\hline \\
Parameter & Value & Units \\
\hline \\
$M_s$ & $1.25\pm0.13$ & $\mbox{M}_{\mbox{sun}}$ \\[2pt]
$R_s$ & $1.216^{+0.067}_{-0.054}$ & $\mbox{R}_{\mbox{sun}}$ \\[2pt]
$J-K$ & $0.278\pm0.032$ & \\[2pt]
$T_{eff}$ & $6400\pm100$ & K \\[2pt]
$v\sin\,I$ & $10.77\pm0.04$ & km\,s$^{-1}$ \\[2pt]
$P_{rot,s}$ & $5.64\pm0.28$ & days \\[2pt]
System age & $0.630^{+0.950}_{-0.530}$ & Gyr \\[2pt]
$M_p$ & $10.30\pm0.69$ & $\mbox{M}_{\mbox{jup}}$ \\[2pt]
$R_p$ & $1.106^{+0.072}_{-0.054}$ & $\mbox{R}_{\mbox{jup}}$ \\[2pt]
$a$ & $0.02026\pm0.00068$ & AU \\[2pt]
$e$ & $0.0092\pm0.0028$ & \\[2pt]
$i$ & $86.0\pm2.5$ & $^o$ \\[2pt]
$P_{orb}$ & $0.94145299\pm0.00000087$ & days \\[2pt]
$\lambda$ & $5.0^{+3.1}_{-2.8}$ & $^o$ \\[2pt]
\hline \\
\end{tabular}
\end{table}
An observed system age of $0.630\pm0.530$\,Gyr was used for the calculation of the C-statistic for this system, in line with the age found from stellar isochrones by \citetalias{hellier2009}. WASP-18\,b has a radius of the order of that of Jupiter, but is significantly more massive. Lacking a means of calculating $\alpha_p$ however, we set $\alpha_p=\alpha_{Jupiter}$ as a reasonable estimate.
In light of the observed system age being consistent with the age of the Hyades, we compared $P_{rot,s}$ from Table\,\ref{tab:1} to $P_{rot,s}(t=t_{Hyades})$, as calculated for our derivation of $\kappa$ for the system, finding the observed stellar rotation period to be shorter than the expected rotation period of $7.00$\,days at the age of the Hyades. We therefore set $t_0=150$\,Gyr$\approx t_{M35}$, calculating $\Omega_{s,0}$ by scaling the period colour relation of \citet{cameron2009} to this age.
\begin{figure*}
\includegraphics[width=\textwidth]{Fig1}
\caption{Projection maps of the test statistic probability density for WASP-18. The projections indicate distinct ranges of $\log(Q'_s)$, $\log(Q'_p)$ and $a_0$ that give high probability, corresponding to a low value of the C-statistic. $e_0$ provides a less clear cut high probability region, with several individual values giving good results. The best-fitting initial parameters were found to be $a_0=0.0427$\,AU, $e_0=0.245$, $\log(Q'_s)=7.00$ and $\log(Q'_p)=7.50$, giving a system age of $0.654$\,Gyr.}
\label{fig:1}
\end{figure*}
The C-statistic values returned by the grid search were converted into probability densities according to
\begin{equation}
P(C) = e^{-\frac{C^2}{2}},
\label{eq:probC}
\end{equation} and plotted as a set of projection maps (Fig.\,\ref{fig:1}). These indicate regions of high probability, and therefore low points in the C-statistic surface, for each pair of initial parameters, and appear to show that there are a range of values for each parameter that could produce plausible evolutionary histories with a good quality of fit. This is particularly noticeable in $e_0$ and $\log(Q'_p)'$, indicating that these parameters have a lesser impact on the orbital and spin evolution of the system than $a_0$ and $\log(Q'_s)$. This is, to a certain extent, unsurprising; the current orbital eccentricity for WASP-18\,b is more uncertain than the orbital separation, and so will have less influence on the value of the C-statistic. It is also interesting to note the generally sharp transitions in Fig.\,\ref{fig:1} between regions with $P(C)>0.4$ and those with $P(C)\approx0$. This indicates that although the values of the initial parameters are somewhat uncertain, they are generally well constrained to a smaller region of parameter space than we have allowed for.
The best-fitting combination of initial parameters in these projection maps was found to be $a_0=0.0428$\,AU, $e_0=0.245$, $\log(Q'_s)=7.00$ and $\log(Q'_p)=7.50$, indicating a system age of $0.654$\,Gyr in good agreement with the existing estimate derived from stellar model fitting. Using the period colour relation of \citet{cameron2009} we calculate an age of $0.387\pm0.024$\,Gyr for WASP-18 using the derived rotation period, inconsistent with both the age from isochrone analysis and the result of our grid search. From the evolutionary track we derive an extremely short remaining lifetime of 0.006\,Gyr, implying that the planet is currently spiralling in towards its host and is on the verge of reaching the Roche limit. This seems incredibly short, and would mean that we have managed to observe WASP-18\,b in a very short window of opportunity. This result is, however, at oodds with the results that we have obtained from our second integration method, the MCMC algorithm.
\begin{figure*}
\includegraphics[width=\textwidth]{Fig2}
\caption{The posterior probability distributions of each pair of jump parameters produced by the MCMC analysis of the WASP-18 system. Note the clear correlations between several of the integration parameter pairs; these arise from the dependence of tidal energy dissipation on the orbital eccentricity, and from the strong coupling that exists between the orbital eccentricity and the orbital separation. The location of the best-fitting parameters in the parameter space explored are denoted by the dashed lines.}
\label{fig:2}
\end{figure*}
The posterior probability distributions produced by the MCMC integration scheme (Fig.\,\ref{fig:2}) indicate significant correlation between many of the pairs of initial orbital parameters, in particular $a_0$ and $\log(Q'_s)$ although it is also noticeable in several of the other distributions. Some correlation is to be expected; the orbital circularisation time-scale depends on energy dissipation within both the star and the planet \citep{miraldaescude2002}, and so a correlation between the eccentricity and the quality factors that govern tidal energy dissipation is unsurprising. The correlation between $a_0$ and $e_0$ may arise as a result of the strong coupling between eccentricity and separation noted by \citet{jackson2008}. The rapidity with which the orbital separation can decrease is intrinsically linked to the efficiency with which energy is dissipated within the system, a process that is governed by the tidal quality factors. Moreover inspection of (\ref{eq:dlnadt}) indicates that $\dot{a}\propto Q^{-1}$, so the form of the correlation between $a_0$ and $\log(Q'_s)$ that is observed is also as expected. In fact it is surprising that little to no correlation is observed between $a_0$ and $\log(Q'_p)$
The best-fitting parameters found using the MCMC code are set out in Table\,\ref{tab:2}, and imply a stellar age of $0.579^{+0.305}_{-250}$\,Gyr, in agreement with the existing age obtained through isochrone fitting, and the age derived from our grid search results. Fig.\,\ref{fig:3} displays the evolution of the stellar rotation for the set of best-fitting parameters; it is important to recall that the age estimate is not evaluated solely on the basis of this parameter, and that the observed orbital eccentricity and semi-major axis play a part as well.
The value of $\log(Q'_s)$ returned by the MCMC-derived integration scheme does not agree with the best-fitting value obtained from the grid search of the defined parameter space. Nor do the values of $a_0$ and $e_0$. We attribute this to the adoption of the set of median parameters from the MCMC scheme to avoid chance encounters with local minima, as well as the discrete nature of the grid search. Comparison of the solution with the lowest value of the C test statistic shows a value of $\log(Q'_s)$ that fits more closely to the grid search result, albeit still in disagreement, although we note that the minimum test statistic value for the grid search is an order of magnitude lower than that for the MCMC exploration of the available parameter space. We also note that although Figs.\,\ref{fig:1} and \ref{fig:2} are broadly similar in form, they differ somewhat in detail. The same correlations between parameters are visible in both figures, but the cutoff in $\log(Q'_s)$ occurs at a slightly greater value in Fig.\ref{fig:2} than Fig.\,\ref{fig:1}. Additionally, the MCMC algorithm explores a more narrow range of both $a_0$ and $e_0$ parameter space than the grid search, but does explore the lower end of $\log(Q'_p)$ space, a region which the grid search suggests gives poor results. We again attribute this to the differences in the manner in which the two algorithms explore the parameter space that we have delineated.
\begin{table}
\caption{The initial orbital parameters and tidal quality factors of the best-fitting tidal evolution histories produced by the grid search and MCMC integration schemes for the WASP-18 system. The $1-\sigma$ error bars for the MCMC derived values are estimated from the parameter values that encompass the central $68.3$\,percent of the final parameter distributions, and in some cases are inflated by the presence of short `tails' in the distributions.}
\label{tab:2}
\begin{tabular}{lcccc}
\hline \\
Parameter & Grid search & MCMC & Units \\
\hline \\
$a_0$ & $0.0427$ & $0.0258^{+0.0052}_{-0.0044}$ & AU \\[2pt]
$e_0$ & $0.245$ & $0.0399^{+0.1023}_{-0.0351}$ & \\[2pt]
log($Q'_s$) & $7.00$ & $8.21^{+0.90}_{-0.52}$ & \\[2pt]
log($Q'_p$) & $7.50$ & $7.77^{+1.54}_{-1.25}$ & \\[2pt]
age & $0.654$ & $0.579^{+0.305}_{-0.250}$ & Gyr \\[2pt]
$\mbox{t}_{\mbox{remain}}$ & $0.006$ & $0.076^{+0.790}_{-0.044}$ & Gyr \\[2pt]
\hline \\
\end{tabular}
\end{table}
Using (\ref{eq:tremain}) with the data from Table\,\ref{tab:2} and the best-fitting value of $\log(Q'_s)$ from the MCMC scheme, we estimate that WASP-18\,b will reach its Roche limit $0.086$\,Gyr from now. The evolutionary track produced using the MCMC results implies a remaining lifetime of $0.076^{+0.742}_{-0.044}$\,Gyr for the planet, consistent with this value. These are, respectively, $2.40$ and $2.12^{+20.73}_{-1.23}$\,percent of the expected main sequence lifetime of the host star, which we estimate to be $3.58$\,Gyr. These remaining lifetimes strongly imply that the WASP-18 planetary system has an short life expectancy, as found by \citetalias{hellier2009}. Although still quite short, these values are far more reasonable than the value of 6\,Myr derived from the grid search.
The evolution of the stellar rotation period (Fig.\,\ref{fig:3}) for the set of initial parameters returned by the MCMC integration scheme indicates that the rotational evolution of the star under the influence of tides initially differs little from our simulation in which the rotation period is governed purely by a Skumanich-type magnetic braking law. However the evolution of the rotation period rapidly diverges from this ideal case, with the rate of spin-down slowing gradually as the star ages and the planet migrates towards its host under the influence of tidal interactions. The final spiral-in of the planetary companion causes a rapid and substantial spin-up of the star; during this process the stellar rotation period is reduced by almost a factor of two from its maximum $6.26$\,day period to $3.16$\,days at the Roche limit, suggesting significant spin up of the host star. Furthermore, Fig.\,\ref{fig:3} clearly shows that, for the age estimate returned by our MCMC integration scheme, the observed rotation period of the host star is quite inconsistent with the rotation period of $7.96$\,days that would be expected at the same age if tidal interactions played no part in the evolution of stellar rotation.
\begin{figure}
\includegraphics[width=0.48\textwidth]{Fig3}
\caption{The evolution of the rotation period of WASP-18 from the best-fitting initial conditions found by the MCMC exploration of the allowed parameter space (solid line), along with the evolution that would occur in the absence of tidal interactions (dashed line). For reference, the observed parameters are plotted with associated errors, and the best-fitting age given by the MCMC integration scheme is denoted (dotted line). The tide-governed evolution diverges from the ideal rotational evolution early on, with the effect of tidal interactions on the the rotation of the host star increasing with time until the rotation period dramatically decreases during the final spiral-in of the planetary body.}
\label{fig:3}
\end{figure}
Although the best-fitting parameters that we have adopted show WASP-18\,b in the process of its final spiral in, the true picture is slightly more ambiguous, as evidenced by the large upper uncertainty on the remaining lifetime. Within the $1-\sigma$ parameter ranges there exist combinations of the initial parameters that produce evolutionary tracks of a different form to that displayed in Fig.\,\ref{fig:3}. These evolutionary tracks follow the stellar rotation curve produced by a purely magnetic braking scenario for a much longer period of time, and exhibit little spin up of the host star. Furthermore, in semi-major axis space they indicate that the planet stays at approximately the same semi-major axis for the duration of the star's main sequence lifetime. In these cases the age of the system lies towards the lower end of the adopted range, owing to the necessity of conforming to the calculated stellar rotation period, and the planet has a long remaining lifetime. We note however that these solutions form only a minority of the evolutionary tracks consistent with the parameter ranges that we adopt from the MCMC results. More prevalent were tracks of the same form as that displayed in Fig.\,\ref{fig:3}, but with the spin-up taking place much later such that the stellar rotation implies an age consistent with the gyrochronological estimate and magnetic-braking only scenario.
We attempted to further constrain the range of possible tidal quality factors by visually fitting to the observed parameters and their associated $1-\sigma$ uncertainties. Starting from the best-fitting initial orbital separation and eccentricity, we investigated the evolution of the system across the range of values for both $\log(Q'_s)$ and $\log(Q'_p)$. We found that changing the value of $\log(Q'_p)$ makes no difference to the evolution of the orbital separation or rotation period, but has a strong effect on the eccentricity evolution of the planetary orbit. $\log(Q'_s)$, in contrast, strongly affects the evolution of all integration parameters. Moreover we found that changing the initial semi-major axis, within the adopted range, made a significant difference to the evolution of the stellar rotation period, whilst modifying $e_0$ merely affected the evolution of the orbital eccentricity. We therefore conclude that the two most important parameters with respect to the stellar rotation are $Q'_s$ and $a_0$.
We were unable to constrain the permissible ranges of the two tidal quality factors any further, owing to the range of values that $a_0$ was able take, but found that for the majority of the possible combinations the rotational evolution of WASP-18\,A gradually diverges from that expected of a purely magnetic braking scenario, and that the star is eventually spun-up by a substantial amount during the final in-spiral of its planetary companion. Fig.\,\ref{fig:4a} shows the rotational period evolution resulting from several values of $\log(Q'_s)$ within the permissible range, with $\log(Q'_p)$, $a_0$ and $e_0$ set to values adopted from the MCMC search.
\begin{figure}
\subfloat{
\includegraphics[width=0.48\textwidth]{Fig4a}
\label{fig:4a}}\\
\subfloat{
\includegraphics[width=0.48\textwidth]{Fig4b}
\label{fig:4b}}
\caption{Evolutionary tracks for WASP-18 produced using several different values of $\log(Q'_s)$ within the range consistent with the observed system parameters. The tracks were calculated using the values of $e_0$, $a_0$ and $\log(Q'_p)$ adopted from the MCMC solution and, from left to right (at the end of the track): $\log(Q'_s)=8.02$, $\log(Q'_s)=8.21$, $\log(Q'_s)=8.70$ and $\log(Q'_s)=9.11$. \emph{Upper panel:} The evolution of stellar rotation period with time for WASP-18. The dashed line shows the evolution expected from a Skumanich-type magnetic braking law. Changing the value of $Q'_p$ has no effect on the evolution of the rotation period with time. \emph{Lower panel:} The evolution of the C test statistic with time for the tracks in the upper panel. The time at which C is a minimum is taken to be the age of the system given the set of initial parameters used for that evolutionary track.}
\label{fig:QrangeW18}
\end{figure}
\section{The WASP-19 system}
\label{sec:wasp19}
The transiting hot Jupiter WASP-19\,b orbits a late G-dwarf star with an orbital period of $0.7888399\pm0.0000008$\,days. The host star has been measured to have $T_{eff}=5500\pm100$\,K, periodic sinusoidal flux variations that indicate a detectable level of intrinsic variability and activity, and a metallicity slightly greater than solar at $\mbox{[M/H]}=0.1\pm0.1$\,dex \citep{hebb2010} (hereafter \defcitealias{hebb2010}{H10}\citetalias{hebb2010}). Measurement of the Rossiter-McLaughlin angle for the system indicates that it is aligned \citep{hellier2011}.
The age of the system is still somewhat uncertain, with \citetalias{hebb2010} able to determine a constraint of $>1.0$\,Gyr. We note that several studies (e.g. \citet{hansen2010,weidner2010}), as well as the exoplanet encyclopedia, cite an age of $0.6\pm0.5$\,Gyr for the star and attribute it to \citetalias{hebb2010}. We believe that this originates from the abstract of \citetalias{hebb2010} which quotes that age as one of two possibilities. However the text of that paper favours the older age constraint. For our analysis we used an age of $t_{age}=5.5\pm4.5$\,Gyr, consistent with the isochronal fit and lower bound on the age quoted by \citetalias{hebb2010}, and set $\alpha_p=\alpha_{Jupiter}=0.26401$. Comparing the stellar rotation period from Table\,\ref{tab:3} to the rotation period expected at the age of the Hyades, we found that $P_{rot,s}>P_{rot,Hyades}=8.60$\,days and thus set $t_0=t_{Hyades}$ and $\Omega_{s,0}=\Omega_{s,Hyades}$.
\begin{table}
\caption{System parameters and $1-\sigma$ limits for WASP-19, taken from the free eccentricity fit of \citetalias{hebb2010} with the exceptions of the J-K colour, which is derived from data taken from SIMBAD, and the rotation period (Collier Cameron, \emph{priv. comm.}).}
\label{tab:3}
\begin{tabular}{lcc}
\hline \\
Parameter & Value & Units \\
\hline\\
$M_s$ & $0.95\pm0.10$ & $\mbox{M}_{\mbox{sun}}$\\[2pt]
$R_s$ & $0.93^{+0.05}_{-0.04}$ & $\mbox{R}_{\mbox{sun}}$\\[2pt]
$J-K$ & $0.43$ & \\[2pt]
$T_{eff}$ & $5500\pm100$ & K \\[2pt]
$v\sin\,I$ & $4.0\pm0.2$ & km\,s$^{-1}$ \\[2pt]
$P_{rot}$ & $10.5\pm0.2$ & days \\[2pt]
System age & $5.5^{+9.0}_{-4.5}$ & Gyr \\[2pt]
$M_p$ & $1.14\pm0.07$ & $\mbox{M}_{\mbox{jup}}$\\[2pt]
$R_p$ & $1.28\pm0.07$ & $\mbox{R}_{\mbox{jup}}$\\[2pt]
$a$ & $0.0164^{+0.0005}_{-0.0006}$ & AU\\[2pt]
$e$ & $0.02^{+0.02}_{-0.01}$ & \\[2pt]
$i$ & $80.8\pm0.8$ & \\[2pt]
$P_{orb}$ & $0.7888399\pm0.0000008$ & days \\[2pt]
\hline\\
\end{tabular}
\end{table}
Fig.\,\ref{fig:5} shows the C-statistic probability density projection maps produced from the grid search results for the WASP-19 system. They indicate a fairly small region of high probability for $\log(Q'_s)$ and $\log(Q'_p)$, but appear to show that there are broader ranges of both $e_0$ and $a_0$ that give strong solutions. The absolute maximum probability, and thus the minimum value of the C statistic, was found to occur when the initial conditions were $a_0=0.0939$\,AU, $e_0=0.735$, $\log(Q'_s)=6.25$ and $\log(Q'_p)=10.0$, giving a system age of $2.11$\,Gyr that is consistent with the existing constraint of $>1.0$\,Gyr from \citetalias{hebb2010}. Using the period-colour relation of \citet{cameron2009} we calculate a gyrochronological system age of $0.899\pm0.037$\,Gyr, immediately suggesting that gyrochronological analysis is inappropriate for this system.
\begin{figure*}
\includegraphics[width=\textwidth]{Fig5}
\caption{C-statistic probability density projection maps for WASP-19\,b. A small region of high probability is clearly visible in $\log(Q'_s)$ and $\log(Q'_p)$ parameter space, but in $a_0$ and $e_0$ the range of values that give high probability solutions are much broader. The point of maximum probability, indicating the minimum in the C-statistic surface, was found to occur at parameter space coordinates of $a_0=0.0939$\,AU, $e_0=0.735$, $\log(Q'_s)=6.25$ and $\log(Q'_p)=10.0$, giving a system age of $2.11$\,Gyr.}
\label{fig:5}
\end{figure*}
Comparing the sets of projection maps for the two systems, it is apparent that there are similarities in the forms of Figs.\,\ref{fig:1} and \ref{fig:5}, particularly in the $a_0-\log(Q'_s)$ and $e_0-a_0$ maps. However the range of probability values covered by the greyscale is slightly narrower for WASP-19, where the maximum probability is approximately $0.8$, compared to WASP-18, where the maximum probability is almost $1.0$. From this we deduce that the best-fitting orbital solution for WASP-19 is less certain than our best-fitting grid search solution for WASP-18. This lower maximum probability, coupled with the form of the projection maps, also suggests that the range of tidal quality factors for which valid orbital solutions exist is greater for this system than it was for WASP-18.
\begin{figure*}
\includegraphics[width=\textwidth]{Fig6}
\caption{The posterior probability distributions of each pair of integration parameters produced by the MCMC analysis of the WASP-19 system. The correlations between the parameters are substantially different than those for the WASP-18 system, with that between $\log(Q'_s)$ and $a_0$ being the only particularly noticeable correlation. The location of the best-fitting parameters in the parameter space explored are denoted by the dashed lines.}
\label{fig:6}
\end{figure*}
The posterior probability distributions produced by the MCMC algorithm exhibit a strong correlation between $a_0$ and $\log(Q'_s)$, of a similar form to that observed for the WASP-18 system albeit extended to lower values of $\log(Q'_s)$. Also present is a correlation between $e_0$ and $a_0$ that is somewhat similar to that apparent in Fig.\,\ref{fig:2}, but the other correlations that were present in the distributions for WASP-18 are absent for WASP-19. The region of parameter space explored by the MCMC algorithm is also substantially greater for the WASP-19 system than it was for the WASP-18 system, although this is not unexpected after the comparative appearance of Figs.\,\ref{fig:1} and \ref{fig:5}.
It is also interesting to note the relative lack of agreement between the results from the grid search and the MCMC search methods for this system; the only search parameter for which the two sets of results agree is $\log(Q'_s)$. This is also apparent from inspection of Figs\,\ref{fig:5} and \ref{fig:6}, as the region of greatest probability in the former does not match the location of the adopted parameters in the latter. As for the WASP-18 system, we attribute this to the adoption of the set of median parameters from the MCMC scheme; comparison of the solution with the lowest value of the C test statistic shows that it more closely fits with the location of the high probability region in Fig.\,\ref{fig:5}. Fortunately the stellar tidal quality factor, the most important parameter when considering the evolution of the stellar rotation, is relatively well constrained in our MCMC-derived solution.
The probability distribution for the initial orbital eccentricity of the WASP-19 system also shows weaker clustering than for the WASP-18 system, and the clustering that is present encompasses a much smaller range of eccentricity values above $e_0=0$. This may be symptomatic of the fact that the eccentricity of the system is not well known; previous analysis of the parameters of the system found little difference in the fit to the observed transit lightcurve between the cases in which the orbit was forced to be circular, and in which the eccentricity was allowed to float \citepalias{hebb2010}. In the latter case, the best-fitting eccentricity value was only $0.02^{+0.02}_{-0.01}$, and it is this value that we used in our $C$ statistic calculations.
\begin{table}
\caption{The initial orbital parameters and tidal quality factors of the best-fitting tidal evolution histories produced by the grid search and MCMC integration schemes for the WASP-19 system. The $1-\sigma$ error bars for the MCMC derived values are estimated from the parameter values that encompass the central $68.3$\,percent of the final parameter distributions.}
\label{tab:4}
\begin{tabular}{lcccc}
\hline \\
Parameter & Grid search & MCMC & Updated & Units \\
\hline \\
$a_0$ & $0.0939$ & $0.0317^{+0.0228}_{-0.0146}$ & $0.0317^{+0.0228}_{-0.0089}$ & AU \\[2pt]
$e_0$ & $0.735$ & $0.0017^{+0.0597}_{-0.0016}$ & & \\[2pt]
log($Q'_s$) & $6.25$ & $6.47^{+2.19}_{-0.95}$ & $6.47^{+0.67}_{-0.95}$ & \\[2pt]
log($Q'_p$) & $10.0$ & $6.75^{+1.86}_{-1.77}$ & & \\[2pt]
age & $2.11$ & $1.60^{+2.84}_{-0.79}$ & & Gyr \\[2pt]
$\mbox{t}_{\mbox{remain}}$ & $0.0038$ & $0.0067^{+1.1073}_{-0.0061}$ & & Gyr \\[2pt]
\hline \\
\end{tabular}
\end{table}
The best-fitting parameters given by the MCMC integration scheme, set out in Table\,\ref{tab:4}, imply a stellar age of $1.60^{+2.84}_{-0.79}$\,Gyr, in broad agreement with the loose constraint of $\mbox{age}>1.0$\,Gyr found by \citetalias{hebb2010}. The gyrochronological age that we have calculated for the system also agrees with this age estimate but lies at the lower end of the range, suggesting that gyrochronology provides a possible, if unlikely estimate for the age of WASP-19\,A. The data in Table\,\ref{tab:4} further imply a remaining lifetime for the planet of $0.0067^{+1.1073}_{-0.0061}$\,Gyr, a mere $0.4$\,percent of the estimated system age. This strongly suggests that WASP-19\,b is in the final, spiral-in stage of its orbital evolution, a conclusion supported by the stellar rotation period evolutionary track that results from integration of the best-fitting parameters (Fig.\,\ref{fig:7}). As with the WASP-18 system we attempted to further constrain the range of possible tidal quality factors by visually fitting to the observed parameters; in this case we were able to reduce the upper limit on $\log(Q'_s)$ and raise the lower limit on $a_0$. The updated values are given in Table\,\ref{tab:4}. Updating the limits on these parameters have no effect on the $1-\sigma$ limits of the age or remaining lifetime.
\begin{figure}
\includegraphics[width=0.5\textwidth]{Fig7}
\caption{The evolution of the rotation period of WASP-19 from the best-fitting initial conditions found by the MCMC exploration of the allowed parameter space, along with the evolution that would occur in the absence of tidal interactions (dashed line). The tide-governed evolution follows the gradual spin-down of the isolated evolution closely until spiral-in of the planet begins, at which point the stellar rotation period rapidly decreases. The observed rotation period is plotted with associated errors, and implies that WASP-19\,b is in the final spiral-in stage of its orbital evolution. Furthermore it is irreconcilable with the expected spin down from magnetic braking alone, and can only be explained by invoking spin-up during tidal interactions.}
\label{fig:7}
\end{figure}
The observed spin rate of WASP-19\,b is inconsistent with evolution governed only by magnetic braking; Fig.\,\ref{fig:7} shows that, at the system age implied by our MCMC solution, the observed period of $10.5\pm0.2$\,days is substantially less than the period of $15.9$\,days implied by our magnetic braking only model. It is possible that WASP-19\,b is in the process of spiraling-in to the Roche limit, spinning up its host as it does so.
Fig.\,\ref{fig:8} shows the rotational period evolution resulting from $\log(Q'_s)$ values within the $1-\sigma$ limits returned by the MCMC search scheme. These reinforce the idea that WASP-19\,b is on the verge of reaching its Roche limit; for the best-fitting tidal quality factors the end of the evolutionary track is at $P_{rot}=10.5$\,days compared to the observed $P_{rot}=10.5\pm0.2$\,day, placing the planet precisely on the Roche limit and suggesting that we have been fortunate to observe WASP-19\,b at all.
\begin{figure}
\includegraphics[width=0.5\textwidth]{Fig8}
\caption{The evolution of stellar rotation period with time for WASP-19 for a range of $\log(Q'_s)$ values. The dashed line shows the evolution expected from a Skumanich-type magnetic braking law, whilst the three solid lines are the evolution that results from the best-fitting $e_0$, $a_0$ and $\log(Q'_p)$ given by the MCMC algorithm, and, from left to right (at the observed system age), $\log(Q'_s)=6.14$, $\log(Q'_s)=6.47$ and $\log(Q'_s)=6.52$. These tracks imply that WASP-19\,b is very close to the Roche limit, and will reach it in less than 10\,Myr.}
\label{fig:8}
\end{figure}
Much as with the WASP-18 system though, this is not the entire picture presented by our results. The upper limit on $t_{remain}$ implied by our adopted solution is $1.114$ \,Gyr, a significant 69.6\,percent of the estimated age and 13.7 percent of the MS lifetime of the host star. This strongly suggests that there exist solutions that more closely follow the stellar rotation evolution expected when only magnetic braking is acting on the star. To investigate this we characterised the orbital evolutionary tracks for each $\log(Q'_s)-a_0$ node from the grid search that was consistent with our results from the MCMC algorithm. Since our investigation into WASP-18\,b's future evolution led us to the conclusion that $e_0$ and $\log(Q'_p)$ have little effect on the evolution of the stellar rotation, we set these parameters to the MCMC values from table\,\ref{tab:4}.
We found that for each value of $a_0$ there was only a very narrow range of $\log(Q'_s)$ values that gave solutions with an age $\geq1.0$\,Gyr \emph{and} a rotation period at that age consistent with the observed value. In fact only $13$\,percent of the nodes investigated gave such solutions. The majority of the nodes gave solutions that either did not reach an age of $1.0$\,Gyr, instead showing the distinctive spiral-in signature at younger ages, or that followed the magnetic braking-only track for much longer such that the spin-up induced by the final spiral-in was insufficient to reduce the stellar rotation period back to the observed value. Tracks that fell into this latter category often returned ages consistent with our gyrochronological calculation, although some did return older ages consistent with the $1.0$\,Gyr lower limit owing to the influence of the orbital eccentricity and semi-major axis.
This raises questions concerning WASP-19\,A's true age. There is evidence from isochrone fitting that the star is older than 1\,Gyr. This is supported by the investigation of \citetalias{hebb2010} into the space velocity of the star, which found that 65\,percent of simulated stars with similar stellar properties in a small volume around WASP-19\,A were older than 1.0\,Gyr. But the rotation period of WASP-19\,A implies a younger age of $0.899$\,Gyr through gyrochronology. The lithium abundance of $\log(A[Li])<1.0$ quoted by \citetalias{hebb2010}, while consistent with either estimate as it only constrains the age to greater than $0.6$\,Gyr, would appear to support a younger age, as the presence of any lithium tends to rule out a more evolved star. There therefore appear to be two main possibilities; either the star is old and has been spun up by the infall of the planet, or the star is younger, still following its natural spin down and more dense and/or hotter than expected for its age. Both of these are consistent with the results that we have obtained from our search methods, so which is the more plausible?
\subsection{Is W19 young or old?}
\label{sec:W19dicuss}
We consider our results for WASP-19 in the context of a larger ensemble of exoplanetary systems with similar stellar and planetary properties, assuming that the value of $\log(Q'_s)=6.47$ adopted for WASP-19 is applicable to the other systems under consideration. We find that for 7 of the 9 planetary systems in the sample the remaining lifetime is $<1$\,Gyr. Furthermore we find that WASP-4\,b has a remaining lifetime of $0.017$\,Gyr, and calculate that it lies at a distance of only 1.4 times the Roche limit from its host star; this places it in apparently similar circumstances to WASP-19\,b. WASP-2\,b and WASP-10\,b have remaining lifetimes at least one order of magnitude greater than those of the rest of the sample, but they also have significantly longer orbital periods.
\begin{table*}
\begin{threeparttable}
\caption{A comparison of the WASP-19 system to a sample of transiting hot Jupiter systems with $M_p>M_J$ in close orbits around stars similar to, or cooler than, WASP-19\,A. We calculate the remaining lifetime for each planet using equation\,\ref{eq:tremain}, assuming that the stellar tidal quality factor of $\log(Q'_s)=6.47$ adopted for WASP-19\,A is applicable across the entire sample. Note that the remaining lifetime for WASP-19\,b quoted here does not agree with the value derived from the evolutionary track displayed in Fig.\,\ref{fig:7}.}
\label{tab:5}
\begin{tabular}{lllllllll}
\hline \\
System & $\mbox{M}_p$ /$\mbox{M}_J$ & $\mbox{R}_p$ /$\mbox{R}_J$ & $\mbox{M}_s$ /$\mbox{M}_\odot$ & $\mbox{R}_s$ /$\mbox{R}_\odot$ & $a$ /AU & $P$ /days & $\mbox{t}_{remain}$ /Gyr & Age /Gyr\\
\hline \\
WASP-19 & $1.14$ & $1.28$ & $0.95$ & $0.93$ & $0.0164$ & $0.7888399$ & $0.012$ & $1.60^{+2.84}_{-0.79}$ \\[2pt]
WASP-2 & $0.847$ & $1.079$ & $0.84$ & $0.834$ & $0.03138$ & $2.1522254$ & $1.708$ & $11.9^{+8.1}_{-4.3}$ \tnote{a} \\[2pt]
WASP-4 & $1.12$ & $1.416$ & $0.93$ & $1.365$ & $0.023$ & $1.3382282$ & $0.017$ & $7.0^{+5.2}_{-4.5}$ \tnote{a} \\[2pt]
WASP-5 & $1.63$ & $1.171$ & $1.00$ & $1.15$ & $0.02729$ & $1.6284246$ & $0.080$ & $3.0\pm1.4$ \tnote{b} \\[2pt]
WASP-10 & $3.06$ & $1.08$ & $0.71$ & $0.783$ & $0.0371$ & $3.0927616$ & $1.818$ & $0.8\pm0.2$ \tnote{c} \\[2pt]
CoRoT-1 & $1.03$ & $1.49$ & $0.95$ & $1.11$ & $0.0254$ & $1.5089557$ & $0.093$ & \\[2pt]
CoRoT-2 & $3.31$ & $1.465$ & $0.97$ & $0.902$ & $0.0281$ & $1.7429964$ & $0.159$ & \\[2pt]
OGLE-TR-113 & $1.24$ & $1.11$ & $0.78$ & $0.77$ & $0.0229$ & $1.4324772 $ & $0.224$ & $>0.7$ \tnote{d}\\[2pt]
TrES-3 & $1.91$ & $1.305$ & $0.92$ & $0.813$ & $0.0226$ & $1.30618608$ & $0.111$ & $0.1^{+0.7}_{-0.0}$ \tnote{a} \\[2pt]
\hline \\
\end{tabular}
\begin{tablenotes}
\item(a)\citet{southworth2010} \item(b)\citet{anderson2008} \item(c)\citet{christian2009} \item(d)\citet{melo2006}
\end{tablenotes}
\end{threeparttable}
\end{table*}
It is interesting to consider the range of the stellar tidal quality factor that produces certain threshold values of $\mbox{t}_{remain}$. We find that, if we exclude the longer period WASP-2 and WASP-10 systems, $\log(Q'_s)>7.10$ leads to an estimated $\mbox{t}_{remain}\geq1.0$\,Gyr for all planets in the sample, whilst a remaining lifetime greater than $0.1$\,Gyr requires $\log(Q'_s)>6.12$. For WASPs -2 and -10 we find that the values are $6.24$ and $5.24$, and $6.22$ and $5.22$ respectively.
The tidal evolution solution that we presented previously would seem to support the hypothesis that the star is old, with stellar spin up accounting for the rotation period. But the results in Table\,\ref{tab:5} would seem to add weight to the idea that the system is actually younger than expected. The likelihood of observing one system in such a condition is low, but to observe two such systems as Table\,\ref{tab:5} implies that we have done with WASPs -4 and -19, seems incredulous.
Considering the remaining lifetimes in the context of the total planetary lifetime suggests a different scenario. \ref{eq:tremain} can also be used to calculate the total planetary lifetime if the initial semi-major axis is known. We calculated the total planetary lifetimes assuming that the value adopted for WASP-19\,b, $a_0=0.0317$\,AU applied unilaterally across our sample, and in Fig.\,\ref{fig:9} plot remaining lifetime as a function of total lifetime. It is immediately apparent that WASP-19\,b is a special case, being clearly separated from the rest of the sample. The remaining lifetime of WASP-19\,b is only between 1 and 2\,percent of the total lifetime, whilst for all of the other planets in the sample this figure is greater than 10\,percent. The short remaining lifetime of WASP-4\,b is thus somewhat misleading, as it represents a significant portion ($\approx12.6$\,percent) of the total lifetime of the planet.
\begin{figure}
\includegraphics[width=0.5\textwidth]{Fig9}
\caption{Remaining lifetime as a function of total planetary lifetime for the sample of transiting planets in Table\,\ref{tab:5}, as calculated using \ref{eq:tremain} assuming that the values of $\log(Q'_s)$ and $a_0$ adopted for the WASP-19 system from our MCMC results apply universally. We plot the results for $\log(Q'_s)=5.52$\,(squares), $\log(Q'_s)=6.47$\,(circles) and $\log(Q'_s)=7.14$\,(triangles). The data for WASP-19\,b are denoted by filled symbols. The lines represent $t_{remain}=1$\,percent (red, dash), $5$\,percent (green, dot-dash) and $10$\,percent (dark blue, dot) of the total planetary lifetime. WASP-19\,b is clearly separate from the rest of the sample, indicating that it is in a unique situation.}
\label{fig:9}
\end{figure}
There are several ways in which we can analyse the remaining lifetimes calculated here, one of which is to examine the probability that we have managed to observe the system in its present configuration. We calculated this for the systems in Table\,\ref{tab:5}, disregarding those with no literature age estimate. For WASP-19 we calculate the probability using both our own age adopted age estimate, and the younger age of $0.6\pm0.5$ that can be found in the literature. Fig.\,\ref{fig:10} displays the results for a range of $\log(Q'_s)$ values consistent with the $1-\sigma$ limits derived from the MCMC posterior probability distributions. For the value of $\log(Q'_s)$ adopted from the MCMC solution we find that two systems, including WASP-19 at our older age estimate, have a probability of observation of less than 1\,percent. Using the younger age estimate for WASP-19 pushes the probability up to 2\,percent, which is more plausible (see analysis in \citet{hellier2011}). Increasing the value of $\log(Q'_s)$ to our upper limit of 7.14 increases the probability of observing WASP-19\,b to approximately 3.5\,percent for the older age, and approximately 9\,percent for the younger age. However if we consider the lower limit of $\log(Q'_s)=5.52$ then four systems, including both ages of WASP-19, have observation probabilities less than 1\,percent. This is an entirely implausible situation.
However this analysis assumes that all of the planets in our sample started at the same distance from their respective host stars, and that they experience tidal interactions of the same strength. It seems somewhat unlikely that this accurately represents reality, as the planetary systems in our sample show significant variation in their properties. Previous studies \citep{matsumura2010,hansen2010} have found that different planetary systems are likely to experience different strengths of stellar and planetary tide, so describing these systems with a single value is almost certainly unphysical. This would, of course, mean that the observational probabilities for several of these hot Jupiters could be significantly greater.
\begin{figure}
\includegraphics[width=0.5\textwidth]{Fig10}
\caption{The probability of observing the systems in Table\,\ref{tab:5} in their present configuration, as a function of $\log(Q'_s)$ within the range encompassed by the $1-\sigma$ errors from the MCMC results. Systems without a stellar age in the literature were disregarded. The data are offset slightly from the true value of $\log(Q'_s)$ to aid clarity for the error bars. Systems for which $P>1$ are plotted at $P=1$. WASP-19 is denoted by filled symbols; triangles for the probability using the age estimate from the MCMC results, squares for the probability calculated using the estimate of $0.6\pm0.5$\,Gyr prevalent in the literature. At $\log(Q'_s)=6.47$ two systems, including WASP-19 with its older age, show a probability of observation that is less than 1\,percent. Increasing the stellar tidal quality factor increases the probability that we have observed the systems in the state implied by their remaining lifetimes.}
\label{fig:10}
\end{figure}
To further muddy the waters we add evidence for an old age by carrying out additional stellar model fits. We used the updated Padova models \citep{marigo2008}, which imply an age of $7.94^{+4.65}_{-2.93}$\,Gyr, and the latest version of the Yonsei-Yale isochrones \citep{demarque2004}, which return an age of $7^{+2}_{-3}$\,Gyr. Both estimates agree with the earlier isochrone fit of \citetalias{hebb2010}. We also investigated the values of $T_{eff}$ and $\rho_s$ required for a stellar age of $0.6\pm0.5$\,Gyr in these two models, assuming the same metallicity; the results are given in Table\,\ref{tab:6}. The correction required to the effective temperature if the stellar density is correct is not improbable, but the converse is not necessarily true. Modifying the stellar density requires a change to either the stellar mass or radius, or both, either of which could have a dramatic effect on our estimates of the planetary parameters. This is notwithstanding the more likely case, which is that both the temperature and density are in need of a slight adjustment. A younger age for the star does not seem, therefore, outwith the realms of possibility.
For completeness we also attempted to force an age of 0.6\,Gyr by adjusting the stellar metallicity whilst leaving the effective temperature and stellar density at the values from \citetalias{hebb2010}. We found that the upper limit of $Z=0.03$ imposed on the Padova isochrones prevented us from reaching such a low age, leaving us with a lowest age estimate of $3.98\pm3.96$\,Gyr, which could be consistent with a star the age of the Hyades. When fitting to the Yonsei-Yale isochrones we found that a value of $[Fe/H]=0.32$ gave an age of $2.5^{+3.5}_{-1.9}$\,Gyr, just consistent with the younger age that we were looking for. Fitting to the 0.6\,Gyr isochrone itself would require a much greater metallicity, but given that this value of the iron abundance is already 0.05\,dex greater than the maximum elemental abundance found by \citetalias{hebb2010} (specifically the upper $1-\sigma$ limit on the Calcium abundance), it is clear that such a fit would be utterly inconsistent with the current spectral analysis.
It therefore seems that a combination of slightly increased elemental iron abundance, greater stellar density and higher effective temperature could produce stellar model fits more consistent with the young stellar age suggested by gyrochronology. Indeed if one is using the Padova models then such a combination lies just within the existing $1-\sigma$ ranges for the three parameters.
\begin{table}
\caption{The age estimates obtained through fitting the stellar parameters of WASP-19\,A to stellar models. Adjusting the stellar density, effective temperature or iron abundance in isolation lowers the age that the model fit returns. Obtaining an age of $0.6$\,Gyr required the parameters to be increased beyond their $1-\sigma$ limits when done in isolation. If adjusted as a set, the required age could be obtained with more plausible values.}
\label{tab:6}
\begin{tabular}{lllll}
\hline \\
Model & $\rho_s /\rho_\odot$ & $T_{eff}$ /K & [Fe/H] & age /Gyr \\
\hline \\
Padova & $1.13\pm0.12$ & $5500\pm100$ & $0.02$ & $7.94^{+4.65}_{-2.93}$ \\[2pt]
Padova & $1.45\pm0.10$ & $5500\pm100$ & $0.02$ & $0.631^{+3.35}_{-0.627}$ \\[2pt]
Padova & $1.13\pm0.12$ & $5805\pm100$ & $0.02$ & $0.631^{+3.35}_{-0.626}$ \\[2pt]
Padova & $1.13\pm0.12$ & $5500\pm100$ & $0.2$ & $3.98\pm3.96$ \\[2pt]
Padova & $1.25\pm0.10$ & $5585\pm100$ & $0.12$ & $0.631^{+3.35}_{-0.625}$ \\[2pt]
Yonsei-Yale & $1.13\pm0.12$ & $5500\pm100$ & $0.02$ & $7^{+2}_{-3}$ \\[2pt]
Yonsei-Yale & $1.52\pm0.10$ & $5500\pm100$ & $0.02$ & $0.60^{+1.90}_{-0.54}$ \\[2pt]
Yonsei-Yale & $1.13\pm0.12$ & $5885\pm100$ & $0.02$ & $0.60^{+1.90}_{-0.54}$ \\[2pt]
Yonsei-Yale & $1.13\pm0.12$ & $5500\pm100$ & $0.32$ & $2.5^{+3.5}_{-1.9}$ \\[2pt]
Yonsei-Yale & $1.30\pm0.10$ & $5625\pm100$ & $0.12$ & $0.60^{+1.90}_{-0.40}$ \\[2pt]
\hline \\
\end{tabular}
\end{table}
As noted previously, the constraint placed on the age by the lithium abundance has the potential to rule out the young stellar age. The detection of a substantial lithium abundance in a stellar spectrum strongly implies a young age for the star, even one that hosts planets, as recent work by \citet{baumann2010} shows that there is no correlation between the presence of planets and reduced stellar lithium content. The original analysis in \citetalias{hebb2010} utilised only 34 spectra from the CORALIE spectrograph \citep{queloz2000}, but thanks to an ongoing radial velocity observation program we now have access to additional data on WASP-19\,A. Since the publication of \citetalias{hebb2010} a further 3 spectra have been taken using CORALIE, and 36 spectral measurements have been obtained using the HARPS high precision echelle spectrograph \citep{mayor2003}. We co-added the individual spectra into a single spectrum with a higher signal-to-noise ratio, from which we were able to improve the constraint on the lithium abundance. Hints of a lithium line were present but only at the level of the spectral noise, limiting us to an upper limit of $\log(\epsilon_{Li})<0.5$. This allows us to place a lower limit on the age of WASP-19\,A of $2.0$\,Gyr \citep{sestito2005}, making an older age for the star more palatable.
The final piece of evidence to be considered is the proximity of WASP-19,\b to the Roche limit. Using the observed system parameters and their associated uncertainties with the formulation of \citet{eggleton1983}, we calculate that the Roche limit for WASP-19\,b is at $0.0148\pm0.0025$\,AU, in agreement with the observed orbital separation of $0.0164^{+0.005}_{-0.0006}$\,AU. \citet{hellier2011} place the planet slightly further from the star at 1.21 times the Roche tidal radius, but we note that they have used a different formulation for calculating the limit.
We conclude that the evidence is in favour of WASP-19\,A being old. Whilst there is evidence for the star being young, it is more circumstantial than that which points to an older star. The upper limit on the lithium abundance and the results from stellar model fitting in particular point towards an age in excess of $1.0$\,Gyr. If WASP-19\,A is indeed old, then the exploration of tide-governed evolution presented herein suggests that WASP-19\,b has spun up its host star, and might be in the final stages of spiralling into the Roche limit. During our exploration of possible evolutionary histories we found that those which returned a stellar age of $>1.0$\,Gyr tended to exhibit a very short remaining lifetime. There were, however, some histories that married an older age for the star to a long remaining lifetime for the planet, so we are unable to completely rule out that scenario.
\section{Summary and conclusions}
\label{sec:conclude}
We have calculated an age for the WASP-18 system of $0.579^{+0.305}_{-0.250}$\,Gyr, in agreement with the $0.630^{+0.950}_{-0.530}$\,Gyr age found from stellar isochrones by \citetalias{hellier2009}. Using an MCMC algorithm we find tidal quality factors of $\log(Q'_s)=8.21^{+0.90}_{-0.52}$ and $\log(Q'_p)=7.77^{+1.54}_{-1.25}$. Our results imply that WASP-18\,b will reach the Roche limit in $0.076^{+0.790}_{-0.044}$\,Gyr, and that in most cases it will cause its host to spin up as it does so. We are unable to place stronger constraints the status of the system with respect to planetary infall owing to the range of evolutionary histories that we find to be compatible with its observed parameters, but a large number of the evolutionary tracks investigated imply that the planet is in the process of spiralling in to its host star. Our results for the WASP-18 system seem roughly consistent with the work of \citet{matsumura2010}.
For the WASP-19 system we found tidal quality factors of $\log(Q'_s)=6.47^{+0.67}_{-0.95}$ and $\log(Q'_p)=6.75^{+1.86}_{-1.77}$. These values give a stellar age of $1.60^{+2.84}_{-0.79}$\,Gyr, in broad agreement with the constraint of $\mbox{age}\geq1.0$\,Gyr found by \citetalias{hebb2010}, and imply a remaining lifetime of $0.0067^{+1.1073}_{-0.0061}$\,Gyr. We have investigated the possibility that WASP-19\,A is younger than previously estimated, in line with the predictions of gyrochronology. After considering the evidence for both the old and young stellar age possibilities, we conclude that the older age is more probable based on a reanalysis of the spectral lithium abundance and updated stellar model fits. We therefore suggest that WASP-19\,b could be in the final stages of its spiral-in, and could be on the verge of reaching the Roche limit. We found that this scenario was more prevalent amongst those evolutionary histories with a stellar age $>1.0$\,Gyr, but that there were some instances in which the older stellar age coincided with a substantial remaining planetary lifetime. We are therefore unable to rule out the possibility that it will be some time before the planet falls into the star.
Tidal interactions between these two hot Jupiters and their host stars will dramatically affect the evolution of the stellar rotation periods, counteracting and then reversing the spin-down that is expected from evolution according to a Skumanich-type magnetic braking law. The observed rotation periods are irreconcilable with such an evolution of the rotation period, and strongly suggest that falling in hot Jupiters cause their host stars to spin up during their inward, tidal interaction governed migration. It is worth noting that the our results seem to point toward a more diverse range of stellar tidal dissipation strength than is commonly considered in the literature. The ranges of $\log(Q'_s)$ that we have attributed to the two systems investigated herein are slightly disparate, for which there may be several possible explanations. We turn to the work of \citet{pinsonneault2001}, which shows that the mass of the convective zone is a function of $T_{eff}$, and therefore of spectral type. From their fig.\,1, we note that the effective temperature of WASP-18\,A lies close to the point at which the mass of the convective zone becomes negligible, implying a convective zone mass of $M_{CZ}\approx0.001$\,$\mbox{M}_{\mbox{sun}}$. WASP-19\,A, with its much lower effective temperature, would have $M_{CZ}\approx0.030$\,$\mbox{M}_{\mbox{sun}}$. In our work we have assumed that the star rotates as a single body, but we have made no assumption about where the majority of tidal dissipation takes place. If this process occurs in the convective zone then the discrepancy in the masses, and therefore depths, of the convective zones of the two stars could provide an explanation for the disagreement as to the value of $Q'_s$, with a larger convective zone allowing for more efficient dissipation and hence a smaller quality factor.
This study also provides a warning against using gyrochronology to estimate the ages of hot Jupiter host stars. Owing to the tidal spin-up of its host star by the in-falling planet, the age that we have found for the WASP-19 planetary system is greater than the age found using gyrochronology alone. The situation for the WASP-18 system is less clear cut, but there is no doubt that gyrochronology will not provide an accurate estimate of the system age at all points during its evolution. We therefore suggest that care should be taken when applying gyrochronology to hot Jupiter systems. The two systems studied herein are extreme examples of this class of planet, with extremely short orbital periods and very close orbits. It is likely that there will be a critical point in semi-major axis space, beyond which tidal interactions will be sufficiently weak so as not to influence the stellar rotation period. For systems in this region gyrochronology may well work very well. But for systems such as WASP-18 and WASP-19 which have $a<a_{crit}$, gyrochronology should be applied with care. Determination of this critical point will require further work, and it is likely to depend on a host of factors.
\citet{lanza2010} also suggests that gyrochronology may not always provide accurate age estimates for exoplanetary systems, finding that plotting $P_{rot}t^{-\alpha}$ as a function of $T_{eff}$ for planet hosting stars gives a poor fit to the period-colour relation of \citet{barnes2007}. This implies that exoplanet host stars are systematically faster rotators than stars with similar ages and properties that do not appear to have any associated planets. \citeauthor{lanza2010} also finds that the rotation period evolution of F- and G-type planet hosts does not differ substantially from similar stars without hot Jupiters, a conclusion with which we disagree although we note that the final spiral-in of the planet does not appear to have been considered.
The substantially reduced rotation period that results from tidal spin-up may provide a means of detecting stars that have either been planet hosts in the past, or that have unseen planetary companions that are in the process of spiraling in to the Roche limit. Measurement of such an anomalous rotation period would provide a strong indication of the current or previous existence of a hot Jupiter around that star. Searching for such systems could help to pinpoint targets for extra-solar planet searches.
\section*{Acknowledgments}
\label{sec:acknowledge}
DJAB would like to thank Keith Horne and Moira Jardine for useful comments and suggestions made during an early draft of this manuscript. The authors are grateful to the referee Brad Hansen, for insightful and thought-provoking comments that greatly improved the quality of the manuscript. The WASP Consortium consists of representatives from the Universities of Keele, Leicester, The Open University, Queens University Belfast and St Andrews, along with the Isaac Newton Group (La Palma) and the Instituto de Astrofisca de Canarias (Tenerife). The SuperWASP and WASP-S Cameras were constructed and operated with funds made available from Consortium Universities and PPARC/STFC. This research has made use of NASA's Astrophysics Data System Bibliographic Services, the ArXiV preprint service hosted by Cornell University, and the SIMBAD database, operated at CDS, Strasbourg, France.
\bibliographystyle{mn2e}
|
\section{Introduction}
Throughout by a \emph{foliation} we mean a singular foliation
(Sussmann \cite{sus}, Stefan \cite{st0}), and by a \emph{regular
foliation} we mean a foliation whose leaves have the same
dimension. Introducing the notion of foliations, Sussmann and
Stefan emphasized that they play a role of collections of
"accessible" sets. Alternatively, they regarded
foliations as integrable smooth distributions. Another point of
view is to treat foliations as by-products of non-transitive
geometric structures, c.f. \cite{dlm}, \cite{va} and examples in
\cite{ry2}. In Molino's approach some types of singular foliations
constitute collections of closures of leaves of certain regular
foliations (\cite{mol}, \cite{wol}). In this note we regard
foliations as a special type of diffeomorphism groups.
Given a $C^{\infty}$ smooth paracompact boundaryless
manifold $M$, $\diff^r(M)_0$ (resp. $\diff^r_c(M)_0$), where
$1\leq r\leq\infty$, is the subgroup of the group of all $C^r$
diffeomorphisms $\diff^r(M)$ on $M$ consisting of diffeomorphisms
that can be joined to the identity through a $C^r$ isotopy (resp.
compactly supported $C^r$ isotopy) on $M$. A diffeomorphism group
$G\leq \diff^r(M)$, is called \emph{isotopically connected}
if any
element $f$ of $G$ can be joined to $\id_M$ through a $C^r$
isotopy in $G$. That is, there is a mapping $\R\t M\ni(t,x)\mapsto
f_t(x)\in M$ of class $C^r$ with $f_t\in G$ for all $t$ and such
that $f_0=\id$ and $f_1=f$. It is well-known that any
isotopically connected group $G\leq \diff^r(M)_0$ defines uniquely
a foliation of class $C^r$, designated by $\F_G$ (see sect. 2).
Our first aim is to establish a correspondence between the class
$\frak F^r(M)$ of all $C^r$-foliations on $M$ and a subclass of
diffeomorphism groups on $M$, and, by using it, to interpret some
results and some open problems concerning non-transitive
diffeomorphism groups. The second aim is to prove new results
(Theorems 1.1 and 1.2) illustrating this correspondence.
A group $G\leq\diff^r(M)$ is called \emph{factorizable} if for
every open cover $\U$ and every $g\in G$ there are $g_1\ld g_r\in
G$ with $g=g_1\ldots g_r$ and such that $g_i\in G_{U_i}$, $i=1\ld
r$, for some $U_1\ld U_r\in\U$. Here for $U\s M$ and
$G\leq\diff^r(M)$, $G_U$ stands for the identity component of the
group of all diffeomorphisms from $G$ compactly supported in $U$.
Next $G$ is said to be \wyr{non-fixing} if $G(x)\neq \{ x \}$ for
every $x \in X$.
\begin{thm}
Assume that $G\leq
\diff^r_c(M)_0$, $ 1\leq r\leq\infty$, is isotopically connected,
non-fixing and factorizable group of diffeomorphisms of smooth
manifold $M$. Then the commutator group $[G,G]$ is simple if and
only if the corresponding foliation $\F_{[G,G]}$ admits no proper
(i.e. not equal to $M$) minimal set.
\end{thm}
In early 1970's Thurston and Mather proved that the group
$\diff^r_c(M)_0$, where $1\leq r\leq\infty$, $r\neq \dim(M)+1$, is
perfect and simple (see \cite{Thu74},\cite{Mat}, \cite{Ban97}).
Next, similar results were proved for classical diffeomorphism
groups of class $C^{\infty}$ (\cite{Ban97},
\cite{ry5}).
For the significance of these
simplicity theorems, see, e.g., \cite{Ban97}, \cite{ry5} and
references therein.
Let $(M_i,\mathcal F_i)$, $i=1,2$, be foliated manifolds. A map
$f:M_1\rightarrow M_2$ is called \emph{foliation preserving}
if~$f(L_x)=L_{f(x)}$ for any $x\in M_1$, where $L_x$ is the leaf
meeting $x$. Next, if~$(M_1,\mathcal F_1)=(M_2,\mathcal F_2)$ then
$f$ is \emph{leaf preserving} if~$f(L_x)=L_x$ for all $x\in M_1$.
Throughout $\diff^r(M,\mathcal F)$ will stand for the group
of~all leaf preserving $C^r$-diffeomorphisms of~a~foliated
manifold $(M,\mathcal F)$. Define $\diff^r(M,\mathcal F)_0$ and
$\diff^r_c(M,\mathcal F)_0$ analogously as above. Observe that a
perfectness theorem for the compactly supported identity component
$\diff^{\infty}_c(M,\mathcal F)_0$, being a non-transitive
counterpart of Thurston's theorem,
has been proved by the author in
\cite{ry1} and by Tsuboi in \cite{Tsu1}. Next, the author in \cite{ry2},
following Mather \cite{Mat}, II,
showed that $\diff^r_c(M,\mathcal F)_0$ is perfect provided
$1\leq r\leq\dim\F$. Observe that, in general, the group $\diff^r_c(M,\mathcal F)_0$
is not simple for obvious reasons.
\begin{thm}
Let $(M,\F)$ be a foliation on a $C^{\infty}$ smooth manifold $M$
with no leaves of dimension $0$. Then the commutator subgroup
$$\D=[\diff^r_c(M,\F)_0,\diff^r_c(M,\F)_0]$$ is simple if and only
if $\F_{\D}$ does not have any proper minimal set. In particular,
if $\F$ is regular, and $1\leq r\leq \dim\F$ or $r=\infty$, then
$\diff^r_c(M,\F)_0$ is simple if and only if $\F$ has no proper
minimal sets.
\end{thm}
In the proof of Theorem 1.1 in sect. 3 some ideas from Ling
\cite{li} are in use.
\section{Foliations correspond to a subclass of the class of diffeomorphism groups}
Let $1\leq r\leq\infty$ and let $L$ be a~subset of
a~$C^r$-manifold $M$ endowed with a~$C^r$-differen\-tiable
structure which makes it an~immersed submanifold. Then $L$ is
\emph{weakly imbedded} if for any locally connected topological
space $N$ and a continuous map $f:N\rightarrow M$ satisfying
$f(N)\subset L$, the map $f:N\rightarrow L$ is continuous as well.
It~follows that in this case such a~differentiable structure is
unique . A \emph{foliation of class} $C^r$ is a partition
$\mathcal{F}$ of $M$ into weakly imbedded submanifolds, called
leaves, such that the following condition holds. If~$x$~belongs to
a $k$-dimensional leaf, then there is a local chart $(U,\phi)$
with $\phi(x)=0$, and $\phi(U)=V\times W$, where $V$ is open in
$\mathbb{R}^k$, and $W$ is open in $\mathbb{R}^{n-k}$, such that
if~$L\in \mathcal{F}$ then $\phi(L\cap U)=V\times l$, where
$l=\{w\in W| \phi^{-1}(0,w)\in L\}$. A foliation is called
\emph{regular} if all leaves have the same dimension.
Sussmann (\cite{sus}) and Stefan (\cite{st0},\cite{st2}) regarded foliations
as collections of accessible sets in the following sense.
\begin{dff}
A smooth mapping $\phi$ of a open subset of $\mathbb{R}\times M$
into $M$ is said to be a \emph{$C^r$-arrow}, $1\leq r\leq\infty$, if\\
(1) $\phi(t,\cdot)=\phi_{t}$ is a local $C^r$-diffeomorphism for
each $t$, possibly with empty domain,\\
(2) $\phi_0=\id$ on its domain,\\
(3) $\dom(\phi_t)\subset \dom(\phi_s)$ whenever $0\leq s<t$.
\end{dff}
Given an arbitrary set of arrows $\mathcal A$, let $\mathcal A^*$
be the totality of local diffeomorphisms $\psi$ such that $\psi =
\phi(t,\cdot)$ for some $\phi\in \mathcal A$, $t\in \mathbb{R}$.
Next $\hat{\mathcal A}$ denotes the set consisting of all local
diffeomorphisms being finite compositions of elements from
$\mathcal A^*$ or $(\mathcal A^*)^{-1}=\{\psi^{-1}|\psi \in
\mathcal A^*\}$, and of the identity. Then the orbits of
$\hat{\mathcal A}$ are called \emph{accessible} sets of $\mathcal
A$.
For $x\in M$ let $\mathcal A(x)$, $\bar{\mathcal A}(x)$ be the
vector subspaces of $T_x M$ generated by
$$\{\dot{\phi}(t,y)|\phi\in \mathcal A,\phi_t(y)=x\},
\quad \{d_y\psi(v)|\psi\in \hat{\mathcal A},\psi(y)=x,v\in
\mathcal A(y)\},$$ respectively. Then we have (\cite{st0})
\begin{thm}
Let $\mathcal A$ be an arbitrary set of $C^r$-arrows on M. Then
\begin{enumerate} \item every accessible set of $\mathcal A$ admits a (unique)
$C^r$-differentiable structure of~a~connected weakly imbedded
submanifold of~$M$; \item the collection of accessible sets
defines a foliation $\mathcal{F}$; and \item
$\D(\F):=\{\bar{\mathcal A}(x)\}$ is the tangent distribution of
$\mathcal{F}$.\end{enumerate}
\end{thm}
Let $G\leq \diff^{r}(M)$ be an isotopically connected group of
diffeomorphisms. Let $\A_G$ be the totality of restrictions of
isotopies $\R\t M\ni(t,x)\mapsto f_t(x)\in M$ in $G$ to open
subsets of $\R\t M$. Then by $\F_G$ we denote the foliation
defined by the set of arrows $\A_G$. Observe that $\hat\A_G=\A_G$
and, consequently, $\bar\A_G(x)=\A_G(x)$.
\begin{rem}
(1) Of course, any subgroup $G\leq \diff^{r}(M)$ determines
uniquely a foliation. Namely, $G$ defines uniquely its maximal
subgroup $G_0$ which is isotopically connected.
(2) Denote by $G_c$ the subgroup of all compactly supported
elements of $G$. Then $G_c$ need not be isotopically connected
even if $G$ is so. In fact, let $G=\diff^r(\R^n)_0$, $1\leq
r\leq\infty$. Then every $f\in G_c$ is isotopic to the identity
but the isotopy need not be in $G_c$. That is, $G_c$ is not
isotopically connected. Observe that the exception is the $C^0$
case: due to Alexander's trick for $r=0$ (see, e.g., \cite{ed-ki},
p.70) $G_c$ is isotopically connected.
Likewise, let $C=\R\t\mathbb S^1$ be the annulus and let
$G=\diff^r(C)_0$. Then there is the twisting number epimorphism
$T:G_c\r\mathbb Z$. It is easily seen that $f\in G_c$ is joined to
id by a compactly supported isotopy iff $T(f)= 0$. Consequently,
$G_c$ is not isotopically connected.
\end{rem}
Denote by $\frak G^r(M)$ (resp. $\frak G^r_c(M)$), $1\leq
r\leq\infty$, the totality of isotopically connected (resp.
isotopically connected through compactly supported isotopies)
groups of $C^r$ diffeomorphisms of $M$. Next the symbol $\frak
F^r(M)$ will stand for the totality of foliations of class $C^r$
on $M$. Then each $G\in\frak G^r(M)$ determines uniquely a
foliation from $\frak F^r(M)$, denoted by $\F_G$. That is, we have
the mapping $\beta_M: \frak G^r(M)\ni G\mapsto \F_G\in\frak
F^r(M)$. Conversely, to any foliation $\F\in\frak F^r(M)$ we
assign $G_{\F}:=\diff^r_c(M,\F)_0$ and we get the mapping
$\alpha_M:\frak F^r(M)\ni \F\mapsto G_{\F}\in \frak G^r_c(M)$. The
following is obvious.
\begin{prop}
One has $\beta_M\ci\alpha_M=\id_{\frak F^r(M)}$. In particular
$$\alpha_M:\frak F^r(M)\ni \F\mapsto
G_{\F}\in\frak G_c^r(M)$$ is an injection identifying the class
of $C^r$-foliations with a subclass of $C^r$-diffeo\-mor\-phism
groups.
\end{prop}
Observe that
usually $(\alpha_M\ci\beta_M)(G)\in\frak G^r_c(M)$ is not a
subgroup of $G$ even if $G\in\frak G^r_c(M)$. For instance, take
the group of Hamiltonian diffeomorphisms of a Poisson manifold,
see \cite{va}. See also examples in \cite{ry0}.
\begin{rem}
Note that we can also define $\alpha'_M:\frak F^r(M)\ni \F\mapsto
G'_{\F}\in \frak G^r(M)$, where $G'_{\F}:=\diff^r(M,\F)_0\in\frak
G^r(M)$, and we get another identification of the class of
$C^r$-foliations with a subclass of $C^r$-diffeo\-mor\-phism
groups. However we prefer $\alpha_M$ to $\alpha'_M$ because of
Proposition 2.11 below.
\end{rem}
For $\F_1,\F_2\in\frak F^r(M)$ we say that $\F_1$ is a
\emph{subfoliation} of $\F_2$ if each leaf of $\F_1$ is contained
in a leaf of $\F_2$. We then write $\F_1\prec\F_2$. By a
\emph{flag structure} we mean a finite sequence
$\F_1\prec\cdots\prec\F_k$ of foliations of $M$. Next, by the
\emph{intersection} of $\F_1, \F_2$ we mean the partition
$\F_1\bar\cap\F_2:=\{L_1\cap L_2:\, L_i\in\F_i, i=1,2\}$ of $M$.
Clearly, if $\F_1\bar\cap\F_2$ is a foliation then
$\F_1\bar\cap\F_2\prec\F_i$, $i=1,2$.
It is a rare phenomenon that $\F_1\bar\cap\F_2$ would be a regular
foliation, provided $\F_1, \F_2$ are regular. In the category of
(singular) foliations it may happen more often.
\begin{prop}
\begin{enumerate}
\item If the distribution $\D(\F_1\bar\cap\F_2)$ is of class $C^r$
(\cite {st0}) then $\F_1\bar\cap\F_2$ is a foliation. \item If
$G_1,G_2\in\frak G^r(M)$ have the intersection $G=G_1\cap G_2$
isotopically connected then $\F_G=\F_{G_1}\bar\cap\F_{G_2}$. \item
For $\F_1,\F_2\in\frak F^r(M)$, if the intersection
$\F_1\bar\cap\F_2$ is a foliation then there is $G\in\frak G^r(M)$
such that $G\leq G_{\F_1}\cap G_{\F_2}$ and
$\F_G=\F_1\bar\cap\F_2$. \item For $\F_1,\F_2\in\frak F^r(M)$, if
$G_{\F_1}\cap G_{\F_2}$ is isotopically connected then the
intersection $\F_1\bar\cap\F_2$ is a foliation.
\end{enumerate}
\end{prop}
\begin{proof}
(1) In fact, the distribution of $\F_1\bar\cap\F_2$ is then
integrable.
(2) Denote by $\I G$ the set of all isotopies in $G$. Clearly,
$\I(G_1\cap G_2)=\I{G_1}\cap\I{G_2}$ for arbitrary $G_1,
G_2\in\frak G^r(M)$. For $x\in M$, set $\I G(x):=\{y\in M:
(\exists f\in\I G)(\exists t\in I)\, y=f_t(x)\}$. By definition,
$L_x=\I G(x)$, where $L_x\in\F_G$ is a leaf meeting $x$.
Therefore, since $ G_1, G_2, G$ are isotopically connected we have
$L_x=\I G(x)=\I G_1(x)\cap \I G_2(x)=L^1_x\cap L_x^2$, where
$L^i_x\in\F_{G_i}$, $i=1,2$.
(3) Set $\F=\F_1\bar\cap\F_2$ and $G=G_{\F}$. Use Prop. 2.4.
(4) In view of Prop. 2.4 we have $\F_{G_{\F_0}}=\F_0$ for all
$\F_0\in\frak F^r(M)$. Put $G=G_{\F_1}\cap G_{\F_2}$. Therefore,
in view of (2),
$\F_1\bar\cap\F_2=\F_{G_{\F_1}}\bar\cap\F_{G_{\F_2}}=\F_G$ is a
foliation.
\end{proof}
Let $\F_1\prec\cdots\prec\F_k$ be a flag structure on $M$ and let
$x\in M$. If $x\in L_i\in \F_i$ we write
$p_i(x)=\dim L_i$, $\bar p_i(x)=p_i(x)-p_{i-1}(x)$ $(i=2\ld k)$ and
$q_i(x)=m-p_i(x)$.
\begin{dff} A chart $(U,\phi)$ of $M$ with $\phi(0)=x$ is called a {\it
distinguished chart} at $x$ with respect to
$\F_1\prec\cdots\prec\F_k$ if $U=V_1\t\cdots\t V_k \t W$ such that
$V_1\s \R^{p_1(x)}$,
$V_i\s\R^{\bar p_i(x)}$ $(i\geq 2)$
and $W\s\R^{q_k(x)}$ are open balls and for any $L_i\in \F_i$ we
have
$$
\phi(U)\cap L_i=\phi (V_1\t\cdots\t V_i\t l_i), $$ where
$l_i=\{w\in V_{i+1}\t\cdots\t V_k\t W:\ \phi(0,w)\in L_i\}$ for
$i=1\ld k$.
\end{dff}
Observe that actually the above $\phi$ is an inverse chart;
following \cite{st2} we call it a chart for simplicity. Notice as
well that in the above definition one need not assume that $\F_i$
is a foliation but only that it is a partition by weakly imbedded
submanifolds; that $\F_i$ is a foliation follows then by
definition.
\begin{thm} Let $G_1\leq\ldots\leq G_k\leq\diff^r(M)$ be an increasing sequence of
diffeomorphism groups of $M$. Then
$\F_{G_1}\prec\cdots\prec\F_{G_k}$ admits a distinguished chart at
any $x\in M$. \end{thm}
In fact, it is a straightforward consequence of Theorem 2 in
\cite{ry0}.
\begin{cor} Let $G_1\leq\ldots\leq G_k\leq\diff^r(M)$ and let
$(L,\sigma)$ be a leaf of $\F_{G_k}$. Then all $G_i$ preserve $L$,
and $\F_{G_1|L}\prec\cdots\prec\F_{G_{k-1}|L}$ is a flag structure
on $L$. Moreover, a distinguished chart at $x$ for
$\F_{G_1|L}\prec\cdots\prec\F_{G_{k-1}|L}$ is the restriction to
$L$ of a distinguished chart at $x$ for
$\F_{G_1}\prec\cdots\prec\F_{G_k}$.
\end{cor}
The following property of paracompact spaces is well-known.
\begin{lem}
If $X$ is a paracompact space and $\U$ is an open cover of $X$,
then there exists an open cover $\V$ starwise finer than $\U$,
that is
for all $V\in \V$ there
is $U\in\U$ such that $\st^{\V}(V)\s U$. Here
$\st^{\V}(V):=\bigcup\{V'\in\V:\; V'\cap V\neq\emptyset\}$. In
particular, for all $V_1, V_2\in \V$ with $V_1\cap
V_2\neq\emptyset$ there is $U\in\U$ such that $V_1\cup V_2\subset
U$.
\end{lem}
\begin{prop}
If $\F\in\frak F^r(M)$ then $G_{\F}=\a(\F)$ is factorizable.
\end{prop}
\begin{proof}
Let $\frak X_c(M,\F)$ be the Lie algebra of all compactly
supported vector fields on $M$ tangent to $\F$. Then there is a
one-to-one correspondence between isotopies $f_t$ in $G_{\F}$ and
smooth paths $X_t$ in $\frak X_c(M,\F)$ given by the equation
\begin{equation*}
\frac{df_t}{dt}=X_t\circ f_t \quad \mathrm{with} \quad f_0=\id.
\end{equation*}
Let $f=(f_t)\in\I G_{\F}$ and let $X_t$ be the corresponding
family in $\frak{X}_c(M,\F)$. By considering
$f_{(p/m)t}f^{-1}_{(p-1/m)t}$, $p=1,\dots ,m$, instead of $f_t$ we
may assume that $f_t$ is close to the identity.
Let $\U$ be an open cover of $M$.
We choose a
family of open sets, $(V_j)_{j=1}^s$, which
is starwise finer than $\U$, and satisfies $\supp(f_t)\subset
V_1\cup\dots \cup V_s$ for each $t$. Let $(\lambda_j)_{j=1}^s$ be
a partition of unity subordinate to $(V_j)$, and let
$Y_t^j=\lambda_jX_t$. We set
$$X_t^j=Y_t^1+\dots +Y_t^j,\quad j=1,\dots ,s,$$
and $X_t^0=0$. Each of the smooth families $X_t^j$ integrates to
an isotopy $g_t^j$ with support in $V_1\cup\dots \cup V_j$. We get
the fragmentation
$$f_t=g_t^s=f_t^s\circ \dots \circ f_t^1,$$
where $f_t^j=g_t^j\circ (g_t^{j-1})^{-1}$, with the required
inclusions
$$\supp(f_t^j)=\supp(g_t^j\circ (g_t^{j-1})^{-1})\subset \st(V_j)\subset U_{i(j)}$$
which hold if $f_t$ is sufficiently small. Thus the group of
isotopies of $G_{\F}$ is factorizable. Consequently, $G_{\F}$
itself is factorizable.
\end{proof}
\begin{rem} The identification $\a_M$ enables us to consider several
new properties of foliations from $\frak F^r(M)$. For instance,
one can say that a foliation $\F$ is \emph{perfect} if so is the
corresponding diffeomorphisms group $G_{\F}=\a_M(\F)$. As we
mentioned before it is known that $G_{\F}=\diff^r_c(M,\mathcal
F)_0$ is perfect provided $\F$ is regular and
$1\leq r\leq\dim\F$ or $r=\infty$ (\cite{ry1}, \cite{Tsu1}, \cite{ry2}).
It is not known whether $G_{\F}$ is perfect for singular foliations and
a possible proof seems to be very difficult. In turn, a possible perfectness of $G_{\F}=\diff^r_c(M,\mathcal
F)_0$ with $r$ large is closely related to the simplicity of
$\diff^{n+1}_c(M^n)_0$, see \cite {le-ry}.
Likewise, one can consider \emph{uniformly perfect} or
\emph{bounded} foliations by using the corresponding notions for
groups, see \cite{ry6} and references therein.
\end{rem}
Finally consider the following important feature of subclasses of
the class $\frak F^r(M)$, depending also on $M$ and $r$. A
subclass $\frak K$ of $\frak F^r(M)$ is called \emph{faithful} if
the following holds: For all $\F_1,\F_2\in\frak K$ and for any
group isomorphism $\Phi:\alpha_M(\F_1)\cong\alpha_M(\F_2)$ there
is a $C^r$ foliated diffeomorphism $\phi:(M,\F_1)\cong(M,\F_2)$
such that $\forall f\in\alpha_M(\F_1)$, $\Phi(f)=\phi\ci
f\ci\phi^{-1}$. From reconstruction results of Rybicki \cite{ry4}
and Rubin \cite{Rub} it is known that the class of regular
foliations of class $C^{\infty}$, $\frak F_{reg}^{\infty}(M)$, is
faithful.
\section{Proof of Theorem 1.1 and 1.2}
\emph{Proof of Theorem 1.1}. First observe that the fact that a
foliation $\F$ has no proper minimal set is equivalent to the
statement that all leaves of $\F$ are dense.
$(\Rightarrow)$ Assume that $\emptyset\neq L\s M$ is a proper
closed saturated subset of $M$. Choose $x\in M\setminus L$. We
prove the following statement:
$(*)$ there are a ball $U\s M\setminus L$ with $x\in U$ and $g\in
[G_U,G_U]$ such that $g(x)\neq x$.
We are done in view of $(*)$ by setting $H:=\{g\in [G,G]:
g|_L=\id_L\}$. To prove $(*)$, choose balls $U$ and $V$ in $M$
such that $x\in V\s \overline{V}\s U$. Take $f\in G$ such that
$f(x)\neq x$. In light of the assumption, for $\U=\{U,
\setminus\overline V\}$
we may write $g=g_r\ldots g_1$,
where all $g_i$ are supported in elements of $\U$. Let
$s:=\min\{i\in\{1\ld r\}:\; \supp(g_i)\s U \text{\; and\;}
g_i(x)\neq x\}$. Then $g_s\in G_U$ satisfies $g_s(x)\neq x$.
Now take an open $W$ such that $x\in W\s U$ and $g_s(x)\not\in W$.
Choose $f\in G_W$ with $f(x)\neq x$ by an argument similar to the
above. It follows that $f(g_s(x))=g_s(x)\neq g_s(f(x))$ and,
therefore, $[f,g_s](x)\neq x$. Thus $g=[f,g_s]$ satisfies the
claim.
$(\Leftarrow)$ First observe the following commutator formulae for
all $f,g,h\in G$
\begin{equation}
[fg,h]=f[g,h]f^{-1}[f,h],\quad [f,gh]=[f,g]g[f,h]g^{-1}.
\end{equation}
Next, in view of a theorem of Ling \cite{li} we have that $[G,G]$
is a perfect group, that is
\begin{equation}
[G,G]=[[G,G],[G,G]].
\end{equation}
Suppose that $H$ is a nontrivial normal subgroup of $[G,G]$. Let
$x\in M$ satisfy $h(x)\neq x$ for some $h\in H$. Fix a ball $U_0$
such that $h(U_0)\cap U_0=\emptyset$. By the definition of
$\F_{[G,G]}$ and the assumption that each leaf $L\in\F_{[G,G]}$ is
dense, for every $y\in M$ there are a ball $U_y$ with $y\in U_y$
and $f_y\in [G,G]$ such that $f_y(U_y)\s U$. Let $\U=\{U_y\}_{y\in
M}$.
Due to Lemma 2.10 we can find an open cover $\V$ starwise finer
than $\U$. We denote $\U^G=\{g(U)|\, U\in\U,\, g\in[G,G]\}$ and
\begin{equation*}G^{\U}=\prod\limits_{U\in \U^G}[G_U,G_U].\end{equation*}
By assumption $G$ is factorizable with respect to $\V$.
First we show that
$[G,G]\s G^{\U}$, i.e. that any $[g_1,g_2]\in[G,G]$ can be
expressed as a product elements of $G^{\U}$ of the form
$[h_1,h_2]$, where $h_1,h_2\in G_U$ for some $U\in\U^G$. In view
of (3.1) and (3.2) we may assume that $g_1, g_2\in[G,G]$. Now the
relation $[G,G]\s G^{\U}$ is an immediate consequence of (3.1)
and the fact that $\V$ is starwise finer than $\U$.
Next we have to show that $G^{\U}\s H$. It suffices to check that for
every $f,g\in G_U$ with $U\in \U$ the bracket $[f,g]$ belongs to
$H$. This implies that for every $f,g\in G_U$ with $U\in\U^G$ one
has $[f,g]\in H$, since $H$ is a normal subgroup in $[G,G]$.
We have fixed $h\in H$ and $U_0$ such that $h(U_0)\cap
U_0=\emptyset$. If $U\in\U$ and $f,g\in G_U$, take $k\in[G,G]$
such that $k(U)\s U_0$, and put $\bar f=kfk^{-1}$, $\bar
g=kgk^{-1}$. It follows that $[h\bar fh^{-1}, \bar g]=\id$.
Therefore, $[\bar f,\bar g]=[[h,\bar f],\bar g]\in H$, and we have
also that $[f,g]\in H$. Thus we have $G^{\U}\s H$ and,
consequently, $[G,G]\leq H$, as required. \quad $\square$
\medskip
\emph{Proof of Theorem 1.2}. By assumption and Prop. 2.11
$\diff^r(M,\mathcal F)_0$ is factorizable and non-fixing. Since
$\diff^r(M,\mathcal F)_0$ is isotopically connected, the first
assertion follows from Theorem 1.1. The second assertion is a
consequence of theorems stating that $\diff^r(M,\mathcal F)_0$ is
perfect (\cite{ry1} and \cite{Tsu1} for $r=\infty$, and \cite{Mat}
and \cite{ry2} for $1\leq r\leq\dim\F$). \quad $\square$
|
\section{Introduction}
\subsection{Stellar winds of O stars}
Stellar winds of early (O or B) type stars are driven by the radiation pressure of copious absorption lines
present in the ultraviolet part of the spectrum on material in the stellar atmosphere \citep{castor75}.
Therefore the winds are very strong; common mass loss rates are $\sim$10$^{-6} M_{\odot}/\mathrm{year}$.
Since primaries of high-mass X-ray binaries are O or early B stars \citep{conti78}, which radiate in the UV, this radiation is strong
enough to produce such a wind.
According to simulations of line-driven winds, perturbations are present and dense and cool
inhomogeneities are created in the wind \citep{feldmeier97}. Larger density, velocity, and temperature variations compress the gas further, creating ``clumps''. Current knowledge about stellar winds assumes two disjunct components of O star winds:
cool dense clumps and hot tenuous gas. Sako et al. \citep{sako02} showed that observed spectra of X-ray binaries can only be explained as originating from an environment, where the cool and dense clumps are embedded in the
photoionized gas.
\subsection{Cygnus X-1}
Cygnus X-1 is a binary system where the X-ray source is a black hole \citep{bolton72, webster72}, and $\sim18\,M_\odot$ \citep{herrero95}, O 9.7 Iab type star HDE 226868 is its companion \citep{walborn73}. Stellar wind accretion plays a major role in the mass transfer process, because \mbox{Cyg X-1} belongs to the High-Mass X-ray Binaries (HMXB), which is in contrast to Low-mass X-ray Binaries (LMXB), where Roche lobe overflow is more important and accretion disk accretion occurs.
There are strong tidal interactions in the system. Moreover, the donor star fills $\sim$90\,\% of its Roche volume \citep{conti78,giesbolton86b}. Therefore the wind is not symmetric, but \emph{focused} towards the black hole \citep{friend82}, such that density and mass loss rate are higher along the binary axis.
The fact that such a high percentage of the Roche lobe is filled, however, means that we cannot exclude Roche lobe overflow taking place as well.
\subsection{Hard and soft state of Cygnus X-1}
Black hole binaries show two principal types of emission called the hard or soft state, which differ in the shape of the X-ray spectrum, the timing properties and the radio emission. Cyg X-1 spends most of the time in the hard state source with a hard, exponentially cut-off powerlaw spectrum, strong short term variability and steady radio emission \citep{pottschmidt03,wilms06}. However, transitions between hard and soft states are observed (Fig.~\ref{RXTE-ASM_0}).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{RXTE-ASM_-1.eps}}
\caption{$1.5 - 12\, \mathrm{keV}$ light curve of Cyg X-1 obtained from 1996 to 2010 by the \textsl{RXTE}-ASM. Over time, transitions from hard to soft state occur. Compared to the time spent in the low-luminosity \emph{hard} state (blue, $\lesssim\:$40\,c/s for ASM countrate) Cyg X-1 spent less time in the high-luminosity \emph{soft} state (red, $\sim$100\,c/s for ASM countrate).
\label{RXTE-ASM_0}}
\end{figure}
How exactly the wind properties differ between states and what triggers state transitions are questions remaining to be answered. One possibility is that states correspond to different configurations of the accretion flow \citep{smith02}, which cause differences in energy dissipation. Another possibility is that changes of the wind properties themselves trigger state transitions \citep{gies03}.
The mass transfer process in either case (HMXB or LMXB) provides extremely efficient energy release and produces luminosities of $\sim$ $10^{37}$ $\mathrm{erg/s}$ in general. For such a luminosity, which is typical for Cyg X-1, the X-ray source produces a considerable feedback on the wind by photoionization of its nearby environment \citep{blondin94}, which contributes to the complex wind structure.
\section{Observations and data analysis}
\subsection{Observations and orbital coverage}
The spectra used for our analysis were obtained by the High Energy Transmission Grating Spectrometer (HETGS -- HETG in combination with ACIS, Advanced CCD Imaging Spectrometer) on board the \textsl{Chandra} observatory. \textsl{Chandra} ACIS observations are performed in two different modes: timed exposure (TE) mode and continuous clocking (CC) mode. In TE mode, the CCD is exposed for some time and then its data are transfered to the frame store, which is read out during the next exposure. The readout time required for the full frame store is 3.2s. In CC mode, the columns are read out continuously, which reduces the readout time to 3ms \citep{garmire03}\footnote{Chandra X-ray Center, The Chandra Proposers' Observatory Guide, 2009, \url{http://cxc.harvard.edu/proposer/POG/}}. When the source is very bright, more than one photon may reach the same pixel in one frame time (pile-up). These photons are misinterpreted as one single event with higher energy. CC mode is usually used to avoid pile-up.
High-resolution spectra of persistently bright sources like Cyg X-1 provide the unique possibility of probing the structure of the wind \emph{directly}.
However, this structure and therefore also the properties of the wind (density, velocity, ionization state) change with different lines of sight, which correspond to the different orbital phases. Thus, a good coverage of the binary orbit is desirable.
\begin{figure
\begin{minipage}[b]{.4\columnwidth}
a)
\includegraphics[width=\columnwidth,angle=0.]{Chandra_coverage_1.eps}
\end{minipage}
\begin{minipage}[b]{.15\columnwidth}
b)
\includegraphics[width=\columnwidth,angle=0.]{Chandra_coverage_xz.eps}
\end{minipage}
\begin{minipage}[b]{.4\columnwidth}
c)
\includegraphics[width=\columnwidth,angle=0.]{Chandra_coverage.eps}
\end{minipage}
\caption{a) Polar view of the Cyg X-1 orbit and illustration of observation coverage with \textsl{Chandra}. Before 2010, January, there were 13 observations available, mostly covering the part of the orbit around phase $\phi{\approx}$0. Another observation (ObsID 11044) was obtained at $\phi{\approx}$0.5. A short observation (ObsID 2742) was the only one at this phase before, but since it was obtained in TE mode during the soft state of the source, it strongly suffered from pile-up.
Full lines (dashed lines) display TE mode (CC mode) observations. Changes from blue to red colour correspond to changes from hard to soft state.
b)~Colorcoded orbital phases corresponding to lines of sight towards Cyg X-1 taking into account an inclination of $\sim$ 35$^{\circ}$.
c)~Highlighted observations at phases of $\phi{\approx}$0.0 (ObsID 3814 and ObsID 8525), 0.2 (ObsID 9847), 0.5 (ObsID 11044) and 0.75 (ObsID 3815), which are analyzed and compared in this work.
\label{Chandra_coverage}}
\end{figure}
Observations (whether in the hard or soft state) which are currently available cover part of the orbit around phase $0.0$, between phases $0.7$ and $0.2$, and around phase $0.5$, with the latter only obtained in January, 2010 (Fig. \ref{Chandra_coverage}a). We focus here on the comparison of observations obtained at four distinct phases of $\phi=0.0$, which is defined at the time of superior conjunction of the black hole (ObsID 3814 and ObsID 8525), $0.2$ (ObsID 9847), $ 0.5$ (ObsID 11044) and $0.75$ (ObsID 3815), see Fig. \ref{Chandra_coverage}c.
The latter was obtained in CC mode, while all others were obtained in TE mode. This difference has no influence on our comparison. While the calibration of CC mode does not allow for an adequate modelling of the whole continuum shape, local absorption lines, which are our primary interest, are not affected.
The phase $0.0$ coverage is extremely important, since due to the inclination of $\sim$ 35$^{\circ}$ of the Cyg X-1 orbital plane, it corresponds to looking through the densest part of the wind close to the stellar surface (Fig. \ref{Chandra_coverage}b). The distribution of X-ray dips with orbital phase peaks around phase $0.0$ \citep{balucinska00}. The observation around phase $0.5$ provides a great opportunity to close a gap in defining the general picture of the wind structure.
While all recent \textsl{Chandra} observations caught Cyg\,X-1 in the hard state at $\lesssim\:$100\,c/s (of \textsl{Chandra} countrate), comparable to the observation at $\phi{\approx}$0 \citep{hanke09}, the spectrum was softer and the flux was more than twice as high during the observation at $\phi{\approx}$0.7. The light curves at $\phi{\approx}$0 are modulated by strong and complex \emph{absorption\, dips}, but dipping occurs already at $\phi{\approx}$0.7 and has not ceased at $\phi{\approx}$0.2, though the dip events seem to become shorter with distance from $\phi{=}$0. The light curve at $\phi{\approx}$0.5 is totally \emph{free of dips}, yielding 30\,ks of remarkably constant flux.
\subsection{Absorption dips}
According to general assumption, absorption dips, during which the soft X-ray flux decreases sharply, originate from inhomogeneities --``clumps''-- present in the wind, where the material is of higher density and lower temperature \citep{castor75,sako02}. According to the softness ratios in the color-color diagram (Fig. \ref{color-color}), different stages of dipping can be classified.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{color-color.eps}}
\caption{ All of these diagrams show a ``soft softness ratio" on the $x$- and a ``hard softness ratio" on the $y$-axis. Dipping produces a clear track in the color-color diagrams:
Both colors harden towards the lower left corner, due to increased absorption. However, during extreme dips, the soft color becomes softer again, which is likely due to partial coverage.
\label{color-color}}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{flux_Si-region3815.eps}}
\caption{ ``Dip'' and ``non-dip'' spectra from the observation at phase $\phi{\approx}$0.7 are shown for comparison of absorption lines present in different stages of dipping. The \emph{reduction in flux} in the spectra is real and due to dips.
In the "non-dip" spectrum, only absorption lines of Si\,\textsc{xiv} and Si\,\textsc{xiii} are present, while in the "dip" spectrum the whole series of Si\,\textsc{xii}--\textsc{vii} appears.
\label{flux_Si-region}}
\end{figure}
Figure~\ref{flux_Si-region} shows the spectrum from the observation at phase $\phi{\approx}$0.7 (ObsID\,3815) split into "dip" and "non-dip" stages in the wavelength interval of the Si-region between 6\,\AA\, and 7.5\,\AA.
While absorption lines of Si\,\textsc{xiv} and Si\,\textsc{xiii} are already present in the non-dip spectrum, the dip spectra contain \emph{additional strong absorption lines} that can be identified with K$\alpha$ transitions of lower ionized Si\,\textsc{xii}--\textsc{vii}.
The strength of the low-ionization lines increases with the degree of dipping, indicating that the latter is related to clumps of lower temperature. Moreover, the clumps are of higher density than their surroundings.
In 2008 our group organized a multi-satellite observational campaign with \textsl{XMM-Newton},\textsl{Chandra} (ObsID 8525 and ObsID 9847), \textsl{Suzaku}, \textsl{RXTE}, \textsl{INTEGRAL} and \textsl{Swift} observing Cyg X-1 simultaneously. Dips shortly after phase $\phi{\approx}$0 were so strong, that they were seen by all the instruments involved in the campaign, even \textsl{RXTE}--PCA or \textsl{INTEGRAL}--ISGRI \citep{hanke08}.
The light curve from XMM-Newton is shown in Fig.~\ref{XMM-EPIC-pn} \citep{hanke10}. Where the dips occur in the light curve, the hydrogen column density, $N_{\mathrm{H}}$, increases strongly.
\begin{figure}[!h]
\resizebox{\hsize}{!}{\includegraphics{XMM-EPIC-pn.eps}}
\caption{Absorption dips and scattering of X-rays as seen with \textsl{XMM-Newton}, EPIC-pn.
a) The light curve in the energy band $0.3 - 10\, \mathrm{keV}$ shows absorption dips, which are identical to dips observed by \textsl{Chandra} shortly after $\phi{\approx}0$.
b) Hydrogen column density of the neutral absorption model increases strongly when dips occur.
c) Relative flux normalization constant, which is consistent with the scattering trough seen in hard X-rays \citep{hanke10}.
\label{XMM-EPIC-pn}}
\end{figure}
As shown in the third panel, however it is not only the pure absorption which causes the dips. Thomson scattering contributes during these times, causing longer time scale variations, also at hard X-rays. A possible explanation is in the existence of dense and (nearly) neutral clumps (causing the sharp dips) embedded in ionized halos (causing the scattering) \citep{hanke10}.
\subsection{Spectroscopy}
We separate the ``non-dip" and the ``dip" parts of the observations. The ``non-dip" spectrum is extracted from the least absorbed phases at the upper right corner of the color-color diagram (except for ObsID\,3815) and the spectroscopic results here refer to these ``non-dip" phases.
The highly photoionized wind is detected at $\phi{\approx}0$ via numerous strong absorption lines at $v_{\rm rad}{\approx}0$ (Fig. \ref{H-lineProfiles}).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{H-lineProfiles.eps}}
\caption{ Absorption and P Cygni profiles of Si\,\textsc{xiv}, Mg\,\textsc{xii} and Ne\,\textsc{x}.
While the observations at $\phi{\approx}0$ and $\phi{\approx}0.75$ show clear absorption profiles, although redshifted for $\phi{\approx}0.75$, the observation at $\phi{\approx}0.5$ shows emission at $v_{\rm rad}{\approx}0$ and blueshifted absorption.
\label{H-lineProfiles}}
\end{figure}
The lack of appreciable Doppler shifts can be explained by the wind flow being orthogonal to the line of sight.
In contrast, the recent observation (ObsID 11044) at $\phi{\approx}0.5$ reveals \emph{for the first time for Cyg X-1} clear \emph{P\,Cygni profiles} with a strong emission component at a projected velocity $v_{\rm rad}{\approx}0$, while the weak absorption components occur at a blueshift of $\approx$500--1\,000\,km/s.
If we observe the same plasma in both cases, this indicates that the real velocity must be small, i.e., we are probing a dense, low-velocity wind close to the stellar surface.
The fact that the absorption line profiles measured at $\phi{\approx}0.75$ are redshifted by $\approx$200--300\,km/s indicates that the wind flow is not radial from the star, as a radial wind (i.e., directing away from the star) would always give a blueshifted velocity when projected onto the line of sight at phases $\phi=0.25 - 0.75$.
\section{Summary}
The new Chandra observation of Cygnus X-1 at orbital phase 0.5 obtained in January 2010 allows us to compare observations at the four distinct orbital phases 0.0, 0.2, 0.5 and 0.75. With such a coverage, the full structure of the wind starts to reveal itself. At phase 0.0 we look through the densest part of the wind, as it is focused towards the black hole. The light curve is modulated by strong absorption dips. The flux decreases strongly during such dips, consistent with being caused by dense and cool clumps of material embedded in the more ionized wind.
While absorption lines of Si\,\textsc{xiv} and Si\,\textsc{xiii} are already present in the non-dip spectrum, in the dip spectra also K$\alpha$ transitions of lower ionized Si appear, whereas the strength of these lines increases with the degree of dipping.
An especially interesting result is the totally flat light curve around phase 0.5. While dipping has started around phase 0.7, is the strongest around 0.0 and still present at 0.2, it has vanished at 0.5. We therefore proposed for the next observation between phases 0.25 and 0.4 to investigate the transition between dipping and non-dipping phases.
Spectroscopic analysis showed another interesting result. In the spectrum at phase 0.5, clear P-Cygni profiles of Lyman $\alpha$ transitions were observed for the first time for Cyg X-1. We observe here strong emission components at a projected velocity $v_{rad}{\approx}0$ in contrast to pure absorption observed at phase 0.0.
Detailed modeling of photoionization and wind structure is in progress.
\paragraph{Acknowledgements}
The research leading to these results was funded by the European Community's Seventh Framework Programme (FP7/2007-2013) under grant agreement number ITN 215212 "Black Hole Universe" and by the Bundesministerium f\"ur Wirtschaft und Technologie under grant number DLR 50 OR 0701.
|
\section{Introduction}
The understanding of spin relaxation dynamics in semiconductor structures is an active area of research related to the field of spintronics \cite{Zutic04a,Wu10a}. Previously, the spin relaxation dynamics has been investigated mainly in infinite two-dimensional (2D) systems with D'yakonov-Perel' \cite{Dyakonov72a,Dyakonov86a} spin relaxation mechanism (see, e.g., Refs. \cite{Dyakonov86a,Sherman03a,Pershin05a,Bernevig06a,Weng08a,Kleinert09a,Tokatly10a,pershin10a}). There are only several examples in the literature
where the effects of boundary conditions \cite{Galitski06a} on D'yakonov-Perel' spin relaxation have been explored. These examples include investigations of spin relaxation in 2D channels \cite{Kiselev00a,Holleitner07a}, 2D half-space \cite{Pershin05c}, 2D systems with antidots \cite{Pershin04a}, small 2D systems \cite{Lyubinskiy06a}, and finite-length one-dimensional (1D) structures \cite{Slipko11a}. The main conclusion of all these studies is that the introduction of boundary conditions results in an increased spin lifetime. In particular, in Ref. \cite{Lyubinskiy06a} Lyubinskiy reports on a strong suppression of spin relaxation in small 2D systems. In Ref. \cite{Slipko11a} the present authors demonstrate that a persistent spin helix spontaneously emerges in course of relaxation of homogeneous spin polarization. As real electronic devices are always of a finite size, the understanding of spin relaxation dynamics in reduced geometries is of a crucial importance.
In this paper, we present our studies of spin relaxation in rings with Bychkov-Rashba \cite{Bychkov84a} spin-orbit coupling. The ring geometry is especially interesting since on the one hand, the electron space motion in rings is confined to a limited space region and on the other hand, the ring geometry does not require any boundary conditions (such as those derived in Ref. \cite{Galitski06a}) for 1D transport that normally have to be used, for example, to describe spin dynamics in finite-length structures \cite{Slipko11a}. Starting with a spin kinetic equation, we formulate a set of equations for spin polarization components on the ring and solve these equations in some particular cases. Specifically, we find that a homogeneous spin polarization transforms into a persistent crown-like spin structure schematically shown in Fig. \ref{fig1}. Moreover, we determine a set of non-trivial spin persistent states that are realized at some specific sizes of the ring. Finally, we derive expression for Green function that can be used to find evolution of spin polarization for any initial spin polarization profile. All these results indicate an unusual character of spin relaxation in finite size structures that is dramatically different from the spin relaxation in the bulk.
\begin{figure} [b]
\begin{center}
\includegraphics[angle=0,width=5.0cm]{fig1}
\caption{(Color online) Schematic diagram of spin relaxation in a ring. Initially homogeneous
spin polarization in $z$ direction (along the ring axis) transforms into a persistent crown-like spin structure. Note that the spin polarization
amplitude decreases in this process according to Eq. (\ref{SzAttenuation}).}
\label{fig1}
\end{center}
\end{figure}
\section{Spin kinetic equation}
In this section, we introduce the model and derive equations governing spin relaxation dynamics in rings. The main results of this section are given by Eqs. (\ref{Sr})-(\ref{Sz}). The derivation of these equations is based on a kinetic approach. It is important to mention that Eqs. (\ref{Sr})-(\ref{Sz}) are more general than traditional spin drift-diffusion equations \cite{Yu02a,Burkov04a,Saikin04a,Pershin04b,arXiv_1007_0853v1}. In particular, Eqs. (\ref{Sr})-(\ref{Sz}) are applicable in both diffusive and ballistic regimes at any value of spin rotation angle per mean free path.
Let us consider spin-polarized electrons that can move along a ring of a sufficiently large radius $r\gg \hbar/p$ so that small-radius corrections \cite{Meijer02a,Berche10a} to spin-orbit Hamiltonian \cite{Bychkov84a} can be neglected. The spin-orbit Hamiltonian is taken in the form $H_{SO}=\alpha\left(\hat{\mathbf\sigma}\times\mathbf{\hat p}\right)\cdot\mathbf{e}_z$. Here, $\mathbf{\hat{p}}=(\hat p_x,\hat p_y)$ is the 2D electron momentum operator, $m$ is the effective electron's mass, $\mathbf{\hat\sigma}$ is the Pauli-matrix vector, $\alpha$ is the spin-orbit coupling constant and $\mathbf{e}_z$ is a unit vector along the ring axis. It can be shown that a motion of an electron with a momentum $\mathbf{p}$ along a straight trajectory is accompanied by a spin rotation with the angular velocity $\mathbf{\Omega_p}=\Omega\mathbf{p}\times\mathbf{e}_z/p$, where
$\Omega=2\alpha p/\hbar$. The spin precession angle per unit length is given by $\eta=2\alpha
m \hbar^{-1}$. The ring geometry, however, changes the character of spin rotations since the spin rotation axis rotates as electron moves along the ring.
Our consideration of spin dynamics in rings is based on the kinetic equation for electron spin polarization (see, e.g., Ref. \cite{Lyubinskiy06a,Lyubinskiy06b}). In quasi-classic approximation such an equation can be written as
\begin{eqnarray}
\left(\frac{\partial }{\partial t}+\frac{\mathbf{p}}{m}\cdot\nabla\right)
\mathbf{S_p}=\mathbf{\Omega_p}\times\mathbf{S_p}+St\{\mathbf{S_p}\},
\label{KinEq}
\end{eqnarray}
where $\mathbf{S_p}(\mathbf{r},t)$ is the vector of spin polarization of electrons, and $St\{\mathbf{S_p}\}$ is the collision integral describing electron scattering processes. Eq. (\ref{KinEq}) describes change in spin polarization of electrons moving with momentum $\mathbf{p}$. The RHS of Eq. (\ref{KinEq}) includes two terms describing rotation of spin polarization with the angular velocity $\mathbf{\Omega_p}$ and change in spin polarization due to scattering processes. In the $\tau$-approximation \cite{Pitaevskii81a} the collision integral is given by
\begin{eqnarray}
St\{\mathbf{S_p}\}=-\frac{1}{\tau}(\mathbf{S_p}-\langle\mathbf{S_p}\rangle),
\label{CollisionIntegral}
\end{eqnarray}
where the angle brackets denote averaging over direction of electron momentum.
The collision integral (\ref{CollisionIntegral}) corresponds to the elastic scattering of electrons by strong scatterers with a characteristic time $\tau$ between the collisions.
For 1D case the average spin polarization simplifies to the following expression $\langle\mathbf{S_p}\rangle=(\mathbf{S}^+ +\mathbf{S}^-)/2 $, where $\mathbf{S}^+$ and $\mathbf{S}^-$ are the spin polarizations of electrons moving along the ring in the clockwise (with momentum $\mathbf{p}=mv\mathbf{e}_\theta$), and counterclockwise ($\mathbf{p}=-mv\mathbf{e}_\theta$) directions with the average velocity $v$ connected to the mean-free path $\ell$ by $\ell=v \tau$. Thus, the kinetic equation (\ref{KinEq}) for 1D ring of a radius $r$ takes the form of the system of two vector equations
\begin{eqnarray}
\left(\frac{\partial }{\partial t}+\frac{v}{r}\frac{\partial }{\partial \theta}\right)
\mathbf{S}^+=\Omega\mathbf{e}_r\times\mathbf{S}^+ -\frac{1}{2\tau}(\mathbf{S}^+-\mathbf{S}^-),
\label{KinEqPlus} \\
\left(\frac{\partial }{\partial t}-\frac{v}{r}\frac{\partial }{\partial \theta}\right)
\mathbf{S}^-=-\Omega\mathbf{e}_r\times\mathbf{S}^- -\frac{1}{2\tau}(\mathbf{S}^--\mathbf{S}^+).
\label{KinEqMinus}
\end{eqnarray}
This system of equations should be complimented by the initial conditions for spin polarizations $\mathbf{S}^+(\theta,t=0)$ and $\mathbf{S}^-(\theta,t=0)$ (for clockwise and counterclockwise moving electrons).
Taking the sum and difference of Eqs. (\ref{KinEqPlus}, \ref{KinEqMinus}) we easily obtain
\begin{eqnarray}
\frac{\partial \mathbf{S}}{\partial t}
&=&-\frac{v}{r}\frac{\partial \mathbf{\Delta}}{\partial \theta}
+
\Omega\mathbf{e}_r\times\mathbf{\Delta},
\label{KinEqS} \\
\frac{\partial \mathbf{\Delta}}{\partial t}
&=&-\frac{v}{r}\frac{\partial \mathbf{S}}{\partial \theta}
+
\Omega\mathbf{e}_r\times\mathbf{S}-\frac{\mathbf{\Delta}}{\tau}
\label{KinEqDelta},
\end{eqnarray}
where the following notations are used: $\mathbf{S}=\mathbf{S}^++\mathbf{S}^-$ and $\mathbf{\Delta}=\mathbf{S}^+-\mathbf{S}^-$.
As we are mainly interested in finding the total spin polarization $\mathbf{S}$, $\mathbf{\Delta}$ can be eliminated from Eqs. (\ref{KinEqS}) and (\ref{KinEqDelta}) via a simple transformation. The final equations for three components of electron spin polarization in cylindrical coordinates, $\mathbf{S}=S_r\mathbf{e}_r+S_\theta\mathbf{e}_\theta+S_z\mathbf{e}_z$, can be presented in the form
\begin{eqnarray}
\frac{\partial^2 S_r}{\partial t^2}+\frac{1}{\tau}\frac{\partial S_r}{\partial t}=
\omega^2 \frac{\partial^2 S_r}{\partial \theta^2}-\omega^2 S_r-2\omega^2\frac{\partial S_\theta}{\partial \theta}-\omega\Omega S_z, \label{Sr} \\
\frac{\partial^2 S_\theta}{\partial t^2}+\frac{1}{\tau}\frac{\partial S_\theta}{\partial t}=\omega^2 \frac{\partial^2 S_\theta}{\partial \theta^2}-(\omega^2+\Omega^2)S_\theta \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \label{Stheta} \\ \nonumber
+2\omega^2\frac{\partial S_r}{\partial \theta}+
2\omega\Omega \frac{\partial S_z}{\partial \theta},\\
\frac{\partial^2 S_z}{\partial t^2}+\frac{1}{\tau}\frac{\partial S_z}{\partial t}=
\omega^2 \frac{\partial^2 S_z}{\partial \theta^2}-\Omega^2 S_z- \omega\Omega S_r -2\omega\Omega \frac{\partial S_\theta}{\partial \theta},\label{Sz}
\end{eqnarray}
where $\omega=v/r$ is the angular velocity.
The initial rate of change of spin polarization $\mathbf{S}$ may be calculated by using the initial condition for $\Delta$ from Eq. (\ref{KinEqS}). In particular, if polarizations of electrons moving clockwise and counterclockwise at the initial moment are the same, i.e. $\Delta(\theta,t=0)=0$, then we find from Eq. (\ref{KinEqS}) that in this case
\begin{eqnarray}
\left(\frac{\partial \mathbf{S}}{\partial t}\right)_{t=0}=0.
\label{IC1}
\end{eqnarray}
The diffusive limit Eqs. (\ref{Sr})-(\ref{Sz}) is realized on time scales much longer than $\tau$. In this case, we can neglect $\partial^2 S_i/\partial t^2$ terms in these equations and notice that the resulting equations are of diffusion type with the diffusion coefficient $D=v^2\tau$.
\section{Relaxation of homogeneous spin polarization}
The most interesting and simple solution of Eqs. (\ref{Sr})-(\ref{Sz}) describes the relaxation of the homogeneous spin polarization initially directed along $z$-axis (see Fig.\ref{fig1})
\begin{eqnarray}
\left(\mathbf{S}\right)_{t=0}=S_0\mathbf{e}_z,
\label{IC2}
\end{eqnarray}
where $S_0$ is the initial spin polarization amplitude. Since the initial condition (\ref{IC2}) is symmetrical with respect to rotations about the ring axis, $S_\theta=0$ at anytime, and the spin polarization components $S_r$ and $S_z$ should not depend on $\theta$. Then Eq. (\ref{Stheta})
satisfies identically, and Eqs. (\ref{Sr}) and (\ref{Sz}) can be simplified to the following relations
\begin{eqnarray}
\frac{d^2 S_r}{d t^2}+\frac{1}{\tau}\frac{d S_r}{d t}
+\omega^2 S_r+\omega\Omega S_z=0,
\label{SrHomEq}\\
\frac{d^2 S_z}{d t^2}+\frac{1}{\tau}\frac{d S_z}{d t}
+\Omega^2 S_z+\omega\Omega S_r=0.
\label{SzHomEq}
\end{eqnarray}
Solving these simple equations with the initial conditions (\ref{IC1}) and (\ref{IC2}), we find that the relaxation of homogeneous spin polarization initially directed along the ring's axis is described by
\begin{eqnarray}
S_r(t)=-\frac{\omega\Omega S_0}{\omega^2+\Omega^2}
\left[1- e^{-\frac{t}{2\tau}}\left(
\cosh\kappa t+
\frac{\sinh\kappa t}{2\tau\kappa}\right)\right], \;\;\;
\label{SrHom}\\
S_z(t)=\frac{\omega^2 S_0}{\omega^2+\Omega^2}
\left[1+\frac{\Omega^2}{\omega^2}e^{-\frac{t}{2\tau}}\left(
\cosh\kappa t+
\frac{\sinh\kappa t}{2\tau\kappa}\right)\right], \;\;\;
\label{SzHom}
\end{eqnarray}
where $\kappa=\sqrt{\frac{1}{4\tau^2}-(\omega^2+\Omega^2)}$. It should be emphasized that the parameter $\kappa$ can take both
real and imaginary values depending on system parameters. In particular, we can define a diffusive ($\tau\sqrt{\omega^2+\Omega^2}\ll 1$, $\textnormal{Im}(\kappa)=0$) and
ballistic ($\tau\sqrt{\omega^2+\Omega^2}\gg 1$, $\textnormal{Re}(\kappa)=0$) limits of spin dynamics. According to these definitions, in the diffusive limit the mean free path is much smaller than the ring radius, $\ell \ll r$, and the spin precession angle per mean free path is small, $\eta \ell \ll 1$. In the opposite ballistic limit, at least one of the following inequalities should hold: $\ell \gg r$, $\eta \ell \gg 1$. Any of these two conditions for the ballistic limit results in oscillations in spin relaxation dynamics. In the diffusive limit, however, the spin polarization decays exponentially without any oscillations.
Eqs. (\ref{SrHom}) and (\ref{SzHom}) can be further simplified in the diffusive and ballistic limits. In diffusive limit, $\tau\sqrt{\omega^2+\Omega^2}\ll 1$, Eqs. (\ref{SrHom}) and (\ref{SzHom}) take the form
\begin{eqnarray}
S_r(t)=-\frac{\omega\Omega S_0}{\omega^2+\Omega^2}
\left[1- e^{-\tau (\omega^2+\Omega^2)t}\right],
\label{SrHom2}\\
S_z(t)=\frac{\omega^2 S_0}{\omega^2+\Omega^2}
\left[1+\frac{\Omega^2}{\omega^2}e^{-\tau (\omega^2+\Omega^2)t}\right].
\label{SzHom2}
\end{eqnarray}
It is clearly seen from the above expressions that in this case the relaxation of spin polarization is characterized by the time
$(\tau(\omega^2+\Omega^2))^{-1}$, which is much longer than $\tau$. In the opposite ballistic limit, when $\tau\sqrt{\omega^2+\Omega^2}\gg 1$, Eqs. (\ref{SrHom}) and (\ref{SzHom}) can be reduced to
\begin{eqnarray}
S_r(t)=-\frac{\omega\Omega S_0}{\omega^2+\Omega^2}
\left[1- e^{-\frac{t}{2\tau}}\cos(\sqrt{\omega^2+\Omega^2}t)\right],
\label{SrHom1}\\
S_z(t)=\frac{\omega^2 S_0}{\omega^2+\Omega^2}
\left[1+\frac{\Omega^2}{\omega^2}e^{-\frac{t}{2\tau}}\cos(\sqrt{\omega^2+\Omega^2}t)\right].
\label{SzHom1}
\end{eqnarray}
Consequently, the time-dependence of spin polarization components involves
an exponential decay with a time constant $2\tau$ modulated by oscillating functions.
Fig. \ref{fig2} shows an example of spin relaxation dynamics in both diffusive and ballistic limits.
\begin{figure} [tbp]
\begin{center}
\includegraphics[angle=0,width=8cm]{fig2}
\caption{(Color online) Relaxation of electron spin polarization in a ring in the (a) diffusive and (b) ballistic regimes. $S_\theta=0$.
This plot was obtained using Eqs. (\ref{SrHom}) and (\ref{SzHom}) with the following parameter values: $\Omega/\omega=-0.5$, $\tau\kappa=0.45$ (in (a)) and $\tau\kappa=i5$ (in (b)).}
\label{fig2}
\end{center}
\end{figure}
The most important feature of the solution given by Eqs. (\ref{SrHom}) and (\ref{SzHom}) is that the electron spin polarization does not decay completely to zero. At long times, it shapes into a persistent crown-like spin polarization structure (schematically presented
in Fig. \ref{fig1}). The long-time values of spin polarization components are
\begin{eqnarray}
S_r=-\frac{\omega\Omega}{\omega^2+\Omega^2}S_0
, ~~
S_z=\frac{\omega^2 }{\omega^2+\Omega^2}S_0.
\label{Szstat}
\end{eqnarray}
The above relations indicate that during the spin relaxation process the amplitude of spin polarization decreases
by the factor
\begin{eqnarray}
\frac{S}{S_0}=\frac{\sqrt{S_r^2+S_z^2}}{S_0}=\frac{\omega}{\sqrt{\omega^2+\Omega^2}}.
\label{SzAttenuation}
\end{eqnarray}
In this persistent crown-like spin polarization
structure the angle between the ring's axis and the direction of spin polarization
does not depend on the magnitude of the initial polarization $S_0$ and is equal to
\begin{eqnarray}
\tan\psi=\frac{S_r}{S_z}=-\frac{\Omega}{\omega}=-\eta r.
\label{SzAngle}
\end{eqnarray}
It follows from Eqs. (\ref{SzAttenuation}) and (\ref{SzAngle}) that the geometric parameters of persistent crown-like
spin structure depend (besides the ring radius $r$) only on the ratio of the frequency of spin-orbit precession
to the frequency of the electron rotation, $\Omega / \omega$. Moreover, we note that in the case of the opposite direction of the initial spin polarization the asymptotic value of $S_r$ changes to the opposite one. In particular, if we consider the situation shown in Fig. \ref{fig1} (realizable at a negative value of $\alpha$), then the change $S_0\rightarrow -S_0$ will result in the opposite directions of the radial and $z$ components of spin polarization. Such an asymmetry can be considered as a result of the symmetry breaking by the spin-orbit interaction.
\section{Persistent spin states}
The crown-like persistent spin polarization structure discussed in the previous section is the simplest example of persistent spin states of spin kinetic equation (\ref{Sr})-(\ref{Sz}). In this section we find the complete set of the persistent spin states of Eqs. (\ref{Sr})-(\ref{Sz}). For this purpose, we search specific solutions of these equations in the complex form
\begin{eqnarray}
S(\theta,t)=\left(
\begin{array}{c}
S_r(\theta,t) \\
S_\theta(\theta,t)\\
S_z(\theta,t) \\
\end{array}
\right)=
\left(
\begin{array}{c}
\tilde{S}_r \\
\tilde{S}_\theta\\
\tilde{S}_z \\
\end{array}
\right)e^{in\theta}e^{\lambda t},
\label{SAnsatz}
\end{eqnarray}
where $n$ is an integer number. The possible values of $\lambda$ and corresponding amplitudes $\tilde{S}_r$, $\tilde{S}_\theta$, $\tilde{S}_z$ have to be found. Substituting Eq. (\ref{SAnsatz}) into Eqs. (\ref{Sr})-(\ref{Sz}) we readily obtain a system of linear homogeneous equations.
The condition of consistency of these equations give rise for the following values of $\lambda$:
\begin{eqnarray}
\lambda_1^{\pm}&=&-\frac{1}{2\tau}\pm \sqrt{\frac{1}{4\tau^2}-n^2\omega^2},
\label{lambda1}\\
\lambda_2^{\pm}&=&-\frac{1}{2\tau}\pm \sqrt{\frac{1}{4\tau^2}-\left( n\omega+\sqrt{\omega^2+\Omega^2}\right)^2},
\label{lambda2}\\
\lambda_3^{\pm}&=&-\frac{1}{2\tau}\pm \sqrt{\frac{1}{4\tau^2}-\left(n\omega-\sqrt{\omega^2+\Omega^2}\right)^2}.
\label{lambda3}
\end{eqnarray}
Solving the linear homogeneous equations for each value of $\lambda$ (given by Eqs. (\ref{lambda1})-(\ref{lambda3})) we find corresponding amplitudes
\begin{eqnarray}
\tilde{S}_1=\left(
\begin{array}{c}
-\frac{\Omega}{\sqrt{\omega^2+\Omega^2}} \\
0 \\
\frac{\omega}{\sqrt{\omega^2+\Omega^2}} \\
\end{array}
\right),
\tilde{S}_2=\frac{1}{\sqrt{2}}\left(
\begin{array}{c}
\frac{\omega}{\sqrt{\omega^2+\Omega^2}} \\
-i \\
\frac{\Omega}{\sqrt{\omega^2+\Omega^2}} \\
\end{array}
\right),\nonumber\\
\tilde{S}_3=\frac{1}{\sqrt{2}}\left(
\begin{array}{c}
\frac{\omega}{\sqrt{\omega^2+\Omega^2}} \\
i \\
\frac{\Omega}{\sqrt{\omega^2+\Omega^2}} \\
\end{array}
\right).~~~~~~~~~~~~~~~
\label{EigenVectors}
\end{eqnarray}
It is interesting to note that the vectors given by Eqs. (\ref{EigenVectors}) do not depend on $\tau$ and $n$. They are orthogonal and normalized (in complex sense), so it is easy to expand any vector (for example, related to initial spin polarization) in this basis.
A spin state is persistent when $\lambda=0$ (this follows directly from Eq. (\ref{SAnsatz})).
It can be seen from Eqs. (\ref{lambda1})-(\ref{lambda3}) that there are three possibilities when the condition $\lambda=0$ is realized.
The first possibility is when $n=0$, in this case, $\lambda^{-}_1$ is equal to zero.
The corresponding persistent spin polarization structure is given by $S=A\tilde{S}_1$, that is
\begin{eqnarray}
S_r=-\frac{\Omega}{\sqrt{\omega^2+\Omega^2}} A,~S_\theta=0,~S_z=\frac{\omega}{\sqrt{\omega^2+\Omega^2}} A,
\label{SpinCrown}
\end{eqnarray}
where $A$ is an arbitrary constant describing the magnitude of spin polarization in this state.
It can be seen that the persistent spin state given by Eq. (\ref{SpinCrown}) is realized in the relaxation
of homogeneous spin polarization initially directed in $z$ direction (this problem was considered in the previous section, see Eq. (\ref{Szstat})).
As it is seen from Eqs. (\ref{lambda2}) and (\ref{lambda3}), two other possibilities ($\lambda^{-}_{2}=0$ or $\lambda^{-}_{3}=0$) may only be realized if $n=-\sqrt{1+\frac{\Omega^2}{\omega^2}}$ or $n=\sqrt{1+\frac{\Omega^2}{\omega^2}}$ correspondingly. Since $n$ is an integer number, these conditions are fulfilled only for certain combinations of system parameters as discussed below. By substituting Eqs. (\ref{EigenVectors}) for $\tilde{S}_2$ or $\tilde{S}_3$ into Eq. (\ref{SAnsatz}) and taking real or imaginary parts, we obtain the persistent spin polarization structures that can be presented as
\begin{eqnarray}
S_r=\frac{B}{n} \sin n(\theta-\theta_0),S_\theta=B\cos n(\theta-\theta_0),\nonumber\\
S_z= \frac{B\sqrt{n^2-1}}{n}\textnormal{sgn}\left( \Omega\right)\sin n(\theta-\theta_0),
\label{SpinExotic}
\end{eqnarray}
where $n=\sqrt{1+\frac{\Omega^2}{\omega^2}}=2,3,4,...$, $B$ and $\theta_0$ are arbitrary real constants, which determine spin polarization amplitude and angular position of the structure respectively. Fig. \ref{fig3} shows schematically the persistent spin polarization structure (\ref{SpinExotic}) for $n=2$.
\begin{figure}
\begin{center}
\includegraphics[angle=0,width=4.0cm]{fig3}
\caption{(Color online) $n=2$ stationary state (as follows from Eq. (\ref{SpinExotic}) at $\theta_0=0$).}
\label{fig3}
\end{center}
\end{figure}
It should be noted that the persistent spin polarization structures (Eq. (\ref{SpinExotic})) exist only at certain relations between the ring's radius $r$ and the strength of spin-orbit interaction $\eta$. Since $\Omega/\omega=\eta r$,
the conditions for persistent spin states are given by
\begin{eqnarray}
|\eta| r=\sqrt{n^2-1},~~ n=2,3,4,...
\label{SpinExoticParameters}
\end{eqnarray}
In addition, we note that at $r\rightarrow \infty$, $n$ becomes much greater than $1$. In this case in accordance with Eq. (\ref{SpinExoticParameters})
$n\approx \eta r\gg 1 $, and persistent spin polarization structure (\ref{SpinExotic}) becomes the usual plain spin helix \cite{Pershin05a} with wave vector $\eta$.
\section{Green function}
Linear combinations of specific solutions (\ref{SAnsatz}) (with amplitudes (\ref{EigenVectors}) and corresponding $\lambda$-s given by Eqs. (\ref{lambda1})-(\ref{lambda3})) of homogeneous linear Eqs. (\ref{Sr})-(\ref{Sz}) are also
solutions of these equations. Therefore, the general solution of the system (\ref{Sr})-(\ref{Sz}) can be presented as a sum
\begin{eqnarray}
S(\theta,t)=\sum_{n,\nu,\sigma}A_\nu^\sigma(n)\tilde{S}_\nu e^{\lambda^\sigma_\nu(n) t}e^{in\theta},
\label{GenSol1}
\end{eqnarray}
where $A_\nu^\sigma(n)$ are complex constants, $\nu=1,2,3$, $\sigma=+,-$, and $n$ runs over all integer values.
In order to find the constants $A_\nu^\sigma(n)$ we should specify the initial conditions for
spin polarization
\begin{eqnarray}
S(\theta,t)_{t=0}\equiv S(\theta,0),
\left(\frac{\partial S(\theta,t)}{\partial t}\right)_{t=0}\equiv\dot{S}(\theta,0).
\label{InitCond}
\end{eqnarray}
Using initial conditions (\ref{InitCond}) we obtain
\begin{eqnarray}
A_\nu^\pm (n)=\int_{0}^{2\pi}\frac{d\theta_0}{2\pi}e^{-in\theta_0}
\frac{\langle\tilde{S}_\nu,\dot{S}(\theta_0,0)-\lambda^\mp_\nu(n)S(\theta_0,0)\rangle}{\lambda^\pm_\nu(n)-\lambda^\mp_\nu(n)}, \;\;
\label{An}
\end{eqnarray}
where $\langle\phi,\psi\rangle=\overline{\phi}_r\psi_r+\overline{\phi}_\theta\psi_\theta+\overline{\phi}_z\psi_z$ is the inner product of amplitudes. It is easy to check that
$\langle\tilde{S}_\nu,\tilde{S}_\mu\rangle=\delta_{\nu\mu}$.
The Green function of spin kinetic equation can be obtained substituting Eq. (\ref{An}) into Eq. (\ref{GenSol1}).
The Green function can be employed to find spin polarization at any moment of time for any given initial conditions as
\begin{eqnarray}
S_{\alpha}(\theta,t)=\left[\frac{\partial}{\partial t}+\frac{1}{\tau} \right]\int_{0}^{2\pi}d\theta_0 G_{\alpha\beta}(\theta-\theta_0,t)S_\beta(\theta_0,0)\nonumber\\
+\int_{0}^{2\pi}d\theta_0 G_{\alpha\beta}(\theta-\theta_0,t)\dot{S}_\beta(\theta_0,0), \;\;\;
\label{GenSol2}
\end{eqnarray}
where $\alpha$ or $\beta$ take the values $r,\theta,z$ and summation over repeated indexes is implied.
In particular, the components of the Green function (it is actually 3x3 matrix) of the system (\ref{Sr})-(\ref{Sz}) can be written as
\begin{eqnarray}
G_{rr}(\theta,t)=\frac{\Omega^2 G_0(\theta,t)+\omega^2G_c(\theta,t)}{\omega^2+\Omega^2},
\label{Grr}
\end{eqnarray}
\begin{eqnarray}
G_{\theta r}(\theta,t)=-\frac{\omega G_s(\theta,t)}{\sqrt{\omega^2+\Omega^2}},
\label{Gtr}
\end{eqnarray}
\begin{eqnarray}
G_{zr}(\theta,t)=-\frac{\omega\Omega [G_0(\theta,t)-G_c(\theta,t)]}{\omega^2+\Omega^2},
\label{Gzr}
\end{eqnarray}
\begin{eqnarray}
G_{r\theta}(\theta,t)=-G_{\theta r}(\theta,t)
\label{Grt}
\end{eqnarray}
\begin{eqnarray}
G_{\theta \theta}(\theta,t)=G_{c}(\theta,t),
\label{Gtt}
\end{eqnarray}
\begin{eqnarray}
G_{z\theta}(\theta,t)=\frac{\Omega G_s(\theta,t)}{\sqrt{\omega^2+\Omega^2}},
\label{Gzt}
\end{eqnarray}
\begin{eqnarray}
G_{rz}(\theta,t)=G_{zr}(\theta,t)
\label{Grz}
\end{eqnarray}
\begin{eqnarray}
G_{\theta z}(\theta,t)=-G_{z\theta }(\theta,t),
\label{Gtz}
\end{eqnarray}
\begin{eqnarray}
G_{zz}(\theta,t)=\frac{\omega^2 G_0(\theta,t)+\Omega^2G_c(\theta,t)}{\omega^2+\Omega^2},
\label{Gzz}
\end{eqnarray}
where
\begin{eqnarray}
G_0(\theta,t)=\frac{e^{-\frac{t}{2\tau}}}{2\pi}\sum_{n=-\infty}^{n=+\infty}
\cos n\theta\frac{\sin\left(t\sqrt{\omega^2 n^2-1/(2\tau)^2}\right)}{\sqrt{\omega^2 n^2-1/(2\tau)^2}}, \;\;\;
\label{G0}
\end{eqnarray}
\begin{eqnarray}
G_c(\theta,t)=\frac{e^{-\frac{t}{2\tau}}}{4\pi}\sum_{n=-\infty}^{n=+\infty}
\cos n\theta
\left\{
\frac{\sin\Omega^{+}_n t}{\Omega^{+}_n}
+
\frac{\sin\Omega^{-}_n t}
{\Omega^{-}_n}
\right\},\;\;\;
\label{Gc}
\end{eqnarray}
\begin{eqnarray}
G_s(\theta,t)=-\frac{e^{-\frac{t}{2\tau}}}{4\pi}\sum_{n=-\infty}^{n=+\infty}
\sin n\theta
\left\{
\frac{\sin\Omega^{+}_n t}{\Omega^{+}_n}
-
\frac{\sin\Omega^{-}_n t}
{\Omega^{-}_n}
\right\},\;\;\;
\label{Gs}
\end{eqnarray}
\begin{eqnarray}
\Omega^{\pm}_n=\sqrt{\left(\omega n\pm\sqrt{\omega^2+\Omega^2}\right)^2-1/(2\tau)^2}.
\label{OmegaPM}
\end{eqnarray}
\begin{figure} [tbp]
\begin{center}
\includegraphics[angle=0,width=8cm]{fig4}
\caption{(Color online) Spin polarization components at $\theta=\pi /4$ induced by a delta function excitation at $\theta=0$ ($S_\beta(\theta,t=0)=0$, $\dot{S}_\beta(\theta,t=0)=\delta(\theta)\delta_{\beta,z}$, where $\delta(..)$ is the delta function and $\delta_{i,j}$ is the Kronecker delta). This plot was obtained using the parameter values $\tau \Omega=-0.1$, $\tau \omega=1$ (in (a)), and $\tau \omega=0.1$ (in (b)).}
\label{fig4}
\end{center}
\end{figure}
The above equations can be further simplified. In fact, the sums appearing in Eqs. (\ref{G0})-(\ref{Gs}) can be calculated in the following way.
First of all, we employ the well-known formula from the Bessel function theory
\begin{eqnarray}
\frac{\sin \left(t\sqrt{\omega^2-q^2} \right)}{\sqrt{\omega^2-q^2}}= \nonumber \\
\frac{1}{2}
\int_{-\infty}^{+\infty}d\xi e^{i\omega\xi}\Phi(t-|\xi|)I_0\left( q\sqrt{t^2-\xi^2}\right)
\label{Formula},
\end{eqnarray}
where $\Phi(t)$ is the Heaviside step function, and $I_0(t)$ is the modified Bessel function
of zero order. Substituting Eq. (\ref{Formula}) into Eqs. (\ref{G0})-(\ref{Gs}) we perform summation in
these equations taking into account that \cite{VladimirovEqMathPhys}
\begin{eqnarray}
\sum_{n=-\infty}^{n=+\infty}e^{in\xi}=2\pi\sum_{n=-\infty}^{n=+\infty}\delta(\xi-2\pi n).
\label{Formula2}
\end{eqnarray}
Then, the integrals in Eqs. (\ref{G0})-(\ref{Gs}) modified by Eq. (\ref{Formula}) are trivially evaluated yielding the following results
\begin{eqnarray}
G_0(\theta,t)=\frac{e^{-\frac{t}{2\tau}}}{2\omega}\sum_{n=-\infty}^{n=+\infty}
I_0
\left(
\frac{\sqrt{\omega^2t^2-(\theta-2\pi n)^2}}{2\omega\tau}
\right)\nonumber\\
\times\Phi(\omega t-|\theta-2\pi n|),~
\label{G02}
\end{eqnarray}
\begin{eqnarray}
G_c(\theta,t)=\frac{e^{-\frac{t}{2\tau}}}{2\omega}\sum_{n=-\infty}^{n=+\infty}
I_0
\left(
\frac{\sqrt{\omega^2t^2-(\theta-2\pi n)^2}}{2\omega\tau}
\right)\nonumber\\
\times\cos\left[
\sqrt{1+\frac{\Omega^2}{\omega^2}}\left(\theta-2\pi n\right)\right]
\Phi(\omega t-|\theta-2\pi n|),
\label{Gc2}
\end{eqnarray}
\begin{eqnarray}
G_s(\theta,t)=\frac{e^{-\frac{t}{2\tau}}}{2\omega}\sum_{n=-\infty}^{n=+\infty}
I_0
\left(
\frac{\sqrt{\omega^2t^2-(\theta-2\pi n)^2}}{2\omega\tau}
\right)\nonumber\\
\times\sin\left[
\sqrt{1+\frac{\Omega^2}{\omega^2}}\left(\theta-2\pi n\right)\right]
\Phi(\omega t-|\theta-2\pi n|).
\label{Gs2}
\end{eqnarray}
In fact, all sums in Eqs. (\ref{G02})-(\ref{Gs2}) are finite sums because of the Heaviside functions $\Phi(\omega t-|\theta-2\pi n|)$.
For any given values of time $t$ and angle $\theta$ only the terms with $n$ in the range $(\theta-\omega t)/(2\pi)<n<(\theta+\omega t)/(2\pi)$
give non-zero contribution to the R.H.S. of Eqs. (\ref{G02})-(\ref{Gs2}).
Fig. \ref{fig4} shows an example of application of the Green function formalism to calculations of spin polarization dynamics.
Specifically, we plot spin polarization components at $\theta=\pi /4$ induced by spin polarization excitation at $\theta=0$. Such an excitation
induces clockwise and counterclockwise propagating waves of spin polarization that finally evolve into a steady homogeneous spin polarization.
Each time when clockwise and counterclockwise propagating waves of spin polarization reach $\theta=\pi /4$, the spin polarization at this point changes in steps (see Fig. \ref{fig4}(a)). The step amplitudes, however, decrease in time because of the diffusive component in spin transport. This type of behavior resembles a multiple echo.
In addition, in the diffusive limit shown \ref{fig4}(b) we did not observe clear steps in spin polarization. At long times, $S_\theta \rightarrow 0$ and the crown-like spin polarization configuration is formed.
\section{Conclusions}
Spin relaxation in confined structures exhibits some remarkable properties and behavior that are not found in infinite 2D systems.
In this paper we have studied the electron spin relaxation in the ring. It has been found that the homogeneous spin polarization along the ring axis
transforms into a persistent crown-like spin structure. Moreover, a family of persistent spin states in the ring has been identified. We have also derived the Green function
of spin kinetic equation and used it to investigate the propagation of a point spin excitation.
|
\subsection*{1. Introduction}
According to Bohmian interpretation of quantum mechanics (see \cite{Bohm1} and \cite{Bohm2}), the wave and the particle co-exist as two separate substances. The wave evolves according to Schrodinger's equation, while particle moves according to \emph{guidance equation} given by
\beq \frac{d \vec{x}}{dt} = \frac{1}{m} \; \vec{\nabla} \; Im \; ln \; \psi \eeq
Suppose at the beginning of the universe, a "biased coin" was tossed in order to determine where to place a particle in space. If the bias of a coin happened to be $\rho = \vert \psi \vert^2$, then the "probability current" ($\vec{J}$) associated with \emph{classical} probability ($\rho$) of finding a particle will coincide with the "probability current" ($\vec{j}$) associated with $\vert \psi \vert^2$ that we obtain from Schrodinger's equation. For this reason $\rho$ and $\vert \psi \vert^2$ will continue to coincide at the time $t+ dt$. Then, for the same reason, they will coincide at $t+2dt$, and so forth. This is the key argument in favor of Born's rule arising out of Bohmian model.
According to the Bohmian theory, the collapse of wave function is a result of wave function splitting into non-overlapping branches. In such scenario, a particle will "happen" to occupy only \emph{one} of them, and the zero-probability regions will prevent it from going back to any other branches. These branches are interpreted as different outcomes of a measurement. In this language, the "collapse" of the wave function is equivalent to these branches disappearing. According to the theory, they do not disappear and, therefore, no true collapse occurs. But, due to their lack of overlap, they have no effect on a Bohmian particle. Thus, the particle "thinks" they disappeared, which results in \emph{appearance} of collapse of wave function. This phenomenon is commonly referred to as "effective collapse".
However, the reliance on the claim of "lack of overlap" raises several problems. First of all, the fact that branches do not overlap was disputed by Leggett. According to his argument, the only thing we know is the emergence of sum rule, but the latter implies the \emph{lack} of evidence of the \emph{absence} of a "split" which is \emph{not} the same as the evidence of its presence (see \cite{Leggett1} and \cite{Leggett2}). Secondly, even if we believe that the splitting of the branches occurs, the particle can "cross" that "gap" once the theory is non-local in configuration space.
Such "non-local" theories have been recently proposed in \cite{Epstein} and \cite{Sverdlov}. Both \cite{Epstein} (Chapter 3) and \cite{Sverdlov}, in fact, acknowledge the "traveling between universes" that would occur as a result of this non-locality (although in \cite{Sverdlov} rather artificial mechanism was presented in "getting rid" of these unwanted universes). Finally, if one hopes to introduce gravity into the the theory (which is beyond the scope of this paper) the non-linear nature of gravity might imply non-zero gravitational interaction between these branches.
As Leggett pointed out (see \cite{Leggett1} and \cite{Leggett2}) the above problems are completely absent in GRW theory. The GRW theory postulates "real" collapse which is is \emph{exclusively} based on the distance in configuration space (and, therefore, no appeal to the shape of the curve, including the "branching" is made). The "real" collapse also would get rid of any unwanted branches \emph{if} such were created (although, of course, if they are not created to begin with, this is "even better"), so the issues such as "particle crossing the gap between the branches" doesn't arise either.
The jist of the theory is that a wave function is being periodically multiplied by Gaussians around random points at random times; these events are called "hits". These Gaussians are very wide, and occur very rarely. Thus, their effects on few particle systems are negligible. In multi-particle case, however, the entanglement implies that any particle is being affected by "hit" performed on any other particle. This makes the number of "hits" it experiences far more frequent, which makes the degree of localization significant. The same is true for external particle having short term interaction with the multiparticle system, such as electron interacting with the screen.
However, GRW theory is stochastic in its original form, which is in conflict with the goal of some physicists (including the author of this paper) to come up with deterministic theory. There have been "continuous spontaneous localization models" (see \cite{continuous}) that postulated extra classical fields that "trigger" the above described localization behavior. Thus, instead of random "hits", the "multiplication by Gaussian" is continuous and its (very slow) rate varies according to the behavior of the above mentioned classical field. This enables the theory to be deterministic.
In this paper, we will propose a different alternative of imposing determinism upon GRW model. Instead of claiming that the "classical trigger" involves these fields, I will claim that the Bohmian particle, itself, triggers a "continuous localization" around itself, as a center. In other words, the evolution of wave function is given by
\beq \frac{\partial \psi}{\partial t} = i H \psi - a (\vec{x} - \vec{x}_B)^2 \eeq
where $\vec{x}_B$ is a location of "beable particle". Upon integration over finite time interval, the above equation also produces the multiplication by Gaussian. The key difference between this dynamics and GRW model is that in our case we no longer have discrete hits, which were the only sources of randomness.
There are different ways of summarizing what we are proposing to do. On the one hand, we can say that we plan to use GRW as a "supplement" to Bohmian dynamics that "gets rid" of unwanted parts of wave function (and, therefore, renders the argument about "splitting into branches" unnecessary). At the same time, it can also be viewed as Bohm supplementing GRW in a sense that the latter no longer needs to postulate extra classical fields as "sources" of localization. In the former case we claim that GRW is "a little addition" to the "main Bohmian context" while in the latter case we claim just the opposite. We can also think of it in "unbiased" way and say we are simply "merging" Bohm and GRW into a single "hybrid" between the two. This is the preferred view of the author.
However, the author of this paper happens to have some other, unrelated, objections against GRW. The first objection is that the "multiplication by Gaussian" goes against momentum conservation. The second, and unrelated, subjection is that the "smooth dynamics" summarized in the above equation can be re-phrased in terms of "imaginary Hamiltonian" $H = -iax^2$. While neither of these issues affect the validity of the main point of the paper, we will still attempt to address those in Chapters 6 and 7, respectively, for the sake of completeness of presentation.
\subsection*{2. The problems created by particle's influence on the wave}
One key implication of coupling Bohm to GRW is that the wave-particle interaction happens \emph{both} ways. This should be contrasted with purely Bohmian case when the wave influences a particle and the particle has no effect on the wave. The \emph{lack} of the influence of a particle on the wave is precisely the reason why unwanted branches continue to exist. In our case, the \emph{presence} of that influence is claimed to get rid of these branches.
However, once we allow a particle to influence a wave, one serious question arises. In order to write down Born's rule, we have to make two key assumptions. One is that we \emph{do} know the exact value of $\psi$ and the other is that we do \emph{not} know the exact location of a particle. If we did not know the value of $\vert \psi \rangle$, we would not be able to use it in Born's rule. On the other hand, if we knew the location of a particle, the "probability distribution" would have been a $\delta$-function, in violation of Born's rule. This raises a question: how can we "know" $\psi$ despite the fact that it is being influenced by $\vec{x}$ which we do \emph{not} know?
This question can also be asked in another way. Suppose we don't know $\psi$ at any other time, but we are "told" what $\psi$ is at time $t=t_0$. We are now trying to estimate the probability of finding a particle at different locations in $\vec{x}$ at this particular time. Now, \emph{if} a particle happens to be located at $\vec{x} = \vec{x}_0$, then we can do "reverse dynamics" and say that, at time $t- \delta t$, the wave function was $\psi +iH \psi \delta t - G(\vec{x}_0) \psi$, where $G(\vec{x}_0)$ is an "additional influence" the particle exerts onto a wave function.
Now, if we again look at the "forward dynamics", in order for particle to "end up" being at $\vec{x}_0$ we have to \emph{first} produce $\psi +iH \psi \delta t - G(\vec{x}_0) \psi$ and \emph{if} a particle \emph{happens to be} at $\vec{x}_0$, then this would ultimately produce the situation we are looking for after the interval $\delta t$ passes. This means that if it happens that, for the time $t_0 - \delta t$, the state $\psi +iH \psi \delta t - G(\vec{x}_1) \psi$ is "more likely" to be produced than $\psi +iH \psi \delta t - G(\vec{x}_2) \psi$, then at time $t_0$, the particle is "more likely" to be found at $\vec{x}_1$ than at $\vec{x}_2$ at the "ideal" situation that $\vert \psi (\vec{x}_1) \vert ^2 = \vert \psi (\vec{x}_2)^2$.
Of course, it is \emph{possible} that the probability of producing $\psi +iH \psi \delta t - G(\vec{x}_0) \psi$ at time $t_0 - \delta t$ is independent of the choice of $\vec{x}_0$, but we do not \emph{know} that. Finding out whether or not that is the case is quite difficult since the probability of production of $\psi +iH \psi \delta t - G(\vec{x}_0) \psi$ depends on the situation \emph{before} $t_0 - \delta t$ which means that we can not use the "back dynamics" from time $t_0$ to assist ourselves. On the other hand, the "forward" dynamics for the times $t < t_0 - \delta t$ depends on the entire quantum mechanics (or, in non-relativistic case, quantum field theory) to which $\psi$ is subject to, which makes the proof of any kind of assertion extremely difficult.
One way to illustrate the difficulty is to consider a problem of a kid and a candy. We are told about the amount of candies in different locations and we are asked to find out probabilities at which a kid is to be found at each of these places. On the one hand, one can argue that kid is more likely to be found where there is more candy, since he is "drawn" to the candy. On the other hand, we can argue that he is more likely to be found where there is \emph{less} candy since he would eat the candy (and, thereby, decrease its number) at whatever location he occupies.
In order to answer this question we need to have an \emph{additional knowledge} about the way candy is distributed. If we happened to know that for some reasons \emph{independent of the child} candies are typically distributed evenly, we would conclude that a kid is to be found where there is \emph{less} candy. After all, the only conceivable reason there is "less candy" to begin with is that he ate them. On the other hand, if we are told that, for some reason, the distribution of candies is \emph{highly uneven} then the kid would be more likely to be found where there is \emph{more} candy. After all, the "unevenness" of candy distribution is so high that a kid can't possibly eat enough candies to "undo" it. Thus, he would arrive at the place with "more candies", eat some, and there would \emph{still} be "more candies than average" after he is done eating.
Furthermore, even if we do know that candies are distributed evenly, the situation might not be as simple. For example, if the "even distribution of candies" happens very rarely, then the kid would eat candies from all the rooms, resulting in their uneven distribution; afterwards, he would "move" to the room where more candies are "left". On the other hand, if the candies are being "evened out" by outside sources quite often, he won't have time to eat candies from all the rooms (provided that we assume that there are a lot of rooms) and, therefore, the room he would occupy would be one of the "few" rooms he "had time" to reach and, therefore, would have "fewer candies". But, at the same time, he might have still had time to reach more than one room and, between the rooms that he could reach, he could "pick" the one where he ate less candies since more candies are "left" for him to eat. From the point of view of this picture, the probability of finding a kid in \emph{either} a room with a lot of candies \emph{or} a room with very few ones is very small, but for very different reasons. The probability "peaks" around the rooms with "fewer but not too few" candies, making the theory "non-linear".
We should notice that the non-linearity of the above paragraphs was produced within "black or white" context of "even distribution of candies". The situation becomes even more complicated if we allow the distribution of candies to be uneven, but, at the same time, not uneven enough to "qualify" for the "other extreme" discussed earlier. Thus, we could not blindly say that the room with more candies attracts a kid since it is \emph{possible} he eats more candies than the "standard deviation" of their distribution; but, at the same time, such criteria might "partly" hold since standard deviation is quite large and a kid \emph{might well} eat fewer candies than that.
Going back to physics, the "kid" is analogous to a "particle", a "candy" is analogous to the wave, "kid wanting to go where there is more candy" is analogous to "wave guiding the particle through Pilot Wave model" and "kid eating the candy" is analogous to the particle influencing the behavior of a wave through GRW-type phenomenon centered around its own location. The "way in which candies are distributed" is analogous to the knowledge of quantum field theory. Thus, the fact that we need to know the way candies are distributed in order to decide how to estimate the probabilities of finding a kid implies that we need to know quantum field theory in order to estimate the probability of finding the particle. This is in sharp context with standard quantum mechanics (or field theory) where, in principle, we can estimate the probability of finding a particle simply by "being told" about the probability amplitude, even if we, ourselves, do not know quantum mechanics and can not reproduce it.
Apart from mathematical difficulties involved, there is also a conceptual one. If we pretend for a moment that a particle doesn't exist and, therefore, wave evolves without its influence, it might still be "reasonable" to say that some shapes of waves are "more likely" to be produced than other shapes. But, at the same time, the unitary evolution is completely deterministic. Thus, in order to talk about different probabilities of various outcomes, we need to be uncertain about initial conditions. Technically, we could describe the uncertainty of wave function through density matrix. But additional notation is not going to answer our conceptual question: just what "makes" one kind of wave function "more likely" than the other kind? These questions is one of the main things I am attempting to tackle in this paper.
\subsection*{3. The "bigger" space consisting of objects of the form $(\vert \psi \rangle, \vec{x}$) }
I propose to address the above questions in the following way. In terms of kid and a candy I would combine them into a single object, $({\rm kid}, {\rm candy})$ that undergoes certain dynamics. Here, by "candy" we mean the \emph{set} of numbers that list all the rooms and list the amounts of candies in every single one of them, and by "kid" I mean a single "number" that tells which room a kid occupies. Then, I would be able to \emph{first} find probability density distribution around any given point $({\rm kid}, {\rm candy})$ and then, later, find a \emph{conditional probability} of kid being in a certain place for a specific distribution of candies. If "kid" is viewed as "$x$-coordinate" and "candy" as "$y$-coordinate", it makes perfect sense to ask about the conditional probability of finding $x$ (that is, a kid) for afore-specified $y$ (that is, candy distribution). This will allow us to use "candy" as a way of determining a "probability" of finding a kid, just like $\psi$ determines a probability of finding a particle.
By the similar logic, I will combine a state $\vert \psi \rangle$ and a particle configuration $\vec{x}$ into a single object, $(\vert \psi \rangle, \vec{x})$ which undergoes random walk in that "bigger space". Thus, $(\vert \psi \rangle, \vec{x})$ will be a \emph{single point} as far as the "bigger space" is concerned. That "bigger space", however, will have $\infty^{3N} + 3N$ dimensions. In particular, we need $3$ dimensions in order to describe one point particle, and, therefore, $3N$ dimensions to describe configuration of $N$ of them. Furthermore, we need $\infty$ dimensions to describe wave function over one dimensional space and, therefore, $\infty^{3N}$ dimensions in order to describe wave function over $3N$ dimensions. But if we combine all of these into a single $\infty^{3N} + 3N$-dimensional space, we would still be able to think of $(\vert \psi \rangle, \vec{x})$ as a single "point" undergoing random walk in that space.
In the standard situation where only $\vec{x}$ was undergoing "random walk" there was an "outside influence" by $\psi$. Thus, there was a \emph{dynamic equilibrium} $\rho = \vert \psi \vert^2$, and both sides were time dependent. In our new situation, however, there is nothing "outside" $(\vert \psi \rangle, \vec{x})$. Thus, the equilibrium is static one. In other words, we \emph{first} come up with evolution equation,
\beq \frac{d (\vert \psi \rangle, \vec{x})}{dt} = u (\vert \psi \rangle, \vec{x}) \eeq
and then determine the time derivative of "classical" probability $\rho (\vert \psi \rangle, \vec{x})$ based on the above dynamics. Finally, we determine what $\rho$ is based on the demand that its time derivative is zero. After that, for any given $\vert \psi \rangle$ I will determine a \emph{conditional probability} of finding $\vec{x}$ at any given point based on
\beq \frac{\rho (\vert \psi \rangle, \vec{x})}{\int d^{3N} y \; \rho (\vert \psi \rangle, \vec{y})} \eeq
where we have taken the $\vert \psi \rangle = \vert \psi_0 \rangle$ "section" of our space, which answers the question of "how do we know $\psi$". The burden of the theory is to prove that the above expression coincides with $\vert \langle x \vert \psi \rangle \vert^2$. This is now a mathematically well defined statement, which one can reasonably expect to answer either affirmatively or otherwise.
Now, in order to obtain the dynamics out of the above "static" equilibrium, one has to realize that the above probabilities are strictly due to our "lack of knowledge". Thus, in principle we can "find out" some information about the time $t=t_0$ which would "allow us" to predict what would happen at some future time $t>t_0$. In particular, I claim that, if we have we have "found out" that, at the time $t=t_0$, the configuration of particles is $\vec{x}_0$, we would be to find out a \emph{new} probability density, $\sigma (\vert \psi \rangle, \vec{x} ; t)$ for some "future time" $t$ based on this information. While $\sigma$ is time-dependent, we would have to use time-independent $\rho$ in our calculation, which is the "static" aspect of the theory.
Let us ask ourselves what is the probability that $(\vert \psi \rangle, \vec{x})$ will happen to reside within the set $S$ at a time $t > t_0$ and then we can "shrink" $S$ around the point we are interested in. Let us denote the evolution under $u$ as $e^{u(t-t_0)}$. Furthermore, let us define $a ((\vert \psi \rangle, \vec{x} \rangle), S)$ to be "yes or no" function that is equal to 1 if $(\vert \psi \rangle, \vec{x})$ is an element of $S$ and zero otherwise. Then the probability of finding $(\vert \psi \rangle, \vec{x})$ inside $S$ at time $t$ is given by
\beq \int_{S} [{\cal D} \psi ] d^{3N} x \; \sigma (\vert \psi \rangle, \vec{x}, t) = \int [{\cal D} \psi ] \rho (\vert \psi \rangle, \vec{x} = \vec{x}_0) a (e^{u(t-t_0)} (\vert \psi \rangle, \vec{x} = \vec{x}_0); S) \eeq
Likewise, if we, instead, "find out" the value of $\vert \psi \rangle$ at a time $t_0$ \emph{as opposed to} the configuration of particles, we can obtain the probability distribution at time $t >t_0$ through
\beq \int_{S} [{\cal D} \psi ] d^{3N} x \; \sigma (\vert \psi \rangle, \vec{x}, t) = \int d^{3N} x \; \rho (\vert \psi \rangle)= \vert \psi_0 \rangle, \vec{x}) a (e^{u(t-t_0)} (\vert \psi \rangle )) = \vert \psi_0 \rangle , \vec{x}); S) \eeq
The choice "which" we "find out" is based on the philosophy of a reader. If a reader believes that we "see" point particles and "use" that information to "infer" the wave function, he would use the former equation. If he believes we "see" the wave function while point particles are the things that guide its evolution which we don't directly see, he will use the latter equation. Our theory works equally well in both cases.
\subsection*{4. Repeated observations}
As we have seen in the previous chapter, the probability $\rho$ is static, while the probability $\sigma$ is dynamic. Its time-dependence depends on our ability to "find out" an additional information at any given time. We have, however, limited the "additional information" we are allowed to find out. In particular we considered an option of "looking at $\vec{x}$" without being allowed to "look at $\vert \psi \rangle$" or visa versa. While, as we said, the choice between these two options is not important, the \emph{existence} of well defined limitation is. After all, if we could measure the exact value of both $\vert \psi \rangle$ and $\vec{x}$ then the determinism would imply the exact prediction for future; in other words, $\sigma$ would be $\delta$-function, which is not what we want.
Now, if we \emph{do} have some limitations on what we are allowed to observe, What \emph{is} important, however, is that we have to be "consistent" in the kind and amount of information we are allowed to "find out". After all, that is what the "consistency" of $\sigma$ depends upon. This raises an important question: does the performance of repeated observations allow us to "infer" extra information and thus ruin this consistency? The answer to this question is that the only way to access past observations is to "look at" our \emph{current} memory. Since our brain is a physical system, our current memory is recorded both in \emph{current} $\vert \psi \rangle$ and $\vec{x}$. Which still reduces to only one observation.
Now, the idea that in the past we never observed anything and just have a "false memory" that we did is, of course, very unpleasant. A way to go around it is as follows. Yes, we \emph{do} make several different observations over time. But every time we make an observation we "forget" what we have seen once a small time interval $\delta t$ passes by. Thus, at a time $t_1 + \delta t$ we no longer remember anything we have seen at $t_1$. Then, at time $t=t_2$ we make another observation. This observation included the configuration of particles in our brain. From the configuration of particles in the brain we can \emph{infer} some aspects of what we have observed at $t=t_1$. But, at the same time, we \emph{still} don't remember that observation; we are simply \emph{inferring it} from observing our brain at $t= t_2$.
It is interesting to note that if we \emph{did} have direct access to our observation at $t=t_1$ then we would indeed get additional clues about the system that we can not obtain from the single observation. At the same time, by observing our brain at $t=t_2$ we can \emph{not} get these clues since, by definition, this is still a "single observation". This implies a fundamental limitation on the amount of information that can be recorded in our brain. In other words, it is possible that at $t_1$ we "observe" a lot more stuff than our brain retains.
This limitation of the brain is due to the fact that a given configuration of particles $\vec{x}_2$ (which includes our brain) at $t_2$ can be reached from many different particle configurations $\vec{x}_1$ at $t_1$. The lack of one to one correspondence is the key to lack of perfect memory. Now, if we could access both $\vert \psi \rangle$ \emph{and} $\vec{x}$, then there would, in fact, be one to one correspondence since the dynamics of $(\vert \psi \rangle, \vec{x})$ is deterministic. There is no fundamental "quantum mechanical" reason for us not to be able to do that. We have simply set up a "classical" rule that we can't, since that is the only source of (classically) "probabilistic" nature of $\sigma$.
Now, if we define our brain through $(\vert \psi \rangle, \vec{x})$, then it would, indeed have unlimited capacity. But, in this case, we are "allowed" only an access to a "part of" our brain, and thus only part of its memory. In this new language the picture is as follows. "We" ourselves do not consist of anything physical and, therefore, we have zero memory. Our brain, consisting of \emph{both} particles ($\vec{x}$) \emph{and} waves ($\vert \psi \rangle$) has absolutely perfect memory. However, we only have access to \emph{part} of our brain (for example, that part consists of \emph{complete} information about $\vec{x}$ and \emph{no} information about $\vert\psi \rangle$). Thus, by making repeated observations of our brain we are lead to believe that we have "imperfect memory".
Let us now go back to the question we asked earlier: if we have a dynamic picture how $\rho$ can be truly static? Let us imagine the following scenario. We first perform an observation at $t=t_1$. We then predict the outcome for the time $t=t_2$. We then "record" both observation and the predicted outcome in our brain. At time $t=t_1 + \delta t$ we completely forget both the observation and predicted outcome. Then at a time $t=t_2$ we make another observation, in order to make a prediction for $t=t_3$.
Now, at a time $t_2 - \delta t$ we have not made the observation \emph{yet}. This means that, at a time $t_2 - \delta t$ we were not looking at \emph{any} part of the universe, including our brain. This means that we have "forgotten" our observation at $t_1$. At the same time, however, we \emph{can} "remember" the "static equilibrium" $\rho (\vert \psi \rangle, \vec{x})$ without "looking" at our brain. Thus, at time $t_2 - \delta t$ we will write down $\rho$; then, at time $t_2$ we will "look" at everything (including our brain) and record $\vec{x} (t_2)$. And then we will use both $\vec{x} (t_2)$ as well as $\rho$ in order to make predictions for $t_3$. Since at a time $t_2 - \delta t$ we had no memory of the observation at $t_1$, we have "written down" the same $\rho$ as we did at $t_1 - \delta t$. That is the ultimate reason why $\rho$ is time independent.
Now, since our observation at $t_2$ includes the observation of our brain that has a record of what happened at $t_1$, the above argument can be stated as follows. The probability distribution $\rho$ is an "a priori knowledge" that is "innate" to us independently of our memories (namely, the "knowledge of physics"). Then when we access our brain at $t_2$, we use our "a priori knowledge" in order to evaluate what we see (in particular, our "physics knowledge" would tell us how likely a wave would look a certain way for a given distribution of particles that we "see"). Thus, while the information we access is time dependent, the "physics knowledge" we have is not. The "static" probability distribution $\rho$ is precisely that knowledge.
Now, if we could access both $\vec{x}$ \emph{and} $\vert \psi \rangle$, we would no longer need $\rho$. The only "physics knowledge" we would need is a \emph{deterministic} dynamics that would predict the future $(\vert \psi \rangle, \vec{x})$ with absolute certainty. The reason we need $\rho$ is precisely because we are not allowed to access $\vert \psi \rangle$, and we can \emph{only} access $\vec{x}$. Thus, $\rho$ is our "guide" on how to "infer" some information about $\vert \psi \rangle$ from $\vec{x}$ (even though that information is probabilistic). In terms of our brain, $\rho$ serves as a guide for us to "guess" the "rest of our brain" (described by $\vert \psi \rangle$) based on our reading of its part described by $\vec{x}$ . This means that we \emph{need} $\rho$ in order to access our memory. Thus, in order to avoid circular arguments, $\rho$ better be memory-independent, which is why it is time-independent.
\subsection*{5. The "dynamic picture" predicted from "static equilibrium"}
In order to satisfy ourselves regarding the "static picture", let us "switch off" collapse mechanism (or, in other words, $\vec{x}$-dependence of $\vert \psi \rangle$) for a while and attempt to reproduce the standard "dynamic" Bohmian argument based on our "static" one. First, let us expand the divergence equation for "static equilibrium" in the following way:
\beq - \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi \rangle, \vec{x})} = \vec{\nabla}_x (\rho \vec{v}_x) + \vert v_{\psi} \rangle \cdot \vec{\nabla}_{\psi} \rho + \rho \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle \eeq
Let us now imagine that there are many different particles in $(\vert \psi \rangle, \vec{x})$ space moving according to the same "deterministic law", but they have different initial conditions, and they do not interact with each other. Suppose we are "sitting" on one of these particles, and measuring the "density" of the other particles in its neighborhood. In this case, the latter will evolve as
\beq \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi (t) \rangle, \vec{x} (t))}= \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi \rangle, \vec{x})} + \vert v_{\psi} \rangle \cdot \vec{\nabla}_{\psi} \rho \eeq
which means that the "static" equation we have written is equivalent to
\beq \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi (t) \rangle, \vec{x}(t))} = \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi \rangle, \vec{x})} - \vec{\nabla}_x \cdot (\rho \vec{v}_x) - \rho \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle \eeq
Furthermore, we recall that "complex divergence" is given by
\beq div \; F \; = \; 2 \; Re \; \Big( \frac{\partial F}{\partial z} \Big) \eeq
If we generalize it to multidimensional case (and further extrapolate it to uncountable dimensionality of $\vert \psi \rangle$ and $\vert v_{\psi} \rangle$) then we will find that the presence of a coefficient $i$ in the unitary evolution equation,
\beq \vert v \rangle = -iH \vert \psi \rangle \eeq
implies that
\beq \vec{\nabla} \cdot \vert v_{\psi} \rangle = 0 \eeq
as long as $H$ is Hermitian. This means that we can "throw away" the last term from the right hand side of the equation for $d \rho /dt$. Thus, we obtain
\beq \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi (t) \rangle, \vec{x}(t))} = \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi \rangle, \vec{x})} - \vec{\nabla}_x \cdot (\rho \vec{v}_x) \eeq
Now, if we substitute a \emph{static equilibrium equation},
\beq \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi \rangle, \vec{x})} = 0 \eeq
which we are "advertising", we then obtain
\beq \frac{\partial \rho}{\partial t} \Big\vert_{(\vert \psi (t) \rangle, \vec{x}(t))} = - \vec{\nabla}_x \cdot (\rho \vec{v}_x), \eeq
which exactly coincides with \emph{dynamic equilibrium} equation we are used to in the context of random walk in $\vec{x}$ \emph{alone}. From this point on the argument is a carbon copy of the standard Bohmian one: if we cleverly define $\vec{v}_x$ in such a way that
\beq \frac{\partial \vert \psi \vert^2}{\partial t} = - \vec{\nabla}_x \cdot (\vert \psi \vert^2 \vec{v}_x) \eeq
is satisfied, then $\rho = \vert \psi \vert^2$ is an "equilibrium point" of the theory.
We notice, however, that the above argument crucially depends on the presence of $i$ in the evolution equation. In particular, we have appealed to the presence of $i$ when we "got rid" of unwanted $\vec{\nabla} \cdot \vec{v}$ term, by identifying it with zero. There are two issues that arise from this. First of all, if we introduce a collapse mechanism in a form of multiplication by Gaussian, this would \emph{not} have the "$i$" that the ordinary unitary evolution equation does. Secondly, and perhaps even more importantly, the "logic" behind \emph{either} dynamic \emph{or} static equilibrium argument is independent of our knowledge of "physics", which includes our knowledge of unitary evolution equation. Thus, \emph{if} no "logical mistakes" were made in either argument, their results should match, even in the case of "wrong physics" that lacks unitarity. The fact that such is not the case seems to suggest that one of these arguments is "wrong". If so, we have to find out which one.
One source of mismatch between the two probabilities is that in a static case we are considering \emph{overall probability}, while in the "dynamic case" we are considering a \emph{conditional probability}, $\rho / \sigma$ where
\beq \sigma (\vert \psi \rangle) = \int d^{3N} x \; \rho (\vert \psi \rangle, \vec{x}) \eeq
Let us, therefore, rewrite our "static equation" for the conditional probabilities, as well. We recall that, right \emph{before} we used "zero divergence" our equation for $\rho$ took the form
\beq \frac{\partial \rho}{\partial t} = - \vec{\nabla}_x \cdot (\rho \vec{v}_x) - \rho \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle, \eeq
Here, we have dropped one of the time derivative terms, because we are assuming from now on that static equilibrium has already been reached. Now, if we "integrate" it over $\vec{x}$, we obtain
\beq \frac{\partial \sigma}{\partial t} = \int d^{3N} x \; (- \vec{\nabla}_x \cdot (\rho \vec{v}_x) - \rho \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle ) \eeq
Now, by Gauss' theorem, the first term on the right hand side is equal to the surface integral of $\rho \vec{v}_x$ over a "very large" sphere that encircles "all of the $\vec{x}$-space". Thus, if we assume a "boundary conditions" that $\vec{v}_x =0$ at $\vec{x} = \infty$, then the first term drops out, and the equation becomes
\beq \frac{\partial \sigma}{\partial t} = - \int d^{3N} x \; \rho \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle \eeq
Let us now write the expression for the time derivative of "conditional probability" $\rho / \sigma$:
\beq \frac{\partial}{\partial t} \frac{\rho}{\sigma} = \frac{1}{\sigma} \frac{\partial \rho}{\partial t} - \frac{\rho}{\sigma^2} \frac{\partial \sigma}{\partial t} \eeq
Now, while one of the two terms of $\partial \sigma / \partial t$ was killed out by the integral, that term is not zero \emph{outside of integration}. Thus, the expression for $\partial \rho / \partial t$ has both terms. If we now substitute "two terms" in place of $\partial \rho / \partial t$, and "one term" in place of $\partial \sigma / \partial t$, we obtain
\beq \frac{\partial}{\partial t} \frac{\rho}{\sigma} = \frac{1}{\sigma} (- \vec{\nabla}_x \cdot (\rho \vec{v}_x) - \rho \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle) - \frac{\rho}{\sigma^2} \Big(- \int d^{3N} x \; \rho \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle \Big) \eeq
In light of the fact that $\sigma$ is defined as an integral over $\vec{x}$, it is a function of $\psi$ alone and, therefore, we are free to move $\sigma$ in and out of differentiation and integration over $\vec{x}$. Thus, we can rewrite the above equation as follows
\beq \frac{\partial}{\partial t} \frac{\rho}{\sigma} = - \vec{\nabla}_x \cdot \Big(\frac{\rho}{\sigma} \vec{v}_x \Big) - \frac{\rho}{\sigma} \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle + \frac{\rho}{\sigma} \Big( \int d^{3N} x \; \frac{\rho}{\sigma} \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle \Big) \eeq
which can be further rewritten as
\beq \frac{\partial}{\partial t} \frac{\rho}{\sigma} = - \vec{\nabla}_x \cdot \Big(\frac{\rho}{\sigma} \vec{v}_x \Big) - \frac{\rho}{\sigma} \Big( \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle - \int d^{3N} x \; \frac{\rho}{\sigma} \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle \Big) \eeq
The last two terms on the right hand side represent the difference between the value of $\vec{\nabla}_{\psi} \vert v_{\psi} \rangle$ for \emph{a given $\vec{x}$} and its "average" over all $\vec{x}$ (but with fixed $\psi$). In order to "set" this difference to zero we have to make sure that $\vert v_{\psi} \rangle$ is $\vec{x}$-independent. Remembering that $\vec{x}$ represents position of a "particle", while $\vert v_{\psi}$ represents velocity of a "wave", this is equivalent to saying that the behavior of a wave is independent of the way the particle influences it.
Now, we have been saying earlier that \emph{any} non-unitary influence on the wave would cause a problem, not just the one coming from the particle. The way particle-independent non-unitary influences are "dealt with" is through normalization factor in the definition of "conditional probability". If non-unitary influence is $\vec{x}$-independent, its impact on pointwise $\rho$ will be identical to its impact on normalization factor (since the latter is an "average" of a former); thus, upon "dividing" the former by the latter we cancel the non-unitary effects. Particle-dependence (or in other words $\vec{x}$-dependence) is precisely what prevents us from doing such cancellation.
This immediately tells us what is the root of "mismatch" between statical and dynamical predictions. Namely, it is the very same thing we were talking about back in Chapter 2. Just to repeat the Chapter 2 argument, we have noticed that dynamical picture assumes that we "know" $\psi$ and, at the same time, "not know" $\vec{x}$. This implicitly assumes $\vec{x}$-independence of $\psi$. \emph{Only} when such assumption is true do we expect the equation produced by "dynamical picture" to hold. In the latter case the "dynamically-produced" equation, in fact, agrees with 'statically-produced" one. In all other cases, the dynamical picture no longer works; but we now have a static picture to "fill in a gap".
However, in light of the fact that divergence "sums" different components, it is possible to hope that $\vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle$ is $\vec{x}$-independent, despite $\vec{x}$-dependence of different "terms" in the "sum". Let us, therefore, attempt to evaluate this divergence to see if that is, in fact, the case. Clearly, $\vec{\nabla}_{\psi} \vert v_{\psi} \rangle$ on the right hand side involves uncountably many degrees of freedom. We can identify their "sum" with corresponding "integral" multiplied by "infinitely large" factor $K$:
\beq \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle = K \int d^{3N} y \; \frac{\delta}{\delta \langle y \vert \psi \rangle} \langle y \vert v_{\psi} (\psi, x) \rangle, \eeq
where $x$ is a position of a "beable particle" while $y$ is a "dummy index" representing "arbitrary component" of $\psi$ as we "take divergence". Thus, $\langle y \vert v_{\psi} (\psi, x) \rangle$ represents an effect on the value of $\psi$ around $\vec{y}$ as a result of "collapse" produced by the particle at $\vec{x}$. The definition of $\vert v_{\psi} (\psi, x) \rangle$ is a "blank space" that is to be "filled in" by a "collapse model". Let us now assume that our "collapse model" produces $\vert v_{\psi} \rangle$ in a form
\beq \vert v_{\psi} (\psi, x) \rangle = a \; \int d^{3N} y \; f(\vec{y} - \vec{x} ) \vert y \rangle \langle y \vert \psi \rangle \eeq
where $a$ is some coefficient. If we substitute this to the right hand side of precious equation, we obtain
\beq \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle = aL \int d^{3N} y \; f (\vec{y} - \vec{x}) \eeq
where $L = K \langle y \vert y \rangle =K \delta^{(3N)} (0)$ is another "infinite" constant. The expression on the right hand side is, indeed, $\vec{x}$-independent. Thus, if we substitute this into
\beq \frac{\partial}{\partial t} \frac{\rho}{\sigma} = - \vec{\nabla}_x \cdot \Big(\frac{\rho}{\sigma} \vec{v}_x \Big) - \frac{\rho}{\sigma} \Big( \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle - \int d^{3N} x \; \frac{\rho}{\sigma} \vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle \Big), \eeq
then the last two terms cancel, resulting in the expression
\beq \frac{\partial}{\partial t} \frac{\rho}{\sigma} = - \vec{\nabla}_x \cdot \Big(\frac{\rho}{\sigma} \vec{v}_x \Big). \eeq
which matches the one we obtained in the scenario free of particle's influence.
\subsection*{6. The additional influence on $\psi$: unitary and non-unitary alternatives}
While at the end of the previous section we seem to have arrived with the "desired" equation for $\rho$, there is a little catch there: $\vert \psi \vert^2$ no longer obeys that equation, since $\psi$ has an additional influence from the position $\vec{x}$ of the particle. In order to conclude that $\rho = \vert \psi \vert^2$ is an equilibrium point we need \emph{both} sides to obey the same dynamics. Yet, up till now we just took it for granted that $\psi$ would be "taken care of" and we have focused exclusively on $\rho$.
Now, there are two alternative ways of making sure that $\vert \psi \vert^2$ continues to obey the dynamics we want it to obey. One such way is to simply "do this by hand". In other words, we no longer appeal to the probability current or any other physical intuition in deriving Pilot Wave model. Instead, we simply use "inverse Laplassian" for the "probability current". We recall that our proposal of "continuous collapse" was
\beq \frac{\partial \psi}{\partial t} = -i H \psi - \frac{a}{2} (\vec{x} - \vec{x}_B)^2 \psi \eeq
which implies that
\beq \frac{\partial \vert \psi \vert^2}{\partial t} = - a \vert \vec{x} - \vec{x}_B \vert^2 \vert \psi \vert^2 \eeq
which implies that the above can be satisfied if we set the probability current to be
\beq \vec{j} = -a \nabla^{-2} (\vert \vec{x} - \vec{x}_B \vert^2 \vert \psi \vert^2 ) \eeq
which, in turn, can be satisfied by setting Pilot Wave model to be
\beq \vec{v}_x = - \frac{a}{\vert \psi \vert^2} \vert \vec{x} - \vec{x}_B \vert^2 \vert \psi \vert^2 \eeq
This approach was, in fact, used both in \cite{Epstein} and in \cite{Sverdlov} for other purposes. For example, in \cite{Sverdlov} the "non-locality" was introduced for the purposes of incorporating creation and annihilation of particles which, due to its discreteness, is "non-local" in nature and, therefore, is difficult to accommodate through "gradient".
However, at least in non-relativistic case we have no other reason of introducing such non-locality. So it makes sense to ask ourselves whether or not we can avoid it. In fact, we can! The dynamics proposed above can be rewritten as
\beq V' = V - ia (\vec{x} - \vec{x}_B)^2 \eeq
Now, we already know that the Pilot Wave model, as defined by Bohm, works for the case of \emph{real} potential:
\beq V \in \mathbb{R} \; \Longrightarrow \; \vec{v} = \frac{1}{m} \; \vec{\nabla} \; Im \; ln \; \psi \; {\rm WORKS} \eeq
The "complex value" is a problem since the evolution is no longer unitary and, therefore, can not be interpreted as "probability current flow" which is what Bohmian mechanics is bases upon. The way to "fix" it is simply to replace $-ia$ with $+a$:
\beq V' = V + a (\vec{x} - \vec{x}_B)^2 \eeq
on physical grounds we would expect the "high potential" to "keep the wave away" from the "far away" region from the particle. At the same time, if $a$ is very small its effect is negligeable as long as the distance is "reasonable"; but, at the same time, it would become important in the event of entanglement since the distance in \emph{configuration space} would become "large". This, in fact, is the only sought-after outcome of GRW collapse scenario and, therefore, can replace the latter in our theory.
Incidentally, setting $V$ to be real allows us to avoid "uncountable divergence" associated with $\vec{\nabla}_{\psi} \cdot \vert v_{\psi} \rangle$. If we recall that complex divergence is given by
\beq div \; F \; = \; 2 \; Re \; \frac{\partial F}{\partial z} \eeq
we will see that by setting $H$ to be Hermitian (which is generalization of "real"), $iH$ will be a generalization of "imaginary", which would set the above real parts (which form "uncountably many terms") to zero. This means that $\vec{\nabla}_{\psi} \cdot (\rho \vert v_{\psi} \rangle)$ will reduce to $\vert v_{\psi} \rangle \cdot \vec{\nabla} \rho$ which can be thought of as "one dimensional" derivative, as opposed to "infinite dimensional" one. Thus, we would be able to get rid of uncountable divergences in a lot less sloppy way, without having to resort to "uncountably large coefficients" that we have been using.
There is, however, a question that one can ask: if we change the factor of $i$, then our dynamic equation becomes
\beq \frac{\partial \psi}{\partial t} = -i H \psi + \frac{ia}{2} (\vec{x} - \vec{x}_B)^2 \psi \eeq
which seems to be saying that additional term is only a phase factor. The good news, however, is that the same question can be asked within the bounds of any textbook quantum mechanics problem: why would the probability of finding a particle within a potential well be "very small" if the potential only contributes to the unitary phase factor? We can, therefore, satisfy ourselves by simply answering that quantum mechanics question within the more familiar context we have just referred to.
The above question, of course, can be easily answered within the context of time independent Schrodinger's equation. In time dependent context it is a bit more tricky for the above stated reason. However, if we apply$-iH \delta t$ \emph{twice} we will get some clues. After we apply $-iH\delta t$ the \emph{first time}, the magnitude $\vert \psi \vert$ would stay unaffected, \emph{but} the $\partial^2 \psi / \partial x^2$ would pick up "imaginary component" relative to $\psi$. Since the second derivative is part of Hamiltonian, we would conclude that $H$ would now have imaginary part as well. Thus, $iH$ would now have a "real part".Thus, after applying $-iH \delta t$ the \emph{second time} we would, in fact, affect magnitude.
Now, consider one dimensional scenario. If $\partial^2 V / \partial x^2>0$, and if we assume that we have "started out" from $\psi$ being purely real and positive, this would imply that the sign of $\partial^3 \psi /\partial t \partial x^2$ is $-i$. Thus, after "very small" time interval, $\partial^2 \psi / \partial x^2$ will pick up a small $-i$ component.Thus, kinetic part of Hamiltonian would pick a small $+i$ component. Thus,$\partial \psi / \partial t$, being proportional to $-iH$ would pick a small $+1$ component. Now, the "potential barrier" is characterized by$\partial^2 V / \partial x^2 <0$ which means that $\partial \psi /\partial t$ will pick $-1$ sign, as desired.
The "time reversal" symmetry is taken care in the following way. We have assumed that "at the beginning" $(t=t_0)$ there is no phase factor. This,however, implies that at $t<t_0$ there is a phase factor "in the opposite direction". Thus, as we go "back in time" we will find our wave function decreasing at the places of high potential. Equivalently, if we go"forward in time" then, before $t=t_0$, the wave function has increased to the "original value" and then later started to decrease. Thus, the "time reversal" symmetry is broken down simply because we have assumed that we are living in $t > t_0$. Again, we have concluded that at $t> t_0$ the time derivative has the same sign as the second space derivative (which implies increase at the low potential and decrease at the high potential),while at $t<t_0$ the situation is the opposite.
This, however, raises another question: what if potential increases away from the center to infinity, while having positive second derivative all along? Our previous argument seems to imply that the wave function would increase away from the center while physically we would expect the opposite to be the case. This question can be answered by noticing that in our qualitative analysis we were exclusively focused on phase and,therefore, assumed that amplitude is constant in space at $t=t_0$. If such is not the case, then \emph{despite} the $-i$ factor in $\partial^2 (e^{i\theta})/ \partial x^2$, if we include amplitude ($r$) we might well have$+i$ in $\partial^2 (r e^{i \theta})/ \partial x^2$.
It is easy to see that it is the logarithm of the coefficient that plays the main role here and. Thus, the fact that logarithm rapidly becomes"very large" with the negative sign towards the boundary of the support of the wave function implies that the coefficient there will, in fact, be $+i$, which means that the function near the boundary will further decrease instead of increasing and we have avoided our problem. This argument, however, relies on the fact that the support of the wave function is, in fact finite.
While the assumption of finite support seems to be justified by normalization, it doesn't have to be the case, as exemplified by "fixed momentum" particle. In the cases of unbounded support we would, in fact,get exotic cases of further increase of wave function we have discussed earlier. Studying these cases is beyond the scope of this paper; but we will investigate these in the upcoming paper (see \cite{oscillator}). for our present purposes, we can be content with "cosmological" assumption that the support of wave function happened to be bound "at the beginning of the universe".
It is interesting to note that our analysis could be copied for a "more standard" GRW theory that involves discrete hits. In this case, we would conclude that a "hit" does not have to be a multiplication by $e^{-ax^2}$as typically assumed. Instead, it could be a multiplication by$e^{-iax^2}$. In this case, the "hit" itself would have no influence on the magnitude of a wave \emph{right after} its occurrence; \emph{but} it would set its "phase factor" in such a way that \emph{subsequent Schr"/odinger's evolution} would result in the desired decrease of the amplitude, similar to the one we have discussed.
The reason in standard GRW approach this was not done is probably because discrete hits don't have much in common with Hamiltonian \emph{even if}the extra $i$-factor was inserted; thus, it is logical to "make things simple" by using $e^{-ax^2}$ since there is not much to be gained by being"clever". In our case, however, by replacing discrete hits with a continuous process, we have made our picture "almost the same" as Hamiltonian; the only difference is precisely that $i$ factor. This factor, further, place deciding role between us using "ugly" $\nabla^{-2}$ in Pilot Wave model versus "beautiful" $\vec{\nabla}$ originally proposed by Bohm. For these reasons we choose to introduce this factor, unless we are dealing with models where we need $\vec{\nabla}^{-2}$ regardless (such as \cite{Epstein} and \cite{Sverdlov}).
\subsection*{7. Momentum conservation}
As was stated in the before, the definition of $\vert v_{\psi} \rangle$ amounts to specifying a "collapse mechanism". We have further found out that our picture would provide the desired result as long as collapse mechanism is "translationally invariant"; or, in other words, if the particle is located at $\vec{x}$, then its influence on the behavior of the wave around $\vec{y}$ is a function strictly of $\vec{y} - \vec{x}$. As long as the latter is true, we do get the desired result, regardless of the specifics of the mechanism.
In light of this, we can, in principle, say that we are done and leave the specification of collapse mechanism to the reader. However, I believe that it is worth elaborating on the mechanism for reasons independent of the main subject of the paper. In particular, if we use the mechanism presented by GRW models, and apply it to a single harmonic, then its multiplication by Gaussian would produce harmonics of other frequencies; thus, "conservation of momentum" would be violated.
In principle, this does not disprove a model: after all, we have never "seen" momentum. We have only seen an arrow of measuring apparatus pointing in a certain direction (which signifies a certain momentum). That direction, of course, is defined in "position space". Thus, one might argue that as long as GRW model makes correct predictions in "position space", it can predict all of the "momentum measurements" and find them consistent with predictions of quantum mechanics. Since stadard quantum mechanics predicts that momentum conservation "holds in the lab", so does GRW model, despite the subtle contradictions with momentum conservation in the setup. However, I still believe that the models that respect momentum conservation are a lot better aesthetically since only in the latter case we can explicitly include the "momentum exchange between particle and a measuring device" into the model at hand.
Consider the following scenario. Suppose we first measure momentum of a particle. Then we measure its position, and then, again measure momentum the second time. The second momentum observation, of course, does not match the first one. This can be explained through the "momentum exchange" the particle has with measuring device. This "momentum exchange" is the result of the fact that, in order to measure position, we have to "localize" measuring device in space. Thus, by uncertainty principle, the latter has uncertain momentum. That is a reason why it can pass "uncertain momentum" onto a particle. On the other hand, if we had "less precise" position measurement, we would not have to "localize" measuring device as much. Thus, the momentum of the latter would be less uncertain, which is why the momentum it "passes" onto a particle would be less uncertain as well.
Now, in the above paragraph we were appealing to the "exchange of momentum" between the particle and measuring apparatus. But, if momentum could be created through multiplication by Gaussian, then the discussion about "momentum exchange" becomes relatively meaningless, since the notion of "exchange" intrinsically appeals to "conservation". Thus, in order to "save" our "momentum exchange" argument, we have to redefine $V(\vec{y}- \vec{x})$ in such a way that it is no longer a "source of momentum" but, rather, a "catalyst of momentum exchange". Here, it is understood that by $V$ we typically mean either "real" or "imaginary" options discussed in previous section:
\beq V = -\frac{ax^2}{2} \; {\rm OR} \; V = -i\frac{ax^2}{2} \eeq
but, since our argument only depends on $V$ being "very large" at the "large distances" we will leave it in a general form.
We recall from Noether's theorem that momentum conservation is equivalent of translational covariance. The reason $V (\vec{y} - \vec{x})$ violates the former is that it violates translational covariance by setting $\vec{x}$ as "preferred value of $\vec{y}$". Strictly speaking, within the "particle-wave" context, the translational symmetry was \emph{not} violated since $\vec{x}$ is simply a position of a particle which means that it is not "better off" than any other location would have been if the particle were to move there. However, the context of Noether's theorem is wave-alone as opposed to particle+wave. In wave-alone context we \emph{do} violate the translational symmetry. Furthermore, we are \emph{not} in a position to "redo" Noether's theorem for particle+wave context, since the "momentum conservation" would require point-like excitations of the wave to "balance" the point-like momentum changes produced by particle's acceleration.
What we can do instead is to take advantage of the fact that $\vec{x}$ represents a \emph{configuration of particles} (as opposed to one single particle) and, therefore, replace the dependence on $\vec{x} = (\vec{x}_1, \cdots \vec{x}_N)$ with a dependence on a \emph{set of vectors},
\beq \{ \vec{x}_j - \vec{x}_i \vert 1 \leq i, j \leq N \}, \eeq
where $N$ is the total number of particles a "configuration space" is meant to describe. The "change" of $\vec{x}_i - \vec{x}_j$ refers to "simultaneous translation" of $\vec{x}_i$ and $\vec{x}_j$ in such a way that their center of mass stays fixed. It is easy to see that fixed center of mass implies the fixed total momentum, while variation of $\vec{x}_i - \vec{x}_j$ implies momentum exchange. Furthermore, in order to avoid the violation of angular momentum conservation as well, we have to demand the rotational symmetry. This can be done by simply replacing $\vec{x}_i - \vec{x}_j$ with $\vert \vec{x}_i - \vec{x}_j \vert$. Thus, we propose the definition of $f$ in the following form:
\beq V (\vec{y} - \vec{x}) = \sum_{i, j} U(\vert \vec{x}_j - \vec{x}_j \vert, \vert \vec{y}_j - \vec{y}_j \vert), \eeq
where $g (\rho, \sigma)$ is designed in such a way that it has a "minimum" when $\rho = \sigma$ and stays "very close" to the minimum whenever $\vert \rho - \sigma \vert$ is "neither small nor large"; but then $g(\rho, \sigma)$ "explodes" when $\rho - \sigma$ becomes "large". This would allow us to "mimic" GRW argument: the effects of $g$ are negligible unless they are "magnified" through entanglement. At the same time, the momentum conservation will be respected.
In the above equation we refrained from replacing $U (\rho, \sigma)$ with $U (\sigma - \rho)$ for the following reason. We would like the momentum exchange to be "stronger" between the particles "nearby" than between the ones "far away". Now, the former is represented by both $\rho$ and $\sigma$ being "small", while the latter is represented by them both being "large". Thus, if $\rho$ and $\sigma$ are relatively small, while $\Lambda$ is "very large", we would expect that
\beq U( \rho + \Lambda, \sigma + \Lambda) \ll U (\rho, \sigma) \eeq
Unfortunately, however, we are unable to remove non-locality completely. While we could set $U (\rho, \sigma) =0$ for "very large" $\rho$ and $\sigma$, we have to allow it to be non-zero when the latter are "small enough"; thus, our momentum exchange is at best "quasi-local". This situation, however, is not any worse from the non-locality we need to adopt \emph{anyway} in the definition of configuration space in order to write down either Schr"/odinger's equation or Pilot Wave model, to begin with.
\subsection*{8. Conclusion}
In this paper we have combined Bohmian particles and GRW collapse model into a single theory, where Bohmian particle serves as a "center" around which a continuous GRW-type collapse occurs. This seems to improve things on both ends. On GRW end, we no longer need to impose a "spontaneous collapse" (as was done in \cite{GRW1} and \cite{GRW2}), nor do we need to come up with any additional mechanism of deterministic continuous collapse (\cite{continuous}). On the Bohmian end, we no longer need to rely on "decoherence" argument which was disputed by Leggett (\cite{Leggett1} and \cite{Leggett2}) and which, if true, can result in "traveling between universes" in non-local Pilot Wave models (\cite{Epstein}, Chapter 3, and \cite{Sverdlov}).
This work is a direct result of my private discussions with Ward Struyve regarding the "collapse mechanism" I proposed in \cite{Sverdlov}. He has suggested to replace that mechanism with "spontaneous collapses" of GRW model, centered around Bohmian particle. However, I found that idea to violate Ocams razor since Bohmian particle doesn't seem to play any role in "improving upon" GRW model, which makes its presence in the theory unnatural. However, I decided that this conflict with Ocam's razor can be fixed if Bohmian particle is a source of \emph{continuous} localization, which would furnish it with far more significant role in the theory. Incidentally, this would also harmonize with my deterministic world view a lot more than spontaneous collapse models (with or without Bohmian particle).
In future work it might be interesting to compare some finer implications of a choice of using Bohmian particle as a "continuous collapse" source as contrasted with other \emph{deterministic} continuous collapse sources (such as presented in \cite{continuous}). For example, the particle follows a continuous one-dimensional trajectory, which can not be said about other kinds of deterministic sources of continuous collapse. This might result in a different kinds of "resonances" produced through the continued collapse process which might either strengthen or weaken different aspects of collapse.
Another offshoot for future work is to explore the modifications of GRW mechanism proposed in Chapters 6 and 7 independently of a "bigger context" of this paper. In particular, \emph{despite} the fact that I favor "continuous collapse" scenario, it might be still interesting to apply Chapters 6 and 7 to the "stochastic" localization events, as originally introduced in \cite{GRW1} and \cite{GRW2}. In particular, it might be interesting to add extra $i$ to the "stochastic localization" hit (which would parallel Chapter 6) and also re-define "stochastic localization" hit in such a way that it would conserve momentum (which would parallel Chapter 7).
I believe both of these are logical things to do: neither extra $i$ nor momentum conservation are "logically related" to a belief in continuous collapse; therefore, both should be adaptable to a discrete collapse settings. I believe it is important to separate logically unrelated modifications to theories and explore the effects of each such modification on its own which would enable us not to confuse the "ups" and "downs" of each such modification with the "ups" and "downs" that would arise from other ones. After that, of course, it might be logical to try different combinations of "including" some modifications and not others. This will be one of the things I will investigate in future work.
{\bf Acknowledgement:} I would like to thank Ward Struyve for suggestion that Bohmian particle is a source of stochastic collapse which ultimately lead me to think of a "continuous and deterministic" version of his suggestion in the form of a present paper.
|
\subsection{Theorem\xspace{\ifx&{}\else{ (#1)}\fi}}\slshape}{\upshape}
\newenvironment{Prop}[1][]{\subsection{Proposition\xspace{\ifx&{}\else{ (#1)}\fi}}\slshape}{\upshape}
\newenvironment{Pro}[1][]{\subsection{Proposition\xspace{\ifx&{}\else{ (#1)}\fi}}\slshape}{\upshape}
\newenvironment{Lem}[1][]{\subsection{Lemma\xspace{\ifx&{}\else{ (#1)}\fi}}\slshape}{\upshape}
\newenvironment{Cor}[1][]{\subsection{Corollary\xspace{\ifx&{}\else{ (#1)}\fi}}\slshape}{\upshape}
\theoremstyle{note}
\newenvironment{Example}[1][]{\subsection{Example\xspace{\ifx&{}\else{ (#1)}\fi}}}{}
\newenvironment{Rem}[1][]{\subsection{Remark\xspace{\ifx&{}\else{ (#1)}\fi}}}{}
\newenvironment{Obs}[1][]{\subsection{Remark\xspace{\ifx&{}\else{ (#1)}\fi}}}{}
\newenvironment{Defn}[1][]{\subsection{Definition\xspace{\ifx&{}\else{ (#1)}\fi}}}{}
\newenvironment{Assum}[1][]{\subsection{Assumption\xspace{\ifx&{}\else{ (#1)}\fi}}}{}
\newenvironment{Hyp}[1][]{\subsection{Assumption\xspace{\ifx&{}\else{ (#1)}\fi}}}{}
\newenvironment{Pbm}[1][]{\subsection{Problem\xspace{\ifx&{}\else{ (#1)}\fi}}}{}
\providecommand{\ddt}{}
\renewcommand{\ddt}{\d_t}
\providecommand{\theHalgorithm}{\arabic{algorithm}}
\providecommand{\semidiscrete}{semidiscrete\xspace}
\providecommand{\Semidiscrete}{Semidiscrete\xspace}
\providecommand{\fullydiscrete}{fully discrete\xspace}
\providecommand{\Fullydiscrete}{Fully Discrete\xspace}
\providecommand{\e}{}
\renewcommand{\e}[1]{\ensuremath{e}^{#1}}
\providecommand{\Le}[1]{\ensuremath{\leb{#1}}}
\providecommand{\hoz}{\sobhz1(\W)}
\providecommand{\hob}{\sobh{1}_b(\W)}
\providecommand{\V}[1]{\fes^{#1}}
\providecommand{\Vb}[1]{\fes^{#1}_b}
\providecommand{\X}[1]{\mathcal{X}^{#1}}
\providecommand{\bi}[2]{\abil{#1}{#2}}
\providecommand{\ltwopbigg}[2]{\ensuremath{\bigg\langle#1,#2\bigg\rangle}}
\providecommand{\ellop}{\ensuremath{\cA}}
\providecommand{\laginterpol}[1]{\La^{#1}
\providecommand{\coarseinterpol}[1]{\mathcal I_0^{#1}}
\providecommand{\clement}{\Pg}
\providecommand{\cleminterpol}[1]{\clement^{#1}}
\providecommand{\threshhold}{\xi}
\providecommand{\coarsethresh}{\kappa}
\providecommand{\EI}{\ensuremath{\operatorname{EI}}\xspace}
\providecommand{\tol}{\ensuremath{\operatorname{tol}}\xspace}
\providecommand{\enorm}[1]{\ensuremath{\Norm{#1}}_a}
\providecommand{\dU}[1]{\ensuremath{\partial U^{#1}}}
\providecommand{\T}[1]{\cT^{#1}}
\providecommand{\s}{\cS}
\providecommand{\coarseT}[1]{{\mathcal T}_0^{#1}}
\providecommand{\patch}[1]{\hat{#1}}
\providecommand{\ie}{i.e.,\xspace}
\providecommand{\linop}{\ensuremath{\cL}}
\providecommand{\nonlinop}[1]{\ensuremath{\cN}\left[ #1 \right]}
\providecommand{\dnonlinop}[1]{\ensuremath{\D \cN}\left[ #1 \right]}
\providecommand{\funq}[1]{\ensuremath{F}\left( #1 \right)}
\providecommand{\dfunq}[1]{\ensuremath{\D F}\left( #1 \right)}
\providecommand{\tr}[1]{\ensuremath{\operatorname{tr}{#1}}}
\providecommand{\hold}[2]{\ensuremath{\CC^{#1,#2}}}
\providecommand{\meas}[1]{\lvert#1\rvert}
\providecommand{\fesh}{\ensuremath{\fespace}}
\providecommand{\rangefromto}[3]{\ensuremath{#1=#2,\ldots,#3}}
\providecommand{\figwidth}{\textwidth} \providecommand{\figscale}{1}
\providecommand{\fe}[1]{{#1}_h}
\providecommand{\fes}{\ensuremath{\mathbb V}}
\renewcommand{\sin}[1]{\ensuremath{\operatorname{sin}\left(#1\right)}}
\renewcommand{\cos}[1]{\ensuremath{\operatorname{cos}\left(#1\right)}}
\renewcommand{\arcsin}[1]{\ensuremath{\operatorname{arcsin}\left(#1\right)}}
\renewcommand{\arctan}[1]{\ensuremath{\operatorname{arctan}\left(#1\right)}}
\providecommand{\Est}[1]{\ensuremath{\cE\qb{#1}}}
\providecommand{\projinter}[1]{\ensuremath{\Pg^{#1}}}
\providecommand{\ltwoproj}[1]{\ensuremath{\operatorname{P}_{\fes} #1}}
\providecommand{\tildevec}[1]{\ensuremath{\geovec{\underline{#1}}}}
\providecommand{\con}[1]{\ensuremath{\kappa(#1)}}
\providecommand{\GZZ}[1]{\ensuremath{\operatorname{G}^{#1}_{ZZ}}}
\providecommand{\GZZsq}{\ensuremath{\operatorname{G}^2_{ZZ}}}
\providecommand{\Gltwo}{\ensuremath{\operatorname{G}_{\ltwoproj{}}}}
\providecommand{\Gltwosq}{\ensuremath{\operatorname{G}_{\ltwoproj{}}^2}}
\providecommand{\Estfunk}{\ensuremath{\cE}}
\providecommand{\myspan}[1]{\ensuremath{\operatorname{span}\left(#1\right)}}
\providecommand{\MA}{Monge--Amp\`ere }
\providecommand{\cof}[1]{\operatorname{Cof} #1}
\renewcommand{\det}[1]{\operatorname{det} #1}
\providecommand{\tensorp}[2]{\ensuremath{\left\langle#1\otimes#2\right\rangle}}
\providecommand{\frob}[2]{\ensuremath{{#1}{:}{#2}}}
\renewcommand{\div}[1]{\operatorname{div}\left(#1\right)}
\renewcommand{\H}{\ensuremath{\vec{H}}}
\providecommand{\Hs}{\ensuremath{\mathcal{H}^s}}
\providecommand{\symm}{\Symmatrices d}
\providecommand{\feszero}{\mathring{\fes}}
\providecommand{\Nzero}{\mathring{N}}
\providecommand{\Nd}{{N}_{\pd{}{}}}
\providecommand{\Phizero}{\mathring{\Phi}}
\providecommand{\A}{\ensuremath{\vec{A}}}
\providecommand{\N}{\ensuremath{\vec{N}}}
\providecommand{\Npert}{\ensuremath{\tilde{\vec{N}}}}
\providecommand{\gpert}{\ensuremath{\tilde{g}}}
\providecommand{\residual}{\cR}
\providecommand{\jresidual}{\cJ}
\providecommand{\matentry}[3]{{#1}_{#2,#3}}
\renewcommand{\matentry}[3]{{#1}_{#2,#3}}
\providecommand{\nlo}{\cN}
\providecommand{\nlop}[1]{\nlo[#1]}
\providecommand{\nlfunk}{F}
\providecommand{\Dnlop}[1]{\D\cN[#1]}
\providecommand{\MAD}{Monge--Amp\`ere--Dirichlet\xspace}
\providecommand{\numPhizero}{\mathring{\smash{\numvec \Phi}}}
\providecommand{\uoo}{{\circ\circ}\atop{U}}
\providecommand{\foo}{{\circ\circ}\atop{F}}
\providecommand{\maresidual}[1]{\cR[#1]}
\newcommand{\eig}[1]{\sigma\qp{#1}}
%
%
%
%
%
%
%
%
\usepackage{ifthen}
\usepackage{amsmath}
\usepackage{amssymb}
\usepackage{stmaryrd}
\usepackage{eufrak}
\usepackage{mathrsfs}
\usepackage{bbold}
\provideboolean{usemathrsfs}
\setboolean{usemathrsfs}{true}
\usepackage{enumerate
\usepackage{xspace
\usepackage[greek,british,english]{babel}
\usepackage{verbatim
\usepackage{fancyvrb}
\usepackage{hyperref}
\usepackage{listings
\usepackage{xcolor}
\provideboolean{usebeamer
\provideboolean{isthesis
\provideboolean{isamsltex
\provideboolean{issiamltex
\usepackage{xcolor}
\colorlet{a}{red}
\colorlet{b}{blue}
\colorlet{c}{green!50!blue}
\colorlet{d}{magenta}
\colorlet{e}{cyan}
\colorlet{f}{yellow!50!black}
\colorlet{g}{white}
\colorlet{h}{black!50}
\colorlet{i}{black}
\colorlet{j}{black!75}
\providecommand{\linkedurl}[1]{\url{#1}}
\providecommand{\linkedemail}[1]{\href{mailto:#1}{#1}}
\providecommand{\email}[1]{{\linkedemail{#1}}}
\providecommand{\Ignore}[1]{}
\providecommand{\ignore}[1]{}
\providecommand{\freeze}[1]{
\providecommand{\crossout}[1]{{\textcolor{red!20}{#1}}}
\providecommand{\highlight}[1]{{\color{blue}#1}}
\providecommand{\standout}[1]{\colorbox{a}{\textcolor{g}{#1}}}
\providecommand{\memotemp}{}
\provideboolean{shownotes}
\setboolean{shownotes}{true
\newcounter{margnote}[page]
\providecommand{\margnotemark}{{\standout{\upshape\texttt{\arabic{margnote}}}}}
\providecommand{\margnote}[2][]{
\ifthenelse{
\boolean{shownotes}
}{
\stepcounter{margnote}
\margnotemark
\marginpar{
\texttt{
\begin{minipage}{2cm}
\raggedright\tiny
\margnotemark
{\ifx&{}\else{#1 says:}\fi}
#2
\end{minipage}
}
}
}{
}
}
\providecommand{\mathnote}[2][]{
\ifthenelse{
\boolean{shownotes}
}{
\stepcounter{margnote}
\text{
\standout{
\texttt{
\tiny
\margnotemark
#1:
#2
}
}
}
}{
}
}
\provideboolean{showtodo}
\providecommand{\inserttext}[1]{%
\stepcounter{margnote}%
\standout{$\mathtt{\arabic{margnote}}\atop\rightthreetimes$}
\marginpar{%
\standout{$\mathtt{\arabic{margnote}}\atop\rightthreetimes$}\highlight{#1}}%
}
\providecommand{\insertmath}[1]{\highlight{#1}}
\providecommand{\typofixtext}[2]{\crossout{#1}\marginpar{\standout{#1}}}
\providecommand{\typofixmath}[2]{
\crossout{#1}\atop{\highlight{#2}}
}
\providecommand{\todo}[1]{\ifthenelse{\boolean{showtodo}}{\margnote[To do.]{#1}}{}}
\providecommand{\Todo}[1]{
\ifthenelse{\boolean{showtodo}}{
\begin{center}
\begin{tikzpicture}
\node[fill=a!17]{
\begin{minipage}{\textwidth}
\texttt{\bfseries{\small #1}}
\end{minipage}
};
\end{tikzpicture}
\end{center}
}{}}
\providecommand{\toreview}[1]{{\color{a}#1}}
\newenvironment{Toreview}{\color{a}\par}{\color{i}}
\providecommand{\Warning}[1]{
\begin{tikzpicture}
\node[fill=a!27]{
\begin{minipage}{\textwidth}
\texttt{\bfseries{\small Warning: #1}}
\end{minipage}
};
\end{tikzpicture}
}
\providecommand{\checkit}{[$\ast$]\marginpar{$[\ast]$: {\tiny Check!}}\xspace}
\providecommand{\omitfirst}{[$\ast$]\marginpar{$[\ast]$: {\tiny May be omitted on 1st reading.}}\xspace}
\provideboolean{showcomments}
\providecommand{\margincomment}[1]{
\ifthenelse{\boolean{showcomments}}{\marginpar{\tiny #1}}{}
}
%
\provideboolean{showchanges}
\setboolean{showchanges}{false}
\providecommand{\changes}[1]{
\ifthenelse{\boolean{showchanges}}{{\highlight{#1}}}{#1}
}
\providecommand{\changesbox}[1]{\par\colorbox{cyan}{\begin{minipage}{\linewidth}#1\end{minipage}}}%
\providecommand{\changefromto}[3][replace with]{
\ifthenelse{\boolean{showchanges}}
{{\crossout{#2}\margnote{#1}}{\highlight{#3}}}
{#3\xspace}
}
\providecommand{\ChangePar}[2]{
\ifthenelse{\boolean{showchanges}}
{{\par$\mapsfrom$ \textcolor{red!20}{#1}}{\par$\mapsto$ \textcolor{blue}{#2}}}
{\par #2}
}
\providecommand{\InsertPar}[1]{
\ifthenelse{\boolean{showchanges}}
{{\par$\mapsto$ \textcolor{blue}{#1}}}
{\par #1}
}
\providecommand{\mathchangefromto}[2]{\crossout{#1}{\color{magenta}\mapsto}\highlight{#2}}
\providecommand{\changedisplayfromto}[2]{
\changefromto{#1\addtocounter{equation}{-1}}{#2}
}
\providecommand{\ChangeText}[2]{\changefromto[sub with]{#1}{#2}}
\providecommand{\InsertText}[1]{\changefromto[insert]{}{#1}}
\providecommand{\DeleteText}[1]{\changefromto[delete]{#1}{}}
\providecommand{\InsertMath}[1]{\mathchangefromto{{}^{\texttt{\tiny{ins}}}}{#1}}
\providecommand{\DeleteMath}[1]{\mathchangefromto{#1}{{}^{\texttt{\tiny{del}}}}}
\usepackage{mathrsfs
\provideboolean{usemathrsfs}
\setboolean{usemathrsfs}{true
\providecommand{\mathscript}
{\mathscr}
\providecommand{\cA}{\ensuremath{\mathscript A}\xspace}
\providecommand{\cB}{\ensuremath{\mathscript B}\xspace}
\providecommand{\cC}{\ensuremath{\mathscript C}\xspace}
\providecommand{\cD}{\ensuremath{\mathscript D}\xspace}
\providecommand{\cE}{\ensuremath{\mathscript E}\xspace}
\providecommand{\cF}{\ensuremath{\mathscript F}\xspace}
\providecommand{\cG}{\ensuremath{\mathscript G}\xspace}
\providecommand{\cH}{\ensuremath{\mathscript H}\xspace}
\providecommand{\cI}{\ensuremath{\mathscript I}\xspace}
\providecommand{\cJ}{\ensuremath{\mathscript J}\xspace}
\providecommand{\cK}{\ensuremath{\mathscript K}\xspace}
\providecommand{\cL}{\ensuremath{\mathscript L}\xspace}
\providecommand{\cM}{\ensuremath{\mathscript M}\xspace}
\providecommand{\cN}{\ensuremath{\mathscript N}\xspace}
\providecommand{\cO}{\ensuremath{\mathscript O}\xspace}
\providecommand{\cP}{\ensuremath{\mathscript P}\xspace}
\providecommand{\cQ}{\ensuremath{\mathscript Q}\xspace}
\providecommand{\cR}{\ensuremath{\mathscript R}\xspace}
\providecommand{\cS}{\ensuremath{\mathscript S}\xspace}
\providecommand{\cT}{\ensuremath{\mathscript T}\xspace}
\providecommand{\cU}{\ensuremath{\mathscript U}\xspace}
\providecommand{\cV}{\ensuremath{\mathscript V}\xspace}
\providecommand{\cW}{\ensuremath{\mathscript W}\xspace}
\providecommand{\cX}{\ensuremath{\mathscript X}\xspace}
\providecommand{\cY}{\ensuremath{\mathscript Y}\xspace}
\providecommand{\cZ}{\ensuremath{\mathscript Z}\xspace}
\providecommand{\hA}{\ensuremath{\mathcal A}\xspace}
\providecommand{\hB}{\ensuremath{\mathcal B}\xspace}
\providecommand{\hC}{\ensuremath{\mathcal C}\xspace}
\providecommand{\hD}{\ensuremath{\mathcal D}\xspace}
\providecommand{\hE}{\ensuremath{\mathcal E}\xspace}
\providecommand{\hF}{\ensuremath{\mathcal F}\xspace}
\providecommand{\hG}{\ensuremath{\mathcal G}\xspace}
\providecommand{\hH}{\ensuremath{\mathcal H}\xspace}
\providecommand{\hI}{\ensuremath{\mathcal I}\xspace}
\providecommand{\hJ}{\ensuremath{\mathcal J}\xspace}
\providecommand{\hK}{\ensuremath{\mathcal K}\xspace}
\providecommand{\hL}{\ensuremath{\mathcal L}\xspace}
\providecommand{\hM}{\ensuremath{\mathcal M}\xspace}
\providecommand{\hN}{\ensuremath{\mathcal N}\xspace}
\providecommand{\hO}{\ensuremath{\mathcal O}\xspace}
\providecommand{\hP}{\ensuremath{\mathcal P}\xspace}
\providecommand{\hQ}{\ensuremath{\mathcal Q}\xspace}
\providecommand{\hR}{\ensuremath{\mathcal R}\xspace}
\providecommand{\hS}{\ensuremath{\mathcal S}\xspace}
\providecommand{\hT}{\ensuremath{\mathcal T}\xspace}
\providecommand{\hU}{\ensuremath{\mathcal U}\xspace}
\providecommand{\hV}{\ensuremath{\mathcal V}\xspace}
\providecommand{\hW}{\ensuremath{\mathcal W}\xspace}
\providecommand{\hX}{\ensuremath{\mathcal X}\xspace}
\providecommand{\hY}{\ensuremath{\mathcal Y}\xspace}
\providecommand{\hZ}{\ensuremath{\mathcal Z}\xspace}
\providecommand{\bbbold}{\mathbb}
\providecommand{\rA}{\ensuremath{\bbbold A}\xspace}
\providecommand{\rB}{\ensuremath{\bbbold B}\xspace}
\providecommand{\rC}{\ensuremath{\bbbold C}\xspace}
\providecommand{\rD}{\ensuremath{\bbbold D}\xspace}
\providecommand{\rE}{\ensuremath{\bbbold E}\xspace}
\providecommand{\rF}{\ensuremath{\bbbold F}\xspace}
\providecommand{\rG}{\ensuremath{\bbbold G}\xspace}
\providecommand{\rH}{\ensuremath{\bbbold H}\xspace}
\providecommand{\rI}{\ensuremath{\bbbold I}\xspace}
\providecommand{\rJ}{\ensuremath{\bbbold J}\xspace}
\providecommand{\rK}{\ensuremath{\bbbold K}\xspace}
\providecommand{\rL}{\ensuremath{\bbbold L}\xspace}
\providecommand{\rM}{\ensuremath{\bbbold M}\xspace}
\providecommand{\rN}{\ensuremath{\bbbold N}\xspace}
\providecommand{\rO}{\ensuremath{\bbbold O}\xspace}
\providecommand{\rP}{\ensuremath{\bbbold P}\xspace}
\providecommand{\rQ}{\ensuremath{\bbbold Q}\xspace}
\providecommand{\rR}{\ensuremath{\bbbold R}\xspace}
\providecommand{\rS}{\ensuremath{\bbbold S}\xspace}
\providecommand{\rT}{\ensuremath{\bbbold T}\xspace}
\providecommand{\rU}{\ensuremath{\bbbold U}\xspace}
\providecommand{\rV}{\ensuremath{\bbbold V}\xspace}
\providecommand{\rW}{\ensuremath{\bbbold W}\xspace}
\providecommand{\rX}{\ensuremath{\bbbold X}\xspace}
\providecommand{\rY}{\ensuremath{\bbbold Y}\xspace}
\providecommand{\rZ}{\ensuremath{\bbbold Z}\xspace}
\providecommand{\Ae}{\ensuremath{\text{ a.e.}}\xspace}
\providecommand{\dashae}{\ensuremath{\text{-a.e.}}\xspace}
\providecommand{\as}{\ensuremath{\text{ a.s.}}\xspace}
\providecommand{\iid}{\ensuremath{\text{i.i.d.}}\xspace}
\providecommand{\qo}{\ensuremath{\text{ a.e.}}\xspace}
%
\providecommand{\ie}{\ensuremath{\text{ i.e., }}\xspace}
\providecommand{\eg}{\ensuremath{\text{ e.g., }}\xspace}
%
\providecommand{\aposteriori}{{aposteriori}\xspace}
\providecommand{\Aposteriori}{{Aposteriori}\xspace}
\providecommand{\NN}{\notag\\}
\providecommand{\MM}{\\[1.5\jot]}
\providecommand{\NM}{\notag\\[1.5\jot]}
\providecommand{\naturals}{\rN\xspace}
\providecommand{\Nat}{\ensuremath{\rZ^+}}
\providecommand{\NO}{\ensuremath{\rN_0}}
\providecommand{\integers}{\rZ}
\providecommand{\Z}[1]{\integers^{#1}}
\providecommand{\rationals}{\rQ}
\providecommand{\reals}{\rR}
\providecommand{\R}[1]{\reals^{#1}}
\providecommand{\RO}{\reals^+_0}
\providecommand{\RP}{\reals_+}
\providecommand{\complex}{\rC}
\providecommand{\C}[1]{\complex^{#1}}
\renewcommand{\C}[1]{\complex^{#1}}
\providecommand{\field}{\rF}
\providecommand{\F}[1]{\field^{#1}}
\providecommand{\sphere}[1]{\rS^{#1}}
\providecommand{\torus}[1]{\rT^{#1}}
\providecommand{\one}{\ensuremath{\bbbold 1}\xspace}
\providecommand{\charfun}[1]{\one_{#1}}
\providecommand{\iverson}[1]{\one\qb{#1}}
\providecommand{\diracdelta}[1][]{\ensuremath{\delta_{#1}}}
\providecommand{\Tolto}{\smallsetminus}
\providecommand{\take}{\Tolto}
\providecommand{\takeset}[1]{\take\setof{#1}}
\providecommand{\closure}[1]{\overline{#1}}
\providecommand{\inner}{\cdot}
\providecommand{\numminner}{{}^\transpose}
\providecommand{\tensor}{\otimes}
\providecommand{\scaprod}[2]{{#1}\inner{#2}}
\providecommand{\Ga}{\ensuremath{\varGamma}\xspace}
\providecommand{\De}{\ensuremath{\varDelta}\xspace}
\providecommand{\Th}{\ensuremath{\varTheta}\xspace}
\providecommand{\La}{\ensuremath{\varLambda}\xspace}
\providecommand{\Ks}{\ensuremath{\varXi}\xspace}
\providecommand{\Pg}{\ensuremath{\varPi}\xspace}
\providecommand{\Si}{\ensuremath{\varSigma}\xspace}
\providecommand{\Y}{\ensuremath{\varUpsilon}\xspace}
\providecommand{\Fi}{\ensuremath{\varPhi}\xspace}
\providecommand{\Ps}{\ensuremath{\varPsi}\xspace}
\providecommand{\W}{\ensuremath{\varOmega}\xspace}
\providecommand{\QW}{\ensuremath{\bar{\W}}\xspace
\providecommand{\Alpha}{\ensuremath{\mathrm A}}
\providecommand{\Beta}{\ensuremath{\mathrm B}}
\providecommand{\Epsilon}{\ensuremath{\mathrm E}}
\providecommand{\Zeta}{\ensuremath{\mathrm Z}}
\providecommand{\Eta}{\ensuremath{\mathrm H}}
\providecommand{\Iota}{\ensuremath{\mathrm I}}
\providecommand{\Kappa}{\ensuremath{\mathrm K}}
\providecommand{\Mu}{\ensuremath{\mathrm M}}
\providecommand{\Nu}{\ensuremath{\mathrm N}}
\providecommand{\Omicron}{\ensuremath{\mathrm O}}
\providecommand{\Rho}{\ensuremath{\mathrm P}}
\providecommand{\Varsigma}{\ensuremath{\mathrm C}}
\providecommand{\Tau}{\ensuremath{\mathrm T}}
\providecommand{\Chi}{\ensuremath{\mathrm X}}
\renewcommand{\a}{\ensuremath{\alpha}\xspace
\renewcommand{\b}{\ensuremath{\beta}\xspace
\providecommand{\g}{\ensuremath{\gamma}\xspace}
\providecommand{\de}{\ensuremath{\delta}\xspace}
\providecommand{\ep}{\ensuremath{\varepsilon}\xspace}
\providecommand{\epsi}{\ensuremath{\epsilon}\xspace}
\renewcommand{\th}{\ensuremath{\theta}\xspace
\providecommand{\thet}{\ensuremath{\vartheta}\xspace}
\renewcommand{\k}{\ensuremath{\kappa}\xspace}
\providecommand{\ka}{\ensuremath{\varkappa}\xspace}
\providecommand{\la}{\ensuremath{\lambda}\xspace}
\providecommand{\lam}{\ensuremath{\lambda}\xspace
\renewcommand{\r}{\ensuremath{\rho}\xspace
\providecommand{\ro}{\ensuremath{\varrho}\xspace}
\providecommand{\sig}{\ensuremath{\varsigma}\xspace}
\providecommand{\fie}{\ensuremath{\varphi}\xspace}
\providecommand{\w}{\ensuremath{\omega}\xspace}
\providecommand{\qp}[1]{\ensuremath{\!\left({#1}\right)}}
\providecommand{\qpreg}[1]{\ensuremath{(#1)}}
\providecommand{\qpbig}[1]{\ensuremath{\big(\!#1\!\big)}}
\providecommand{\qpBig}[1]{\ensuremath{\Big(\!#1\!\Big)}}
\providecommand{\qpbigg}[1]{\ensuremath{\bigg(\!\!#1\!\!\bigg)}}
\providecommand{\qpBigg}[1]{\ensuremath{\Bigg(\!\!#1\!\!\Bigg)}}
\providecommand{\qb}[1]{\ensuremath{\!\left[{#1}\right]}}
\providecommand{\qbreg}[1]{\ensuremath{[#1]}}
\providecommand{\qbbig}[1]{\ensuremath{\big[#1\big]}}
\providecommand{\qbBig}[1]{\ensuremath{\Big[#1\Big]}}
\providecommand{\qbbigg}[1]{\ensuremath{\bigg[#1\bigg]}}
\providecommand{\qbBigg}[1]{\ensuremath{\bigg[#1\Bigg]}}
\providecommand{\qc}[1]{\ensuremath{\left\{{#1}\right\}}}
\providecommand{\qcreg}[1]{\ensuremath{\{#1\}}}
\providecommand{\qcbig}[1]{\ensuremath{\big\{#1\big\}}}
\providecommand{\qcBig}[1]{\ensuremath{\Big\{#1\Big\}}}
\providecommand{\qcbigg}[1]{\ensuremath{\bigg\{#1\bigg\}}}
\providecommand{\qcBigg}[1]{\ensuremath{\bigg\{#1\Bigg\}}}
\providecommand{\qa}[1]{\ensuremath{\left\langle{#1}\right\rangle}}
\providecommand{\qareg}[1]{\ensuremath{\langle#1\rangle}}
\providecommand{\qabig}[1]{\ensuremath{\big\langle#1\big\rangle}}
\providecommand{\qaBig}[1]{\ensuremath{\Big\langle#1\Big\rangle}}
\providecommand{\qabigg}[1]{\ensuremath{\bigg\langle#1\bigg\rangle}}
\providecommand{\qaBigg}[1]{\ensuremath{\Bigg\langle#1\Bigg\rangle}}
\providecommand{\qv}[1]{\ensuremath{\left|{#1}\right|}}
\providecommand{\qvreg}[1]{\ensuremath{|#1|}}
\providecommand{\qvbig}[1]{\ensuremath{\big|#1\big|}}
\providecommand{\qvBig}[1]{\ensuremath{\Big|#1\Big|}}
\providecommand{\qvbigg}[1]{\ensuremath{\bigg|#1\bigg|}}
\providecommand{\qvBigg}[1]{\ensuremath{\Bigg|#1\Bigg|}}
\providecommand{\qe}[1]{\ensuremath{#1}}
\providecommand{\prop}[1]{\qp{#1}
\providecommand{\proptext}[1]{\qp{\text{#1}}
\providecommand{\opinter}[2]{\ensuremath{\left(#1,#2\right)}\xspace}
\providecommand{\clinter}[2]{\ensuremath{\left[#1,#2\right]}\xspace}
\providecommand{\opclinter}[2]{\ensuremath{\left(#1,#2\right]}\xspace}
\providecommand{\clopinter}[2]{\ensuremath{\left[#1,#2\right)}\xspace}
\providecommand{\mathquote}[1]{\ensuremath{\text{``}#1\text{''}}}
\providecommand{\expp}[1]{\ensuremath{\e^{#1}}}
\providecommand{\powqp}[2]{\ensuremath{\qp{#2}^{\kern -.25em\lower .2ex\hbox{\scriptsize $#1$}}\kern-.1em}}
\providecommand{\powqpreg}[2]{\ensuremath{\qpreg{#2}^{\kern -.2em\lower .3ex\hbox{\scriptsize $#1$}}\kern-.3em}}
\providecommand{\powqpbig}[2]{\ensuremath{\qpbig{#2}^{\kern -.2em\lower .3ex\hbox{\scriptsize $#1$}}\kern-.3em}}
\providecommand{\powqpBig}[2]{\ensuremath{\qpBig{#2}^{\kern -.2em\lower .3ex\hbox{\scriptsize $#1$}}\kern-.3em}}
\providecommand{\powqpbigg}[2]{\ensuremath{\qpbigg{#2}^{\kern -.2em\lower .3ex\hbox{\scriptsize $#1$}}\kern-.3em}}
\providecommand{\powqpBigg}[2]{\ensuremath{\qpBigg{#2}^{\kern -.2em\lower .3ex\hbox{\scriptsize $#1$}}\kern-.3em}}
\providecommand{\fsqrt}[1]{{#1}^{1/2}}
\providecommand{\psqrt}[1]{\powqp{1/2}{#1}}
\providecommand{\pregsqrt}[1]{\powqpreg{1\!/\!2}{#1}}
\providecommand{\norm}[1]{\ensuremath{\left|#1\right|}}
\providecommand{\normreg}[1]{\ensuremath{|#1|}}
\providecommand{\abs}[1]{\ensuremath{\left|#1\right|}}
\providecommand{\Norm}[1]{\ensuremath{\left\|#1\right\|}}
\providecommand{\Normreg}[1]{\ensuremath{\|#1\|}}
\providecommand{\Normbig}[1]{\ensuremath{\big\|#1\big\|}}
\providecommand{\Norminfty}[1]{\Norm{#1}_\infty}
\providecommand{\ltwop}[2]{\ensuremath{\qa{#1,#2}}}
\providecommand{\ltwopreg}[2]{\ensuremath{\qareg{#1,#2}}}
\providecommand{\ltwopbig}[2]{\ensuremath{\qabig{#1,#2}}}
\providecommand{\ltwopBig}[2]{\ensuremath{\qaBig{#1,#2}}}
\providecommand{\ltwopbigg}[2]{\ensuremath{\qabigg{#1,#2}}}
\providecommand{\ltwopBigg}[2]{\ensuremath{\qaBigg{#1,#2}}}
\providecommand{\crochet}[2]{\ensuremath{\left\langle #1,#2\right\rangle}}
\providecommand{\duality}[3][]{\ensuremath{#1\langle #2\,#1\vert\,#3#1\rangle}}
\providecommand{\locaverage}[1]{\Lbag#1\Rbag
\providecommand{\pairing}[2]{\ensuremath{\qp{#1,#2}}}
\providecommand{\ensemble}[2]{\ensuremath{\left\{ #1:\;#2 \right\}}}
\providecommand{\eventtext}[1]{\ensuremath{\qc{\text{#1}}}}
\providecommand{\floor}[1]{\ensuremath{\lfloor{#1}\rfloor}}
\providecommand{\ceil}[1]{\ensuremath{\lceil{#1}\rceil}}
\providecommand{\setof}[1]{{\qc{#1}}}
\providecommand{\eventof}[1]{\setof{#1}}
\providecommand{\eventoftext}[1]{{\eventof{\text{#1}}}}
\providecommand{\conditionalto}{\,\vert\,}
\providecommand{\eventcond}[3][]{\eventof{#2#1\conditionalto#3}}
\providecommand{\vecof}[1]{\begin{pmatrix}{#1}\end{pmatrix}}
\providecommand{\vecvd}[3][1]{\vecof{{#2}_{#1},\dotsc,{#2}_{#3}}}
\providecommand{\matof}[1]{\vecof{#1}}
\providecommand{\squmatcustmd}[3][\matentry]{
\begin{bmatrix}
#1{#2}{1}{1}&\dotsc{#2}{1}{#3}
\\
\vdots & \ddots &\vdots
\\
#1{#2}{#3}{1}&\dotsc{#2}{#3}{#3}
\end{bmatrix}
}
\providecommand{\squmatmd}[3][1]{
\begin{bmatrix}
\matentry{#2}{#1}{#1}&\dotsc&\matentry{#2}{#1}{#3}
\\
\vdots & \ddots &\vdots
\\
\matentry{#2}{#3}{#1}&\dotsc&\matentry{#2}{#3}{#3}
\end{bmatrix}
}
\providecommand{\pair}[2]{\qp{#1,#2}}
\providecommand{\sq}[1]{\ensuremath{\boldsymbol{\mathsf{#1}}}}
\providecommand{\sqval}[1]{\ensuremath{\sq{#1}(\naturals)}}
\providecommand{\seqof}[1]{\qp{#1}
\providecommand{\seq}[1]{\seqof{#1}
\providecommand{\seqi}[2]{\seq{#1_{#2}}
\providecommand{\seti}[2]{\setof{#1_{#2}}
\providecommand{\sequ}[2]{\seq{#1^{#2}}
\providecommand{\setu}[2]{\set{#1^{#2}}
\providecommand{\seqival}[2]{\ensuremath{\setof{#1_{#2}}}
\providecommand{\sequval}[2]{\setof{#1^{#2}}
\providecommand{\seqs}[2]{\seq{#1}_{#2}}
\providecommand{\seqsi}[3]{{\seqi{#1}{#2}}_{#2\in#3}}
\providecommand{\setsi}[3]{{\seti{#1}{#2}}_{#2\in#3}}
\providecommand{\seqsu}[3]{{\sequ{#1}{#2}}_{#2\in#3}}
\providecommand{\setsu}[3]{{\sequ{#1}{#2}}_{#2\in#3}}
\providecommand{\seqsval}[2]{\ensemble{#1}{#2}}
\providecommand{\seqsival}[3]{\ensemble{#1_{#2}}{#2\in#3}}
\providecommand{\seqsuval}[3]{\ensemble{#1^{#2}}{#2\in#3}}
\providecommand{\seqsfromto}[4]{\seqs{#1}{\rangefromto{#2}{#3}{#4}}}
\providecommand{\seqsifromto}[4]{\seqi{#1}{#2}_{\rangefromto{#2}{#3}{#4}}}
\providecommand{\setsifromto}[4]{\ensemble{{#1}_{#2}}{\rangefromto{#2}{#3}{#4}}}
\providecommand{\listudotsfromto}[3]{\ensuremath{{#1}^{#2},\dotsc,{#1}^{#3}}}
\providecommand{\seqsufromto}[4]{\sequ{#1}{#2}_{\rangefromto{#2}{#3}{#4}}}
\providecommand{\setsufromto}[4]{\ensemble{{#1}_{#2}}{\rangefromto{#2}{#3}{#4}}}
\providecommand{\setsudotsfromto}[3]{\setof{{#1}^{#2},\dotsc,{#1}^{#3}}}
\providecommand{\seqsivalfromto}[4]{\seqival{#1}{#2}_{\rangefromto{#2}{#3}{#4}}}
\providecommand{\seqsuvalfromto}[4]{\sequval{#1}{#2}_{\rangefromto{#2}{#3}{#4}}}
\providecommand{\sqin}[2]{\seqsival{#1}{#2}{\naturals}}
\providecommand{\sums}[2]{\ensuremath{\sum_{#1\in #2}}}
\providecommand{\sumi}[2]{\ensuremath{\sum_{#2}{#1}_{#2}}}
\providecommand{\sumu}[2]{\ensuremath{\sum_{#2}{#1}^{#2}}}
\providecommand{\sumsi}[3]{\ensuremath{\sum_{#2\in#3}{#1}_{#2}}}
\providecommand{\sumsu}[3]{\ensuremath{\sum_{#2\in#3}{#1}^{#2}}}
\providecommand{\sumifromto}[3]{\ensuremath{\sum_{#1=#2}^{#3}}}
\providecommand{\sumsifromto}[4]{\ensuremath{\sumifromto{#2}{#3}{#4}{#1}_{#2}}}
\providecommand{\sumsufromto}[4]{\ensuremath{\sumifromto{#2}{#3}{#4}{#1}^{#2}}}
\providecommand{\prodifromto}[3]{\ensuremath{\prod_{#1=#2}^{#3}}}
\providecommand{\prodsifromto}[4]{\ensuremath{\prodifromto{#2}{#3}{#4}{#1}_{#2}}}
\providecommand{\seri}[1]{\ensuremath{\sum_{#1=1}^\infty}}
\providecommand{\serifrom}[2]{\ensuremath{\sum_{#1=#2}^\infty}}
\providecommand{\sersi}[2]{\ensuremath{\sum_{#2=1}}^\infty#1_{#2}}
\providecommand{\sersifrom}[3]{\ensuremath{\sum_{#2=#3}}^\infty#1_{#2}}
\providecommand{\sersu}[2]{\ensuremath{\sum_{#2=1}}^\infty#1^{#2}}
\providecommand{\sersufrom}[3]{\ensuremath{\sum_{#2=#3}}^\infty#1^{#2}}
\providecommand{\vecentry}[2]{\qb{{#1}}_{#2}}
\providecommand{\matrow}[2]{\qb{\mat{#1}}_{#2}}
\providecommand{\matcol}[2]{\qb{\mat{#1}}^{#2}}
\providecommand{\bientry}[3]{\ensuremath{\qb{#1}_{#2}^{#3}}}
\providecommand{\matentry}[3]{\bientry{\mat{#1}}{#2}{#3}}
\providecommand{\nummatentry}[3]{\bientry{\nummat{#1}}{#2}{#3}}
\providecommand{\jump}[1]{\ensuremath{\left\llbracket #1\right\rrbracket}}
\providecommand{\Jump}[1]{\jump{#1}
\providecommand{\dgmean}[1]{\ensuremath{{\boldsymbol\{}#1{\boldsymbol\}}}}
\providecommand{\fromto}[2]{\ensuremath{\left[#1:#2\right]}}
\providecommand{\fromstepto}[3]{\ensuremath{\left[#1:#2:#3\right]}}
\providecommand{\integerbetween}[2]{\ensuremath{\in\fromto{#1}{#2}}}
\providecommand{\rangefromto}[3]{\ensuremath{#1\integerbetween{#2}{#3}}}
\providecommand{\loc}{\ensuremath{\mathrm{loc}}}
\providecommand{\cs}{\ensuremath{\mathrm{c}}}
\providecommand{\e}{\ensuremath{\operatorname{e}}\xspace}
\providecommand{\ic}{\ensuremath{\operatorname{i}}\xspace}
\providecommand{\Loneloc}{\ensuremath{\LL_1^{\!\loc}}}
\renewcommand{\d}{\ensuremath{\,\mathrm{d}}}
\providecommand{\ds}[1]{\ensuremath{\,\mathrm d\mathrm s#1}}
\providecommand{\D}{\ensuremath{\,\mathrm{D}}}
\providecommand{\registered}%
{\ensuremath{^\text{\textregistered}}}
\providecommand{\adjoint}[1]{\ensuremath{{#1}^*}}
\providecommand{\AS}{\ensuremath{\quad\text{as }}}
\providecommand{\AND}{\ensuremath{\text{ and }}}
\providecommand{\AAND}{\ensuremath{\times}}
\providecommand{\SAND}{\ensuremath{\quad\AND\quad}}
\providecommand{\loand}{\ensuremath{\wedge}}
\providecommand{\OR}{\ensuremath{\text{ or }}}
\providecommand{\AOR}{\ensuremath{+}}
\providecommand{\NOT}{\ensuremath{\operatorname{not}}\xspace}
\providecommand{\ANOT}{\ensuremath{1-}}
\providecommand{\LTRUE}{\ensuremath{1}}
\providecommand{\LFALSE}{\ensuremath{0}}
\providecommand{\IS}{\ensuremath{\triangleright}}
\providecommand{\arcsec}{\operatorname{arcsec}}
\providecommand{\Arg}{\operatorname{Arg}}
\providecommand{\ball}{\operatorname{B}}
\providecommand{\Ball}[2]{\ball_{{#2}}{\left({#1}\right)}}
\providecommand{\Disc}[2]{\Ball{#1}{#2}}
\providecommand{\dBall}[3]{\ball^{#1}_{#3}{\left(#2\right)}}
\providecommand{\card}{\ensuremath{\#}}
\providecommand{\const}{\operatorname{const}}
\providecommand{\constant}[1]{\ensuremath{C_{#1}}}
\providecommand{\constext}[1]{\constant{\operatorname{#1}}}
\providecommand{\constref}[2][]{\ensuremath{\constant{\text{\ref{#2}{\ifx&{}\else{,#1}\fi}}}}}
\providecommand{\codim}{\operatorname{codim}}
\providecommand{\codomain}{\operatorname{Cod}}
\providecommand{\Cconjugate}[1]{\overline{#1}}
\providecommand{\diag}{\operatorname{diag}}
\providecommand{\Diag}{\operatorname{Diag}}
\providecommand{\diam}{\operatorname{diam}}
\providecommand{\dist}{\operatorname{dist}}
\renewcommand{\div}{\operatorname{div}}
\providecommand{\dom}{\operatorname{Dom}}
\providecommand{\Dom}{\operatorname{Dom}}
\providecommand{\domain}{\dom}
\providecommand{\esssup}{\operatorname{ess\,sup}}
\providecommand{\graph}{\operatorname{graph}}
\providecommand{\hausmeas}{\operatorname{\cH}}
\providecommand{\hcf}[2]{\operatorname{hcf}\left(#1,#2\right)}
\providecommand{\hmone}{\hausmeas^1}
\providecommand{\interior}{\operatorname{int}}
\providecommand{\id}{\operatorname{id}}
\providecommand{\inverse}[1]{\ensuremath{{#1}^{-1}}}
\providecommand{\inverseqp}[1]{\inverse{\qp{#1}}}
\providecommand{\inverseqb}[1]{\inverse{\qb{#1}}}
\providecommand{\inversemat}[1]{\inverse{\mat{#1}}}
\providecommand{\inversenummat}[1]{\inverse{\nummat{#1}}}
\renewcommand{\Im}{\operatorname{im}}
\providecommand{\image}{\operatorname{Img}}
\providecommand{\length}{\operatorname{length}}
\providecommand{\lcm}[2]{\operatorname{lcm}\left(#1,#2\right)}
\providecommand{\lebmeas}[1]{\left|#1\right|}
\providecommand{\fraclpp}[2]{(#1)/(#2)}
\providecommand{\fraclpf}[2]{(#1)/{#2}}
\providecommand{\fraclfp}[2]{#1/(#2)}
\providecommand{\fracl}[2]{{#1}/{#2}}
\providecommand{\fracpp}[2]{\frac{\qp{#1}}{\qp{#2}}}
\providecommand{\fracfp}[2]{\frac{{#1}}{\qp{#2}}}
\providecommand{\fracpf}[2]{\frac{\qp{#1}}{{#2}}}
\providecommand{\Map}[2]{\ensuremath{\operatorname{Map}\left(#1,#2\right)}}
\providecommand{\rank}{\operatorname{rank}}
\providecommand{\Span}{\operatorname{span}}
\providecommand{\old}{\mathrm{old}}
\providecommand{\new}{\mathrm{new}}
\providecommand{\eye}{\operatorname{\bf I}}
\providecommand{\numeye}{\boldsymbol{\mathsf{I}}}
\providecommand{\Eye}[1]{
\begin{bmatrix}
\ifthenelse{#1>1}{
\ifthenelse{#1>2}{
\ifthenelse{#1>3}{
1&0&\dotso&0
\\
0&1&\dotso&0
\\
\vdots&\vdots&\ddots&\vdots
\\
0&0&\dotso&1
}{
1&0&0
\\
0&1&0
\\
0&0&1
}
}{
1&0
\\
0&1
}
}{
1
}
\end{bmatrix}
}
\providecommand{\zeromat}{\operatorname{\bf O}}
\providecommand{\Id}{\operatorname{Id}}
\providecommand{\idon}[1]{\id_{#1}}
\providecommand{\Idon}[1]{\id_{#1}}
\providecommand{\meas}[1]{\operatorname{meas}#1}
\providecommand{\area}[1]{\operatorname{area}#1}
\providecommand{\meshsize}[1]{\operatorname{meshsize}(#1)}
\providecommand{\maxi}[2]{#1\vee#2}
\providecommand{\mini}[2]{#1\wedge#2}
\providecommand{\maxofset}[1]{\max\setof{#1}}
\providecommand{\minofset}[1]{\min\setof{#1}}
\providecommand{\argmin}{\operatorname{argmin}}
\providecommand{\Argmin}{\operatorname{Argmin}}
\providecommand{\argmax}{\operatorname{argmax}}
\providecommand{\Argmax}{\operatorname{Argmax}}
\renewcommand{\neg}[1]{\left[#1\right]_-}
\providecommand{\negpart}[1]{\left[#1\right]_-}
\providecommand{\Oh} {\operatorname{O}}
\providecommand{\oh} {\operatorname{o}}
\providecommand{\pd}[1]{\ensuremath{\partial_{#1}}\xspace}
\providecommand{\qpd}[2]{\ensuremath{\partial_{#1}\qb{#2}}\xspace}
\providecommand{\nthpd}[3]{\ensuremath{\partial_{#1}^{#2}{#3}}}
\providecommand{\secpd}[2]{\nthpd{#1}2{#2}}
\providecommand{\dt}{\ensuremath{\d_t}}
\providecommand{\dtt}{\ensuremath{\d_{tt}}}
\providecommand{\pdt}[1][]{\pd t{{\ifx&{}\else{\qb{#1}}\fi}}}
\providecommand{\qpdt}[1]{\pdt{\qb{#1}}}
\providecommand{\pdtt}[1]{\pd{tt}{#1}}
\providecommand{\dd}[2][]{\ensuremath{\frac{\d{#1}}{\d{#2}}}}
\providecommand{\ddpow}[2]{\ensuremath{\frac{\d^{#2}}{\d{#1}^{#2}}}}
\providecommand{\ddt}[1][]{\dd[{#1}]t}
\providecommand{\dder}[2]{\frac{\d #1}{\d #2}}
\providecommand{\pder}[3][]{\frac{\partial^{#1} #2}{\partial #3}}
\providecommand{\pospart}[1]{\left[#1\right]^+}
\providecommand{\pow}[1]{\operatorname{pow}^{#1}}
\providecommand{\powq}[2]{{#2}^{#1}}
\renewcommand{\Re}{\operatorname{re}}
\providecommand{\reg}{{\operatorname{reg}}}
\providecommand{\sng}{{\operatorname{sng}}}
\providecommand{\sign}{\operatorname{sign}}
\providecommand{\supp}{\operatorname{spt}}
\providecommand{\support}{\supp}
\providecommand{\trace}{\operatorname{trace}}
\providecommand{\wiener}{\operatorname{W}}
\providecommand{\wasserstein}[2][]{\operatorname{Wass}^{#2}_{#1}}
\providecommand{\transpose}{{\boldsymbol\intercal}}
\providecommand{\transinner}{{}^\transpose\,}
\providecommand{\Transpose}[1]{\ensuremath{{#1}^{\transpose}}}
\providecommand{\Transposegeo}[1]{\ensuremath{#1}^*}
\providecommand{\Transposevec}[1]{\Transpose{\vec{#1}}}
\providecommand{\transposevec}[1]{\Transposevec{#1}}
\providecommand{\transnumvec}[1]{\Transpose{\numvec{#1}}}
\providecommand{\Transposemat}[1]{\Transpose{\mat{#1}}}
\providecommand{\transposemat}[1]{\Transposemat{#1}}
\providecommand{\Transinverse}[1]{\Transpose{{{#1}^{-}}}}
\providecommand{\rowvec}[1]{\ensuremath{\vecof{#1}}}
\providecommand{\colvec}[1]{\ensuremath{\Transpose{\rowvec{#1}}}}
\providecommand{\discolvec}[1]{\ensuremath{\begin{bmatrix}#1\end{bmatrix}}}
\providecommand{\rowvecdotsfromto}[3]{\rowvec{{#1}_{#2},\dotsc,{#1}_{#3}}}
\providecommand{\colvecdotsfromto}[3]{\colvec{{#1}_{#2},\dotsc,{#1}_{#3}}}
\providecommand{\discolvecdotsfromto}[3]{\discolvec{{#1}_{#2}\\\vdots\\{#1}_{#3}}}
\providecommand{\disrowvec}[1]{\ensuremath{\begin{bmatrix}#1\end{bmatrix}}}
\providecommand{\dismatrix}[1]{\begin{bmatrix}#1\end{bmatrix}}
\providecommand{\dismatii}[4][1]{
\dismatrix{
{#2}_{#1#1} &\dotso &{#2}_{#1#3}
\\
\vdots &\ddots &\vdots
\\
{#2}_{#4#1} &\dotso &{#2}_{#4#3}
}
}
\providecommand{\dismatiu}[4][1]{
\dismatrix{
{#2}_{#1}^{#1} &\dotso &{#2}_{#1}^{#3}
\\
\vdots &\ddots &\vdots
\\
{#2}_{#4}^{#1} &\dotso &{#2}_{#4}^{#3}
}
}
\providecommand{\mint}{\ensuremath{{\text{\bf{-\!}-}}\!\!\!\!\!\!\int}}
\providecommand{\smint}{\ensuremath{{\text{\textbf{-}}}\!\!\!\!\int}}
\providecommand{\lap}{\ensuremath{\Delta}}
\providecommand{\Harm}[1]{\ensuremath{(-\Delta)^{#1}}
\providecommand{\normal}{\vec n}
\providecommand{\normalto}[1][{}]{\ensuremath{\normal_{#1}}}
\providecommand{\normalder}[1][{}]{\ensuremath{\normalto[#1]}\inner\grad}
\providecommand{\Dnormale}[1]{\ensuremath{\frac{\partial#1}{\partial\nu}}}
\providecommand{\DDnormale}[2]{\ensuremath{\frac{\partial^{#1}#2}{\partial\nu^{#1}}}}
\providecommand{\compose}[2]{\ensuremath{#1\,\circ\:#2}}
\providecommand{\convolution}{\ast}
\providecommand{\convolve}[2]{\ensuremath{#1\convolution#2}}
\providecommand{\pushforward}{\#}
\providecommand{\Hess}{\ensuremath{\D^2}}
\providecommand{\intersected}{\ensuremath{\cap}}
\providecommand{\meet}{\intersected}
\providecommand{\intersection}[1]{\ensuremath{\bigcap}_{#1}}
\providecommand{\united}{\ensuremath{\cup}}
\providecommand{\join}{\united}
\providecommand{\union}[1]{\ensuremath{\bigcup}_{#1}}
\providecommand{\unioni}[2]{\ensuremath{\union{#2}{#1}_{#2}}}
\providecommand{\unionsi}[3]{\ensuremath{\union{#2\in{#3}}{#1}_{#2}}}
\providecommand{\unionsu}[3]{\ensuremath{\union{#2\in{#3}}{#1}^{#2}}}
\providecommand{\intersi}[3]{\ensuremath{\intersection{#2\in{#3}}{#1}_{#2}}}
\providecommand{\unionifromto}[3]{\ensuremath{\union{#1=#2}^{#3}}}
\providecommand{\interifromto}[3]{\ensuremath{\intersection{#1=#2}^{#3}}}
\providecommand{\unionsifromto}[4]{\ensuremath{\unionifromto{#2}{#3}{#4}{#1}_{#2}}}
\providecommand{\intersifromto}[4]{\ensuremath{\interifromto{#2}{#3}{#4}{#1}_{#2}}}
\providecommand{\Parti}[1]{\ensuremath{\cP\left(#1\right)}
\renewcommand{\Parti}[1]{\ensuremath{2^{#1}}}
\providecommand{\complement}[1]{\ensuremath{{#1}^\mathrm{c}}}
\renewcommand{\complement}[1]{\ensuremath{{#1}^\mathrm{c}}}
\providecommand{\symdiff}{\ensuremath{\triangle}}
\providecommand{\PROP}[1]{\ensuremath{\boldsymbol #1}
\renewcommand{\vec}[1]{\ensuremath{\boldsymbol{#1}}}
\providecommand{\geovec}[1]{\vec{#1}}
\providecommand{\geomat}[1]{\vec{#1}}
\providecommand{\arrowvec}[1]{\ensuremath{\overrightarrow{#1}}}
\providecommand{\dotvec}[1]{\ensuremath{\dot{\vec{#1}}}}
\providecommand{\tildevec}[1]{\ensuremath{\widetilde{\vec{#1}}}}
\providecommand{\mat}[1]{\geomat{#1}}
\providecommand{\num}[1]{\ensuremath{\mathsf{#1}}}
\providecommand{\numvec}[1]{{\vec{\num{#1}}}}
\providecommand{\nummat}[1]{\numvec{\MakeUppercase{#1}}}
\providecommand{\dotnumvec}[1]{\ensuremath{\dot{\numvec{#1}}}}
\providecommand{\transnumvec}[1]{\Transpose{\numvec{#1}}}
\providecommand{\Prob}[1][]{\ensuremath{\operatorname{Prob_{#1}}}}
\providecommand{\Probcond}[2]{\Prob\qp{{#1}\conditionalto{#2}}}
\providecommand{\prob}{\ensuremath{P}}
\providecommand{\probof}[1]{\Prob\qp{#1}}
\providecommand{\probofevent}[1]{\Prob\eventof{#1}}
\providecommand{\law}[1]{\ensuremath{\operatorname{law}_{#1}}\xspace}
\providecommand{\cdf}[1]{\ensuremath{\operatorname{cdf}_{#1}}\xspace}
\providecommand{\distrib}[1]{\operatorname{dtrb}_{#1}}
\providecommand{\pdf}[1]{\ensuremath{\operatorname{pdf}_{#1}}\xspace}
\providecommand{\density}[1]{\pdf{#1}}
\providecommand{\mgf}[1]{\operatorname{mgf}_{#1}}
\providecommand{\erf}{\ensuremath{\operatorname{erf}}\xspace}
\providecommand{\chf}[1]{\ensuremath{\operatorname{chf}_{#1}}\xspace}
\providecommand{\straton}{\circ}
\providecommand{\covariance}{\ensuremath{\operatorname{cov}}}
\providecommand{\Cov}{\covariance}
\providecommand{\cov}{\covariance}
\providecommand{\expectation}{\ensuremath{\operatorname{E}}}
\providecommand{\EX}{\expectation}
\providecommand{\QE}[2][]{\expectation_{#1}\qb{#2}}
%
\providecommand{\variance}{\ensuremath{\operatorname{var}}}
\providecommand{\var}{\variance}
\providecommand{\Var}{\variance}
\providecommand{\RV}[1]{\cX(#1)}
\providecommand{\ndistr}{\operatorname{N}}
\providecommand{\gaussrv}[2]{\ensuremath{\ndistr\left({#1},{#2}\right)}}
\providecommand{\uniformrv}[2]{\ensuremath{\operatorname{U}\qp{#1,#2}}}
\providecommand{\boundary}{\partial}
\providecommand{\Ast}{\ensuremath{{\ast\ast}}}
\providecommand{\CC}{\ensuremath{\operatorname C}\xspace
\providecommand{\HH}{\ensuremath{\operatorname H}\xspace}
\providecommand{\LL}{\ensuremath{\operatorname L}\xspace}
\providecommand{\VV}{\ensuremath{\operatorname V}\xspace}
\providecommand{\WW}{\ensuremath{\operatorname W}\xspace}
\providecommand{\dual}[1]{\ensuremath{{#1}'}}
\providecommand{\cont}[1]{\ensuremath{\CC^{#1}}}
\providecommand{\contc}[1]{\ensuremath{\cont{#1}_{\mathrm{c}}}}
\providecommand{\leb}[1]{\ensuremath{\LL_{#1}}}
\providecommand{\lebloc}[1]{\ensuremath{{\LL^{\mathrm{loc}}_{#1}}}}
\providecommand{\sob}[2]{\ensuremath{{\smash\WW}^{#1}_{#2}}}
\providecommand{\sobz}[2]{\ensuremath{{\overset{\smash{\scriptscriptstyle\circ}}\WW}{}^{#1}_{#2}}}
\providecommand{\sobaz}[2]{\ensuremath{{\overset{\smash{\scriptscriptstyle\oslash}}{\smash\WW}}{}^{#1}_{#2}}}
\providecommand{\sobh}[1]{\ensuremath{\HH^{#1}}}
\providecommand{\sobhz}[1]{\sobh{#1}_0}
\providecommand{\sobhaz}[1]{\sobh{#1}_{\varoslash}}
\providecommand{\Lip}{\ensuremath{\operatorname{Lip}}}
\providecommand{\poly}[1]{\ensuremath{\rP}^{#1}}
\providecommand{\Symmatrices}[1]{\ensuremath{\operatorname{Sym}{(\R{d\times d})}}}
\providecommand{\mesh}[1]{\ensuremath{\mathscript{#1}}}
\providecommand{\fespace}{\rV}
\providecommand{\fezerospace}{\ensuremath{\smash{\mathring\fespace}}}
\providecommand{\fezeroaveragespace}{\ensuremath{{\fespace}{\scriptscriptstyle\oslash}}}
\providecommand{\fes}[1]{\ensuremath{\fespace^{#1}}}
\providecommand{\fez}[1]{\ensuremath{\fezerospace^{#1}}}
\providecommand{\feaz}{\fezeroaveragespace}
\providecommand{\fesh}{\ensuremath{\fespace_h}}
\providecommand{\fezh}{\ensuremath{\fezerospace_h}}
\providecommand{\feaz}{\ensuremath{\fezeroaveragespace_h}}
\providecommand{\fe}[2][]{\ensuremath{\uppercase{#2}_{#1}}
\providecommand{\EOC}{\ensuremath{\operatorname{EOC}}\xspace}
\providecommand{\Forall}{\:\forall\:}
\providecommand{\Exists}{\:\exists\:}
\providecommand{\Foreach}{\quad\Forall}
\providecommand{\Forsome}{\quad\text{for some }}
\usepackage{amscd}
\providecommand{\relation}[1]{\,{#1}\,}
\providecommand{\relationnot}[1]{\not\!{#1}\,\,}
\providecommand{\divides}{\mid}
\providecommand{\dividesnot}{\nmid}
\providecommand{\ideq}{\equiv}
\providecommand{\compactsubset}{\Subset}
\providecommand{\funk}[3]{\ensuremath{#1:#2\to#3}}
\providecommand{\funkmapsto}[6][]{\ensuremath{{#2}:{#4}\ni{#3}\mapsto{#5}\in{#6#1}}}
\providecommand{\dfunkmapsto}[6][]{\ensuremath{
\begin{array}{rccl}
{#2}: & {#4} & \to & {#6}
\\
& {#3} &\mapsto & {#5\text{\ #1}}
\end{array}\quad}}
\providecommand{\setmapsto}[4]{\ensuremath{{#2}\ni{#1}\mapsto{#3}\in{#4}}}
\providecommand{\dsetmapsto}[4]{\ensuremath{
\begin{array}{rccl}
& {#2} & \to & {#4}
\\
& {#1} &\mapsto & {#3}
\end{array}\quad}}
\providecommand{\onetoone}[3]{\ensuremath{#1:#2\leftrightarrows#3}}
\renewcommand{\implies}{\ensuremath{\:\Rightarrow\:}\xspace}
\providecommand{\notimplies}{\ensuremath{\:\Arrownot\Rightarrow\:}\xspace}
\providecommand{\THEN}{\implies}
\providecommand{\LIF}{\ensuremath{\:\Leftarrow\:}\xspace}
\providecommand{\equivale}{\ensuremath{\:\Leftrightarrow\:}\xspace}
\providecommand{\IFF}{\equivale}
\providecommand{\LIFF}{\equivale}
\providecommand{\contradiction}{\ensuremath{\:\lightning\:}\xspace}
\providecommand{\downto}{\ensuremath{\searrow}}
\providecommand{\upto}{\ensuremath{\nearrow}}
\providecommand{\Wrightarrow}{\ensuremath{{\rightharpoonup}}}
\providecommand{\convergent}[1]{
\ensuremath{
\begin{CD}
@>{\scriptscriptstyle #1}>>
\end{CD}
}
}
\providecommand{\Convergent}[2]{
\ensuremath{
\begin{CD}
@>{\scriptscriptstyle #1}>{\scriptscriptstyle #2}>
\end{CD}
}
}
\providecommand{\weaklyconvergent}[1]{
\ensuremath{\begin{CD}
@>{\scriptscriptstyle #1}
>{\scriptscriptstyle{\mathrm{weakly}}}>
\end{CD}
}
}
\providecommand{\toas}[1]{\ensuremath{\begin{CD}@>{\phantom{:}}>{#1}>\end{CD}}}
\providecommand{\equipotent}{\rightleftarrows}
\providecommand{\assignvalue}{\ensuremath{\mapsfrom}}
\providecommand{\imbedded}{{\ensuremath{\hookrightarrow}}}
\providecommand{\embedded}{{\ensuremath{\hookrightarrow}}}
\providecommand{\embedsin}{{\ensuremath{\hookrightarrow}}}
\def\lhook\joinrel\lhook\joinrel\rightarrow{\lhook\joinrel\lhook\joinrel\rightarrow}
\providecommand{\compactly}[1]{
\ensuremath{\overset{\mathrm{comp}}{#1}}}
\providecommand{\compactlyimbedded}{\compactly\embedded}
\providecommand{\compactlyembedded}{\compactlyimbedded}
\providecommand{\compactlyembedsin}{\compactlyimbedded}
\providecommand{\finerthan}{\prec}
\providecommand{\rougherthan}{\succ}
\providecommand{\coarserthan}{\succ}
\providecommand{\restriction}[2]{\left.#1\right|_{#2}}
\renewcommand{\restriction}[2]{\left.#1\right|_{#2}}
\providecommand{\restrictionff}[2]{\restriction{#1}{#2}}
\providecommand{\restrictionqf}[2]{\qb{#1}_{#2}}
\providecommand{\restrictionfq}[2]{\restriction{#1}{\qb{#2}}}
\providecommand{\restrictionqq}[2]{\qb{#1}_{\qb{#2}}}
\providecommand{\evalat}[2]{\qb{#1}_{#2}}
\providecommand{\evaldiff}[3]{\qb{#1}^{#2}_{#3}}
\providecommand{\punto}{\ensuremath{\bullet}\ }
\providecommand{\hhat}[1]{\hat{\hat{#1}}}
\providecommand{\ccheck}[1]{\check{\check{#1}}}
\providecommand{\aka}[1]{(also known as {#1})\xspace}
\providecommand{\ode}{ODE\xspace}
\providecommand{\odes}{ODE's\xspace}
\providecommand{\ivp}{IVP\xspace}
\providecommand{\ivps}{IVP's\xspace}
\providecommand{\DNE}{\ensuremath{\text{ DNE}}\xspace}
\providecommand{\CBS}{Cauchy--Bunyakovskii--Schwarz inequality\xspace}
\providecommand\bs{\char '134}
\providecommand{\Program}[1]{\textsf{#1}\xspace}
\providecommand{\Source}[1]{\textsf{#1}\xspace}
\providecommand{\Algoname}[1]{\ensuremath{\text{\textsf{#1}\xspace}}}
\providecommand{\octaveline}[1]{
\begin{center}
\begin{minipage}{\textwidth}
\begin{tikzpicture}
\node[fill=c!20!g]{
\begin{minipage}{\textwidth}
\Verb+octave:**> #1+
\end{minipage}
};
\end{tikzpicture}
\end{minipage}
\end{center}
}
\providecommand{\texcommand}[1]{\texttt{\bs{#1}}}
\providecommand{\codename}[1]{\texttt{\bfseries #1}\xspace}
\providecommand{\filename}[1]{\texttt{#1}\xspace}
\providecommand{\codevarname}[1]{\texttt{#1}\xspace}
\providecommand{\sterling}{\ensuremath{\pounds}\xspace}
\providecommand{\euro}{\textgreek{\euro}}
\providecommand{\euros}{\ensuremath{\text{\textgreek{\euro}}}\xspace}
\providecommand{\matlab}{\Program{Matlab\registered}}
\providecommand{\octave}{\Program{Octave}}
\providecommand{\scilab}{\Program{Scilab}}
\providecommand{\perl}{\Program{Perl}}
\providecommand{\albert}{\Program{ALBERTA}}
\providecommand{\alberta}{\Program{ALBERTA}}
\providecommand{\gltools}{\Program{GL Tools}}
\providecommand{\linux}{\Program{Linux}}
\providecommand{\caesar}{\Program{c\ae sar}}
\providecommand{\ingenious}{Ingenio{\textsc{US}}\xspace}
\providecommand{\codeprint}[1]{
\par
\begin{center}
\framebox{Printout of file \filename{#1}}
\\
\lstinputlisting{#1}
\end{center}
\par
}
\providecommand{\indexit}[1]{#1\index{#1}}
\providecommand{\secsymbol}{\S}
\providecommand{\secsymbols}{\S\S}
\providecommand{\secref}[1]{\secsymbol\ref{#1}}
\providecommand{\charef}[1]{Ch.~\ref{#1}}
\providecommand{\equa}[1]{(#1)}
\providecommand{\proveimplies}[2]{\par[{#1}\implies{#2}]}
\providecommand{\proveequivale}[2]{\par[{#1}\equivale{#2}]}
\providecommand{\eqncomment}[1]{\ensuremath{\qquad\qp{\text{{#1}}}}}
\providecommand{\because}[1]{\ensuremath{\qquad\text{({#1})}}}
\renewcommand{\because}[1]{\ensuremath{\qquad\text{({#1})}}}
\providecommand{\dueto}[1]{\ensuremath{\text{\tiny{(#1)}}\quad}}
\providecommand{\underbracedescribe}[2]{\ensuremath{\underbrace{#1}_{\text{#2}}}}
\providecommand{\underbracedefine}[2]{\ensuremath{\underbrace{#1}_{\displaystyle=:#2}}}
\newenvironment{Hint}{\par\textit{Hint}.}{\par}
\providecommand{\ListParameters}{}
\renewcommand{\ListParameters
{
\setlength{\topsep}{0em}
\setlength{\leftmargin}{0em}
\setlength{\itemsep}{0ex}
\setlength{\parsep}{.5ex}
\setlength{\itemindent}{\labelsep}
\addtolength{\itemindent}{\labelwidth}
}
\newcounter{tmpcounter}
\newcounter{LetterListItem}
\renewcommand{\theLetterListItem}{(\alph{LetterListItem})}
\newenvironment{LetterList}%
{
\begin{list}%
{\theLetterListItem\ }%
{\usecounter{LetterListItem}
\ListParameters
}
}%
{\end{list}}
\newcounter{NumberListItem}
\renewcommand{\theNumberListItem}{\arabic{NumberListItem}}
\newenvironment{NumberList}%
{
\begin{list}%
{\theNumberListItem.\ }%
{\usecounter{NumberListItem}%
\ListParameters
}
}%
{\end{list}}
\newcounter{QuestionListItem}
\renewcommand{\theQuestionListItem}{\textbf{Question \arabic{QuestionListItem}}}
\newenvironment{QuestionList}%
{
\begin{list}%
{\theQuestionListItem.\ }%
{\usecounter{QuestionListItem}%
\ListParameters
}
}%
{\end{list}}
\newcounter{RomanListItem}
\renewcommand{\theRomanListItem}{(\roman{RomanListItem})}
\newenvironment{RomanList}%
{
\begin{list}%
{\theRomanListItem\ }%
{\usecounter{RomanListItem}
\ListParameters
}
}%
{\end{list}}
\newcounter{StepsItem}
\newenvironment{Steps}%
{
\begin{list}%
{Step \theStepsItem.\ }%
{\usecounter{StepsItem}%
\ListParameters
}
}%
{\end{list}}
\providecommand{\claim}[1]{\textsl{#1}}
%
%
\providecommand{\grad}{\nabla}
\renewcommand{\grad}{\nabla}
\usepackage{amsthm}
%
\ifthenelse{\boolean{isthesis}}{%
\setcounter{secnumdepth}{1}%
}{
\setcounter{secnumdepth}{2}
}
%
%
\providecommand{\ListParameters}{}
\renewcommand{\ListParameters}
{
\setlength{\topsep}{0em}
\setlength{\leftmargin}{0em}
\setlength{\itemsep}{0ex}
\setlength{\parsep}{.5ex}
\setlength{\itemindent}{\labelsep}
\addtolength{\itemindent}{\labelwidth}
}
%
%
\newtheoremstyle{plain
{
{
{\mdseries\slshape
{\parindent
{\bfseries
{.
{.5em
{
\newtheoremstyle{note
{
{
{
{\parindent
{\bfseries
{.
{.5em
{
\newtheoremstyle{claim
{
{
{\mdseries\slshape
{
{\bfseries
{
{.5em
{
\newtheoremstyle{exercise
{
{
{
{
{\bfseries
{.
{1em
{
\newtheoremstyle{break
{
{
{
{
{\bfseries
{.
{\newline
{
\providecommand{\ObsName}{Remark
\providecommand{\RemName}{Remark
\providecommand{\NotName}{Notation
\providecommand{\BFNName}{Big~Fat~Note
\providecommand{\DefName}{Definition
\providecommand{\ExaName}{Example
\providecommand{\TheName}{Theorem
\providecommand{\LemName}{Lemma
\providecommand{\ProName}{Proposition
\providecommand{\CorName}{Corollary
\providecommand{\PbmName}{Problem
\providecommand{\HypName}{Hypothesis
\providecommand{\AlgName}{Algorithm
\providecommand{\ExeName}{Exercise
\providecommand{\SolName}{Solution
\providecommand{\ClaName}{Claim
\providecommand{\EsyName}{Essay
\providecommand{\Proofname}{Proof
\ifthenelse{\boolean{isthesis}}{%
\providecommand{\Thecounter}{The}
}{%
\providecommand{\Thecounter}{subsection}
}
\providecommand{\pdfformat}[1]{
\provideboolean{pdfoutput}
\setboolean{pdfoutput}{#1
\ifthenelse{\boolean{pdfoutput}}{
\typeout{using pdf}
\providecommand{\graphext}{pdf}
\renewcommand{\graphext}{pdf}
\providecommand{\graphextex}{pdf_t}
\renewcommand{\graphextex}{pdf_t}
}{
\typeout{using eps}
\usepackage[dvips]{graphicx,xcolor}
\providecommand{\graphext}{eps}
\renewcommand{\graphext}{eps}
\providecommand{\graphextex}{eps_t}
\renewcommand{\graphextex}{eps_t}
}
\usepackage{epsfig}
\usepackage{tikz}
\usepackage{rotating}
\definecolor{SussexFlint}{rgb}{.00,.19,.21}
\definecolor{SussexGrey}{rgb}{.51,.58,.49}
\definecolor{SussexOrange}{rgb}{.94,.29,.00}
\definecolor{SussexYellow}{rgb}{1.00,.73,.00}
\definecolor{SussexRed}{rgb}{.94,.01,.49}
\definecolor{SussexPurple}{rgb}{.48,.06,.44}
\definecolor{SussexGreen}{rgb}{.00,.58,.46}
\definecolor{SussexBlue}{rgb}{.00,.58,.65}
\colorlet{a}{SussexOrange}
\colorlet{b}{SussexRed}
\colorlet{c}{SussexYellow}
\colorlet{d}{SussexPurple}
\colorlet{e}{SussexGreen}
\colorlet{f}{SussexBlue}
\colorlet{g}{white
\colorlet{h}{SussexGrey
\colorlet{i}{black
\colorlet{j}{SussexFlint}
\newcommand{\mausDarkColorTheme}{
\colorlet{a}{SussexYellow!50!yellow}
\colorlet{b}{SussexGreen!50!green}
\colorlet{c}{SussexBlue
\colorlet{d}{SussexOrange!50!yellow}
\colorlet{e}{SussexRed!50!red}
\colorlet{f}{SussexPurple!50!magenta}
\colorlet{g}{black
\colorlet{h}{SussexFlint!50!black}
\colorlet{i}{white
\colorlet{j}{SussexGrey}
}
}
\providecommand{\solution}{\textbf{\SolName.}\xspace}
\newenvironment{sol}{\par\solution}{\par}
\newcounter{phantombox}[enumi
\provideboolean{showphantoms}
\renewcommand{\thephantombox}{\roman{phantombox}}
\newcommand{\phantombox}[1]{\stepcounter{phantombox}
\ensuremath{\boxed{
\ifthenelse
{\boolean{showphantoms}}
{#1^{\phantom{\textup{(\thephantombox)}}}}
{\phantom{#1}^{\textup{(\thephantombox)}}}
}
}
}
\newcommand{\phantomedinput}[1]{\setboolean{showphantoms}{false}\input{#1}}
\newcommand{\ghostbusterinput}[1]{\setboolean{showphantoms}{true}\input{#1}}
\provideboolean{hidesolution}
\newcommand{\consolution}[1]{
\ifthenelse{
\boolean{hidesolution}
}{
}{
{\par \small {\solution}\ #1\par\ \\[5pt]}}
}
\provideboolean{showmarks}
\renewcommand{\marks}[1]{
\ifthenelse{\boolean{showmarks}}{\marginpar{{\tiny [$#1$ marks]}}}{}}
\newcommand{\ifthenelse{\boolean{hidesolution}}{\newpage}{}}{\ifthenelse{\boolean{hidesolution}}{\newpage}{}}
\newcommand{\setcounter{equation}0\item}{\setcounter{equation}0\item}
\usepackage{natbib}
\usepackage[utf8]{inputenc}
\usepackage{tikz}
\providecommand{\nlop}[1]{\cN[#1]}
\providecommand{\nlfunk}{F}
\providecommand{\Dnlop}{\mat N}
\providecommand{\numPhizero}{\circ{\smash{\numvec \Phi}}}
\providecommand{\uoo}{\circ{\numvec u}}
\numberwithin{equation}{section}
\setboolean{showtodo}{true
\setboolean{shownotes}{true
\setboolean{showchanges}{true
\setboolean{usemathrsfs}{false
\setlength{\parindent}{12pt}
\providecommand{\fenics}{FEniCS\xspace}
\providecommand{\petsc}{PETSc\xspace}
\providecommand{\paraview}{ParaView\xspace}
\newcommand{Gnuplot\xspace}{Gnuplot\xspace}
\newcommand{DOLFIN\xspace}{DOLFIN\xspace}
\author{Omar Lakkis}
\address{ Omar Lakkis\newline
Department of Mathematics\newline
University of Sussex\newline
Brighton, England\newline
GB-BN1 9RF} \curraddr{}
\email{\linkedemail{<EMAIL>}}
\urladdr{\linkedurl{http://www.maths.sussex.ac.uk/Staff/OL}}
\author{Tristan Pryer}
\address{Tristan Pryer\newline
School of Mathematics, Statistics and Actuarial Sciences\newline
University of Kent\newline
Canterbury\newline
UK-CT2 7NF, United Kingdom}
\curraddr{}
\email{\linkedemail{<EMAIL>}}
\title[FEM for nonlinear elliptic problems]{A finite element method
for nonlinear elliptic problems}
\date{\today}
\begin{document}
\maketitle
\begin{abstract}
We present a continuous finite element method for some examples of
fully nonlinear elliptic equation. A key tool is the discretisation
proposed in Lakkis \& Pryer (2011) allowing us to work directly on
the strong form of a linear PDE. An added benefit to making use of
this discretisation method is that a \emph{recovered (finite
element) Hessian} is a biproduct of the solution process. We build
on the linear basis and ultimately construct two different
methodologies for the solution of second order fully nonlinear PDEs.
Benchmark numerical results illustrate the convergence properties of
the scheme for some test problems as well as the Monge--Ampère
equation and the Pucci equation.
\end{abstract}
\section{Introduction}
\label{sec:intro}
Fully nonlinear PDEs arise in many areas, including differential
geometry (the \MA equation), mass transportation (the
Monge--Kantorovich problem), dynamic programming (the Bellman
equation) and fluid dynamics (the geostrophic equations). The
computer approximation of the solutions of such equations is thus an
important scientific task. There are at least three main difficulties
apparent to someone attempting to derive numerical methods for fully
nonlinear equations: first, the strong nonlinearity on the highest
order derivative which generally precludes a variational formulation,
second, a fully nonlinear equation does not always admit a classical
solution, even if the problem data is smooth, and the solution has to
sought in a generalised sense (e.g., viscosity solutions), which is
bound to slow down convergence rates, and third, a common problem in
nonlinear solvers, the exact solution may not be unique and
constraints, such as convexity requirements must be included in the
constraints to ensure uniqueness.
Regardless of the problems, the \emph{numerical approximation of fully
nonlinear second order elliptic equations}, as described in
\cite{CaffarelliCabre:1995}, have been the object of considerable recent
research, particularly for the case of \MA of which
\cite{OlikerPrussner:1988, LoeperRapetti:2005, DeanGlowinski:2006,
FengNeilan:09:vanishing, Oberman:2008, Awanou:10, DavydovSaeed:12:techreport,
BrennerGudiNeilanSung:11, Froese:11} are selected examples.
For more general classes of fully nonlinear equations some methods
have been presented, most notably, at least from a theoretical view
point, in \cite{Bohmer:2008} where the author presents a $\cont{1}$
finite element method shows stability and consistency (hence
convergence) of the scheme, following a classical ``finite
difference'' approach outlined by \cite{Stetter:1973} which requires a
high degree of smoothness on the exact solution. From a practical
point of view this approach presents difficulties, in that the
$\cont{1}$ finite elements are hard to design and complicated to
implement, in \cite{DavydovSaeed:10} a useful overview of
Bézier-Bernestein splines in two spatial dimensions is provided and a
full implementation in \cite{DavydovSaeed:12:techreport}. Similar
difficulties are encountered in finite difference methods and the
concept of \emph{wide-stencil} appears to be useful, for example by
\cite{KuoTrudinger:92,KuoTrudinger:05,Oberman:2008,Froese:11}.
In \cite{FengNeilan:09:mixed,FengNeilan:09:vanishing,Awanou:10}
the authors give a method in which they approximate the general second
order fully nonlinear PDE by a sequence of fourth order quasilinear
PDEs. These are quasilinear biharmonic equations which are discretised
via mixed finite elements, or using high-regularity elements such as
splines. In fact for the \MA equation, which admits two solutions, of
which one is convex and another concave, this method allows for the
approximation of both solutions via the correct choice of a
parameter. On the other hand although computationally less expensive
than $\cont{1}$ finite elements (an alternative to mixed methods for
solving the biharmonic problem), the mixed formulation still results
in an extremely large algebraic system and the lack of maximum
principle for general fourth order equations makes it hard to apply
vanishing viscosity arguments to prove convergence. A somewhat
different approach, based on $C^0$-penalty, has been recently proposed
by \cite{BrennerGudiNeilanSung:11}, as well as ``pseudo time'' one by
\cite{Awanou:11}.
It is worth citing also a \emph{least square} approach described by
\cite{DeanGlowinski:2006}. This method consists in minimising the
mean-square of the residual, using a Lagrange multiplier method. Also
here a fourth order elliptic term appears in the energy.
In this paper, we depart from the above proposed methods and explore a
more ``direct'' approach by applying the \emph{nonvariational finite
element method}, introduced in \cite{LakkisPryer:2011}, as a solver
for the Newton iteration directly derived from the PDE. To be more
specific, consider the following model problem
\begin{equation}
\label{eq:modelproblem}
\nlop{u} := \nlfunk(\Hess u) - f = 0
\end{equation}
with homogeneous Dirichlet boundary conditions where $\funk f\W\reals$
is prescribed function and $\funk F\symm\reals$ is a real-valued
algebraic function of symmetric matrixes, which provides an elliptic
operator in the sense of \cite{CaffarelliCabre:1995}, as explained
below in Definition \ref{def:ellipticity}. The method we propose,
consists in applying a Newton's method, given below by equation
\eqref{eq:linearisedequation} of the PDE \eqref{eq:modelproblem},
which results in a sequence of linear nonvariational elliptic PDEs
that fall the framework of the nonvariational finite element method
(NVFEM) proposed in \cite{LakkisPryer:2011}. The results in this
paper are computational, so despite not having a complete proof of
convergence, we test our algorithm various problems that are
specifically constructed to be well posed. In particular, we test our
method on the \MA problem, which is the de-facto benchmark for
numerical methods of fully nonlinear elliptic equations. This is in
spite of \MA having an extra complication, which is conditional
ellipiticity (the operator is elliptic only if the function is convex
or concave. A crucial, empirically observed feature of our method is
that the convexity (or concavity) is automatically preserved if one
uses $\poly2$ elements or higher. For $\poly1$ elements this is not
true and the scheme must be stabilized by reenforcing convexity (or
concavity) at each timestep. This was achieved in \cite{Pryer:2010}
using a semidefinite programming method. In a different spirit, but
somewhat reminiscent, a stabilization procedure was obtained in
\cite{BrennerGudiNeilanSung:11} by adding a penalty term.
The rest of this paper is set out as follows. In
\S\ref{sec:notation_and_discretisation} we introduce some notation,
the model problem, discuss its ellipticity and Newton's method, which
yields a sequences of nonvariational linearised PDE's. In
\S\ref{sec:ndfem} we review of the nonvariational finite element
method proposed in \cite{LakkisPryer:2011} and apply it to discretise
the nonvariational linearised PDE's in Newton's method. In
\S\ref{sec:unconstrained-fnl-pde} we numerically demonstrate the
performance of our discretisation on a class of fully nonlinear PDE,
those that are elliptic and well posed without constraining our
solution to a certain class of functions. In \S\ref{sec:ma} we turn
to conditionally elliptic problems by dealing with the prime example
of such problems, i.e., \MA. We apply the discretisation to the \MA
equation making use of the work \cite{Aguilera:2008} to check
\emph{finite element convexity} is preserved at each iteration.
Finally in \S\ref{sec:Pucci} we address the approximation of Pucci's
equation, which is another important example of fully nonlinear
elliptic equation.
All the numerical experiments for this research, were carried out
using the DOLFIN\xspace interface for \fenics \cite{LoggWells:2010} and
making use of Gnuplot\xspace and \paraview for the graphics.
\section{Notation}
\label{sec:notation_and_discretisation}
\subsection{Functional set-up}
Let $\W\subset \R{d}$ be an open and bounded Lipschitz domain. We
denote $\leb{2}(\W)$ to be the space of square (Lebesgue) integrable
functions on $\W$ together with {its} inner product
$\ltwop{v}{w} := \int_\W v w$ and norm $\Norm{v} :=
\Norm{v}_{\leb2(\W)} = \ltwop{v}{v}^{1/2}$.
We denote by
$\duality{v}{w}$ the action of a distribution $v$ on the function
$w$.
We use the convention that the derivative $\D u$ of a function
$u:\W\to\reals$ is a row vector, while the gradient of $u$, $\nabla u$
is the derivatives transpose (an element of $\reals^d$, representing
$\D u$ in the canonical basis). Hence
\begin{equation}
\nabla u
=
\Transpose{\left(\D u\right)}.
\end{equation}
For second derivatives, we follow the common innocuous abuse of
notation whereby the Hessian of $u$ is denoted as $\Hess u$ (instead
of the more consistent $\D\grad u$) and is represented by a $d\times d$
matrix.
The standard Sobolev spaces are ~\cite{Ciarlet:1978,Evans:1998}
\begin{gather}
\sobh{k}(\W)
:=
\sob{k}{2}(\W)
=
\ensemble{\phi\in\leb2(\W)}
{\sum_{\norm{\vec\alpha} \leq k}
\D^{\vec\alpha}\phi\in\leb2(\W)},
\\
\hoz := \text{closure of }\cont{\infty}_0(\W) \text{ in } \sobh{1}(\W)
\end{gather}
where $\vec\alpha = \{ \alpha_1,...,\alpha_d\}$ is a
multi-index, $\norm{\vec\alpha} = \sum_{i=1}^d\alpha_i$ and
derivatives $\D^{\vec\alpha}$ are understood in a weak sense.
We consider the case when the model problem \eqref{eq:modelproblem} is
uniformly elliptic in the following sense.
\begin{Defn}[{ellipticity \cite{CaffarelliCabre:1995}}]
\label{def:ellipticity}
The operator $\nlop\cdot$ in Problem (\ref{eq:modelproblem}) is
called \emph{elliptic on $\cC\subseteq\symm$} if and only if
for each $\geomat M\in\cC$ there exist $\Lambda\geq\lambda>0$, that
may depend on $\geomat M$ such that
\begin{equation}
\label{eqn:conditional-ellipticity-of-nl}
\lambda \sup_{\norm{\geomat \xi} = 1}\norm{\geomat N\geomat \xi}
\leq
\nlfunk(\geomat M+\geomat N) - \nlfunk(\geomat M)
\leq
\La \sup_{\norm{\geomat \xi} = 1}\norm{\geomat N\geomat \xi}
\Foreach \geomat N\in\symm.
\end{equation}
If the largest possible set $\cC$ for which
(\ref{eqn:conditional-ellipticity-of-nl}) is satisfied is a proper
subset of $\symm$ we say that $\nlop\cdot$ is \emph{conditionally elliptic}.
The operator $\nlop\cdot$ in Problem (\ref{eq:modelproblem}) is called
to be \emph{uniformly elliptic} if and only if for some
$\lambda,\La>0$, called \emph{ellipticity constants}, we have
\begin{equation}
\label{eq:ellipticity-of-nl}
\lambda \sup_{\norm{\geomat \xi} = 1}\norm{\geomat N\geomat \xi}
\leq
\nlfunk(\geomat M+\geomat N) - \nlfunk(\geomat M)
\leq
\La \sup_{\norm{\geomat \xi} = 1}\norm{\geomat N\geomat \xi}
\Foreach \geomat N,\geomat M\in\symm.
\end{equation}
\end{Defn}
If $\nlfunk$ is differentiable (\ref{eq:ellipticity-of-nl}) can be
obtained from conditions on the derivative of $\nlfunk$. A generic
$\geomat M\in\symm$ is written as
\begin{equation}
\geomat M =
\squmatmd md
;
\end{equation}
so the derivative of $\nlfunk$ in the direction $\geomat N$ is given by
\begin{equation}
\D\nlfunk(\geomat M)\geomat N
=
\frob{\nlfunk'(\geomat M)}{\geomat N}
\end{equation}
where the \emph{derivative matrix} $\nlfunk'(\geomat M)$ is defined by
\begin{equation}
\nlfunk'(\geomat M)
:=
\squmatmd{\partial\nlfunk(\geomat M)/\partial m}{d}
.
\end{equation}
Suppose $\nlfunk$ is differentiable. Then
(\ref{eqn:conditional-ellipticity-of-nl}) is satisfied if and only if
for each $\geomat M\in\cC$ there exists $\mu>0$ such that
\begin{equation}
\label{eq:ellipticity-for-derivative}
\Transpose{\geovec\xi}\nlfunk'(\geomat M)\geovec\xi
\geq
\mu
\norm{\geovec \xi}^2
\Foreach
\geovec\xi\in\R d.
\end{equation}
Furthermore $\cC=\symm$ and $\mu$ is independent of $\geomat M$
if and only if (\ref{eq:ellipticity-of-nl}) is satisfied.
\begin{Hyp}[smooth elliptic operator]
\label{hyp:smooth-elliptic-operator}
In the remainder of this paper we shall assume that
$\nlop\cdot$ is conditionally elliptic on $\cC$ and
\begin{equation}
\nlfunk \in \cont{1}(\cC).
\end{equation}
Unless otherwise stated we will also assume that $\cC=\symm$.
\end{Hyp}
\subsection{Newton's method}
\label{sec:newtons-method}
The smoothness assumption \ref{hyp:smooth-elliptic-operator} allows
to apply Newton's method to solve Problem (\ref{eq:modelproblem}).
Given the initial guess $u^0\in\cont2(\W)$, with $\Hess u^0\in\cC$, for
each $n\in\NO$, find $u^{n+1}\in\cont2(\W)$ with $\Hess u^{n+1}\in\cC$ such that
\begin{equation}
\label{eq:newtonsmethod}
\Dnlop{u^n}\left(u^{n+1} - u^n\right) = -\nlop{u^n},
\end{equation}
where $\Dnlop u$ indicates the (Fréchet) derivative, which is formally
given by
\begin{equation}
\label{eq:newtonsmethod2}
\begin{split}
{\Dnlop u}{v}
&=
\lim_{\epsilon\rightarrow 0}
\frac{\nlop{u + \epsilon v} - \nlop{u}}{\epsilon}
\\
&=
\lim_{\epsilon\rightarrow 0}
\frac{\nlfunk(\Hess u + \epsilon \Hess v) - \nlfunk(\Hess u)}{\epsilon}
\\
&=
\nlfunk'(\Hess u) : \Hess v,
\end{split}
\end{equation}
for each $v\in\cont2(\W)$.
Combining (\ref{eq:newtonsmethod}) and (\ref{eq:newtonsmethod2}) then
results in the following nonvariational sequence of linear PDEs. Given
$u^0$ for each $n\in\naturals_0$ find $u^{n+1}$ such that
\begin{equation}
\label{eq:system-of-nonvar-pdes}
\nlfunk'(\Hess u^n) : \Hess\left(u^{n+1}-u^n\right) = f - \nlfunk(\Hess u^n).
\end{equation}
The PDE (\ref{eq:system-of-nonvar-pdes}) comes naturally in a nonvariational
form. If we attempted to rewrite into a variational form, in order, say, to
apply a ``standard'' Galerkin method, we would introduce an advection term
which would depend on derivatives of $\nlfunk'$, \ie for generic $v,w$
\begin{equation}
\frob{\nlfunk'(\Hess v)}{\Hess w}
=
\div\qb{\nlfunk'(\Hess v) \nabla w}- \div\qb{ \nlfunk'(\Hess v) }\nabla w.
\end{equation}
where the matrix-divergence is taken row-wise:
\begin{equation}
\div\qb{\nlfunk'(\Hess v(\vec x))}
:=
\qp{
\sum_{i=1}^d \frac{\partial}{\partial x_i}
\qb{\matentry{F'}i1(\Hess v(\vec x))}
,
\dotsc
,
\sum_{i=1}^d \frac{\partial}{\partial x_i}
\qb{\matentry{F'}id(\Hess v(\vec x))}
}
\end{equation}
and the chain rule provides us, for each $\rangefromto j1d$, with
\begin{equation}
\sum_{i=1}^d \frac{\partial}{\partial x_i}
\qb{\matentry{F'}ij(\Hess v(\vec x))}
=
\sumifromto{k,l}1d
\partial_{k,l}\matentry{F'}ij
(\Hess v(\vec x))
\partial_{ikl}v(\vec x).
\end{equation}
This procedure is undesirable for many reasons. Firstly it requires
$F$ to be twice differentiable and it involves a third order
derivative of the functions $u^{n+1}$ and $u^n$ appearing in
(\ref{eq:newtonsmethod}). Moreover, the ``variational'' reformulation
could very well result in the problem becoming advection dominated and
unstable for conforming FEM, as was manifested in numerical examples
for the linear equation \cite[\S 4.2]{LakkisPryer:2011}. In order to
avoid these problems, we here propose the use of the nonvariational
finite element method described next.
\section[The NVFEM]{The nonvariational finite element method}
\label{sec:ndfem}
In \cite{LakkisPryer:2011} we proposed the \emph{nonvariational finite
element method} (NVFEM) to approximate the solution of problems of
the form (\ref{eq:system-of-nonvar-pdes}). We review here the NVFEM
and explain how to use it in combination with the Newton method to
derive a practical Galerkin method for the numerical approximation of
Problem (\ref{eq:modelproblem})'s solution.
\subsection{Distributional form of (\ref{eq:system-of-nonvar-pdes}) and generalised Hessian}
Let $\A \in \Le{\infty}(\W)^{d\times d}$ and for each $\geovec
x\in\W$, let $\A(\geovec x)\in\symm$, the space of bounded, symmetric,
positive definite, $d \times d$ matrices and $\funk f\W\reals$. The
\emph{Dirichlet linear nonvariational elliptic problem} associated
with $\A$ and $f$ is
\begin{equation}
\frob{\A}\Hess u=f
\AND
\restriction u{\boundary\W}=0.
\end{equation}
Testing this equation, and assuming $u\in\sobh2(\W)\cap\hoz$ such
that $\restriction{\grad u}{\boundary\W}\in\leb2(\boundary\W)$,
we may write it as
\begin{equation}
\begin{split}
\label{Problem}
\ltwop{\frob{\A}{\Hess u}}{\phi}
&= \ltwop{f}{\phi}
\qquad \Foreach
\phi\in\cont\infty_0(\W).
\end{split}
\end{equation}
To allow a Galerkin type discretisation of (\ref{Problem}), we need to
restrict the test functions $\phi$ to finite element function spaces
that are generally \emph{not subspaces} of $\sobh2(\W)$. So before
restricting, we need to extend and we use a traditional
distribution-theory (or generalised-functions) approach.
Given a function $v\in\sobh2(\W)$ and let $\geovec
n:\partial\W\to\reals^d$ be the outward pointing normal of $\W$ then the
Hessian of $v$, $\Hess v$ satisfies the following identity:
\begin{equation}
\label{eq:generalised-hessian}
\ltwop{\Hess v}{\phi}
=
-
\int_\W \nabla v \otimes \nabla \phi
+
\int_{\partial\W} \nabla v \otimes \geovec n\,\phi
\Foreach \phi\in\sobh1(\W),
\end{equation}
where $\geovec a\otimes\geovec b:=\geovec a\transposevec b$ for
$\vec a,\vec b$ column vectors in $\R d$. If $v\in\sobh{1}(\W)$
with $\restriction{\grad v}{\boundary\W}\in\sobh{-1/2}(\partial\W)$ the
right-hand side of (\ref{eq:generalised-hessian}) still makes sense
and defines $\Hess v$ as an element in the dual of $\sobh1(\W)$ via
\begin{equation}
\duality{\Hess v}{\phi}
:=
-
\int_\W \nabla v \otimes \nabla \phi
+
\int_{\partial\W} \nabla v \otimes \geovec n\,\phi
\Foreach \phi\in\sobh1(\W)
,
\end{equation}
where $\duality\cdot\cdot$ denotes the duality action on $\sobh1(\W)$
from its dual. We call $\Hess v$ the \emph{generalised Hessian} of
$v$, and assuming that the coefficient tensor $\A$ is in
$\cont0(\W)^{d\times d}$, for the product with a distribution to make
sense, we now seek $u\in\sobhz1(\W)$ such that $\restriction{\grad
u}{\boundary\W}\in\sobh{-1/2}(\W)$ and whose generalised Hessian
satisfies
\begin{equation}
\label{eq:linear-model}
\duality{\frob\A\Hess v}{\phi}
=
\ltwop f\phi
\Foreach
\phi\in\sobh1(\W).
\end{equation}
\begin{comment}
The mixed formulation of the model problem (\ref{Problem}) we consider is to seek
the pair $(u, \H[u]) \in \sobh{1}(\W)\cap\sobh{1}(\partial\W)
\times \leb{2}(\W)^{d\times d}$ such that
\begin{gather}
\label{eq:mixed-1}
\ltwop{\H[u]}{\phi}
+
\int_\W{\nabla u}\otimes{\nabla \phi}
-
\int_{\partial \W} {\nabla u}\otimes{\geovec{n} \ \phi}
=
\geomat 0
\\
\label{eq:mixed-2}
\ltwop{\frob{\A}{\H[u]}}{\psi}
=
\ltwop{f}{\psi}
\Foreach (\phi,\psi)\in\sobh1(\W)\times \leb{2}(\W).
\end{gather}
\end{comment}
\subsection{Finite element discretisation and finite element Hessian}
We discretise (\ref{eq:linear-model}) for simplicity
with a standard piecewise polynomial approximation for test and trial
spaces for both problem variable, $U$, and auxiliary (mixed-type) variable,
$\H[U]$. Let $\T{}$ be a conforming, shape regular
triangulation of $\W$, namely, $\T{}$ is a finite family of sets such
that
\begin{enumerate}
\item $K\in\T{}$ implies $K$ is an open simplex (segment for $d=1$,
triangle for $d=2$, tetrahedron for $d=3$),
\item for any $K,J\in\T{}$ we have that $\closure K\meet\closure J$ is
a full subsimplex (i.e., it is either $\emptyset$, a vertex, an
edge, a face, or the whole of $\closure K$ and $\closure J$) of both
$\closure K$ and $\closure J$ and
\item $\union{K\in\T{}}\closure K=\closure\W$.
\end{enumerate}
We use the convention where $\funk h\W\reals$ denotes the
\emph{meshsize function} of $\T{}$, i.e.,
\begin{equation}
h(\vec{x}):=\max_{\closure K\ni \vec x}h_K.
\end{equation}
We introduce the \emph{finite element spaces}
\begin{gather}
\label{eqn:def:finite-element-space}
\fes
:=\ensemble{\Phi \in \sobh1(\W)}{\Phi\vert_{K} \in \poly p\Forall
K\in\T{} \AND \Phi\in\cont{0}(\W)},
\\
\feszero
:=\fes \cap \hoz,
\end{gather}
where $\poly k$ denotes the linear space of polynomials in $d$
variables of degree no higher than a positive integer $k$. We
consider $p\geq 1$ to be fixed and denote by $\Nzero :=
\dim{\feszero}$ and $N := \dim{\fes}$.
The discretisation of problem then reads: Find $(U, \H[U]) \in
\feszero \times \fes^{d\times d}$
such that
\begin{equation}
\begin{gathered}
\label{eq:discrete-linear}
\ltwop{\H[U]}{\Phi}
=
-
\int_\W{\nabla U}\otimes{\nabla \Phi}
+
\int_{\partial \W} {\nabla U}\otimes{\geovec{n} \ \Phi}
\Foreach \Phi\in\fes,
\\
\ltwop{\frob{\A}{\H[U]}}{\Psi}
=
\ltwop{f}{\Psi}
\Foreach \Psi\in\feszero.
\end{gathered}
\end{equation}
For an algebraic formulation of (\ref{eq:discrete-linear}) we refer
the reader to \cite[\S 2]{LakkisPryer:2011}. Note that this
discretisation can be interpreted as a mixed method whereby the
first (matrix) equation defines the \emph{finite element Hessian} and the
second (scalar) equation approximates the original PDE (\ref{Problem}).
\subsection{Two discretisation stategies of \eqref{eq:modelproblem}}
\label{sec:two-nonlinear-finite-element-methods}
The finite element Hessian allows us two discretisation strategies.
The first strategy, detailed in \S\ref{sec:unconstrained-fnl-pde},
consists in applying Newton first to set-up
(\ref{eq:system-of-nonvar-pdes}) and then using the NVFEM
(\ref{eq:discrete-linear}) to solve each step. A second strategy
becomes possible, upon noting that given $U\in\fes$ the finite
element Hessian $\H[U]$ is a regular function,\footnote{A
generalised function $v$ is a \emph{regular function}, or just
\emph{regular}, if it can be represented by a Lebesgue measurable
function $f\in\lebloc1$ such that $\duality v\phi=\int_\W f\phi$
for all $\phi\in\cont\infty_0(\W)$. We follow the customary and
harmless abuse in identifying $v$ with $f$.} which the
generalised Hessian $\Hess U$ might fail to be. This allows to
apply nonlinear functions such as $F$ to $\H[U]$ and consider the
following \emph{fully nonlinear finite element method} (FNFEM)
\begin{equation}
\begin{gathered}
\label{eq:fully-nonlinear-FEM}
\ltwop{\H[U]}{\Phi}
=
-
\int_\W{\nabla U}\otimes{\nabla \Phi}
+
\int_{\partial \W} {\nabla U}\otimes{\geovec{n} \ \Phi}
\Foreach \Phi\in\fes,
\\
\ltwop{F(\H[U])}\Psi
=
\ltwop f\Psi
\Foreach \Psi\in\fez.
\end{gathered}
\end{equation}
Of course, in order to solve the second equation, a
finite-dimensional Newton method may be necessary (but this
strategy leaves the door open for other nonlinear solvers, e.g.,
fixed point iterations). A finite element code based on this idea
will be tested in \S\ref{sec:Pucci} to solve the Pucci equation.
In summary the finite element Hessian allows both paths in the
following diagram:
\begin{equation}
\small
\label{eqn:dia:Newton-discretizations}
\begin{tikzpicture}
\path (0,3) node(X){fully nonlinear PDE \eqref{eq:modelproblem}};
\path (6,3) node(N){nonvariational linear PDE's \eqref{eq:system-of-nonvar-pdes}};
\path (0,0) node(F){fully nonlinear FE discretization \eqref{eq:fully-nonlinear-FEM}};
\path (6,.5) node(D){\color adiscrete linear 1};
\path (6,0) node(L){\color bdiscrete linear 2};
\draw[-stealth,color=a] (X)--(N) node[pos=0.5,above]{Newton};
\draw[-stealth,color=a] (N)--(D) node[pos=0.5,above,sloped]{NVFEM};
\draw[-stealth,color=b] (X)--(F) node[pos=0.5,above,sloped]{FNFEM};
\draw[-stealth,color=b] (F)--(L) node[pos=0.5,above]{Newton};
\end{tikzpicture}
\end{equation}
Although the diagram in (\ref{eqn:dia:Newton-discretizations}) does
not generally commute, if the function $F$ is algebraically
accessible, then it is commutative. By ``algebraically accessible''
we mean a function that can be computed in a finite number of
algebraic operations or inverses thereof. In this paper, we use
only algebraically accessible nonlinearities, but, in principle
assuming derivatives are available, our methods could be extended to
algebraically inaccessible nonlinearities, such as Bellman's (or
Isaacs's) operators involving optimums over infinite families, e.g.,
\begin{equation}
F(\geomat M):=\inf_{\alpha\in\cA}\frob{\geomat L_{\alpha}}{\geomat M}
\quad
\qpbig{\OR
F(\geomat M):=\inf_{\alpha\in\cA}\sup_{\beta\in\cB}\frob{\geomat L_{\alpha,\beta}}{\geomat M}}
,
\end{equation}
where $\ensemble{\geomat L_{\alpha}}{\alpha\in\cA}$
(or $\ensemble{\geomat L_{\alpha,\beta}}{(\alpha,\beta)\in\cA\times\cB}$)
is a family of elliptic operators.
\section{The discretisation of unconstrained fully nonlinear PDEs}
\label{sec:unconstrained-fnl-pde}
In this section we detail the application of the method reviewed in
\S\ref{sec:ndfem} to the fully nonlinear model problem
\eqref{eq:modelproblem}. Many fully nonlinear elliptic PDEs must be
constrained in order to admit a unique solution. For example the \MAD
is elliptic and admits a unique solution in the cone of convex (or
concave) functions when $f>0$ (or $f<0$, respectively). Before we
turn our attention to the more complicated constrained PDE's in
\S\ref{sec:ma} and we illustrate the Newton--NVFEM method in the simplest light.
In this section we study fully nonlinear PDEs which have no such
constraint.
\begin{Hyp}[unconditionally elliptic linearisation]
\label{ass:well-posed}
We assume, in this section, that the Newton-step linearisation
(\ref{eq:system-of-nonvar-pdes}) is elliptic. For this assumption
to hold, it is sufficient to assume uniform ellipticity, i.e.,
(\ref{eq:ellipticity-for-derivative}) with $\cC=\symm$ and $\mu>0$
independent of $\geomat M$.
\end{Hyp}
\subsection{The Newton-NVFEM method}
\label{The:nlfem}
Suppose we are given a BVP of the form, finding
$u\in\sobh2(\W)\cap\hoz$ such that
\begin{equation}
\label{eq:nonlinearproblem}
\begin{split}
\nlop{u} = \nlfunk(\Hess u) - f &= 0 \qquad \text{ in } \W,
\end{split}
\end{equation}
which satisfies Assumption \ref{ass:well-posed}.
Upon applying Newton's method to approximate the solution of problem
(\ref{eq:nonlinearproblem}) we obtain a sequence of functions
$(u^n)_{n\in\naturals_0}$ solving the following linear equations in
nonvariational form,
\begin{equation}
\label{eq:linearisedequation}
\frob{\N(\Hess u^n)}{\Hess u^{n+1}} = g(\Hess u^n)
\end{equation}
where
\renewcommand{\thefootnote}{\fnsymbol{footnote}}
\begin{gather}
\label{eq:newton_prob_data:N}
\N(\geomat X) := F'(\geomat X),\\
\label{eq:newton_prob_data:g}
g(\geomat X)
:=
f - \nlfunk(\geomat X) + \frob{F'(\geomat X)}{\geomat X}.
\end{gather}
The nonlinear finite element method to approximate
(\ref{eq:linearisedequation}) is: given an initial guess $U^0 :=
\Pi_0 u^0$ for each $n\in\naturals_0$ find $(U^{n+1},
\H[U^{n+1}])\in \feszero\times\fes^{d\times d}$ such that
\begin{equation}
\begin{gathered}
\label{eq:discretenewtonsmethod}
\ltwop{\H[U^{n+1}]}{\Phi}
+
\int_\W{\nabla U^{n+1}}\otimes{\nabla \Phi}
-
\int_{\partial \W} {\nabla U^{n+1}}\otimes{\geovec{n}}\,\Phi
=
\geomat 0
\Foreach \Phi\in\fes
\\
\AND
\ltwop{\frob{\N(\H[U^{n}])}{\H[U^{n+1}]}}{\Psi}
=
\ltwop{g(\H[U^n])}{\Psi}
\Foreach \Psi\in\feszero.
\end{gathered}
\end{equation}
\begin{Obs}[initial conditions]
\end{Obs}
\begin{comment}
We now give an algorithm for the general method.
\subsection{$\Algoname{The NVFEM for a class of fully nonlinear problems}$}
\label{alg:fully-nonlinear}
\begin{algorithmic}
\Require $(\T{}_0, u^0, p, N, K_{\max}, \tol)$
\Ensure $(U)$ the
NVFE solution of (\ref{eq:discretenewtonsmethod})
\State $k = 0$
\While{$k \leq K_{\max}$}
\State $\fes_k = \Algoname{FE Space}(\T{}_k, p)$
\If{$k = 0$}
\State $U^0 = \laginterpol{\fes_0} u^0$
\State $\H[U^0] = \Algoname{Hessian Recovery}(U^0,\fes_0)$
\EndIf
\State $n = 0$
\While{$n \leq N$}
\State $[U^{n+1}, \H[U^{n+1}]] = \Algoname{NVFEM}(\fes_k, \geomat N(\H[U^n]), g(\H[U^n]))$
\If{$\Norm{U^{n+1}-U^n} \leq \tol$}
\State break
\EndIf
\State $n = n+1$
\EndWhile
\State $\T{}_{k+1} = \Algoname{Global Refine}(\T{}_k)$
\State $k = k+1$
\EndWhile
\end{algorithmic}
\end{comment}
\subsection{Numerical experiments: a simple example}
In this section we detail numerical experiments aimed at demonstrating
the application of \eqref{eq:discretenewtonsmethod} to a simple model
problem.
\begin{Example}[a simple fully nonlinear PDE]
\label{ex:fullynl-abs}
The first example we consider is a fully nonlinear PDE with a very
smooth nonlinearity. The problem is
\begin{equation}
\begin{split}
\nlop{u} := \sin{\Delta u} + 2\Delta u -f &= 0 \text{ in } \W,
\\
u &= 0 \text{ on } \pd{}\W.
\end{split}
\end{equation}
which is specifically constructed to be uniformly elliptic. Indeed
\begin{equation}
\nlfunk'(\Hess u)
=
\left(\cos{\Delta u} + 2\right)\eye.
\end{equation}
which is uniformly positive definite.
The Newton linearisation of the problem is then: Given $u^0$, for
$n\in\naturals_0$ find $u^{n+1}$ such that
\begin{equation}
\frob{\left(\cos{\Delta u^n} + 2\right)\eye}
{\Hess (u^{n+1} - u^n)}
=
f - \sin{\Delta u^n} - 2\Delta u^n.
\end{equation}
and our approximation scheme is nothing but
\ref{eq:discretenewtonsmethod} with
\begin{gather}
\N(\geomat X) = \left(\cos{\trace{\geomat X}} + 2\right)\geomat I
\\
g(\geomat X) = f - \sin{\trace{\geomat X}} - 2\trace{\geomat X}.
\end{gather}
Figure \ref{Fig:nonlin-regular-abs-lap} details a numerical
experiment on this problem when $d=2$ and when $\W =[-1,1]^2$ is a
square which is triangulated using a criss-cross mesh.
\end{Example}
\begin{Rem}[simplification of Example \ref{ex:fullynl-abs}]
\label{rem:variational-simplification}
Example \ref{ex:fullynl-abs} can be simplified considerably by noticing that
\begin{equation}
\int_\W \tr{\H[U]} \Phi = \int_\W \Transpose{\qp{\nabla U}}\nabla \Phi
\Foreach \Phi\in\feszero.
\end{equation}
This coincides with the definition of the \emph{discrete Laplacian}
and makes the NVFEM coincide with the standard conforming FEM. This
observation applies to all fully nonlinear equations, with
nonlinearity of the form \eqref{eq:modelproblem} with
\begin{equation}
F(\geomat M):=a(\tr{\geomat M}),
\end{equation}
for some given $a$. This class of problems, can be solved using a
variational finite element method and can be used for comparison
with our method. Note that in
\cite{JensenSmears:12:techreport:Finite} the authors use this
together with a localisation argument in order to prove convergence
of a finite element method for a specific class of
Hamilton--Jacobi--Bellman equation. Their method coincides with ours,
for an appropriate choice of quadrature.
\end{Rem}
\begin{figure}[h]
\caption[]{\label{Fig:nonlin-regular-abs-lap} Numerical experiments
for Example \ref{ex:fullynl-abs}. Choosing $f$ appropriately such
that $u(\geovec x) = \exp\qp{-10\norm{\geovec x}^2}$. We use an
initial guess $u^0 = 0$ and run the iterative procedure until
$\Norm{U^{n+1} - U^n} \leq 10^{-8}$, setting $U := U^M$
the final Newton iterate of the sequence. Here we are plotting
log--log error plots together with experimental convergence rates
for $\leb{2}(\W), \sobh1(\W)$ error functionals for the problem
variable, $U$, and an $\leb{2}(\W)$ error functional for the
auxiliary variable, $\H[U]$. Notice that there is a
``superconvergence'' of the auxiliary variable for both
approximations.}
\subfigure[][{Taking $\fes$ to be the space of
piecewise linear functions on $\W$ ($p = 1$). Notice that $\Norm{u -
U^M} = \Oh(h^2)$, $\norm{u - U^M}_1 = \Oh(h)$ and $\Norm{\Hess u
- \H[U^M]} = \Oh(h^{0.5})$. }]{
\label{Fig:sin-lap-p1}
\includegraphics[scale=\figscale,width=0.47\figwidth]{Figures/fullynonlinear-abs-sin-linears}
} \hfill \subfigure[][{Taking $\fes$ to be the space of piecewise
quadratic functions on $\W$ ($p = 2$). Notice that $\Norm{u -
U^M} = \Oh(h^3)$, $\norm{u - U^M}_1 = \Oh(h^2)$ and
$\Norm{\Hess u - \H[U^M]} = \Oh(h^{1.5})$}]{
\label{Fig:sin-lap-p2}
\includegraphics[scale=\figscale,width=0.47\figwidth]{Figures/fullynonlinear-abs-sin-quadratics}
}
\end{figure}
\begin{Example}[nonvariational example]
\label{eq:fullynl-unsolvable}
This is a simple example where the \emph{variational trick}
mentioned in Remark \ref{rem:variational-simplification} cannot be applied. We fix
$d=2$ and consider the problem
\begin{equation}
\begin{split}
\nlop{u}:=
\qp{\partial_{11} u}^3
+
\qp{\partial_{22} u}^3
+
\partial_{11} u
+
\partial_{22} u
-
f
&=
0 \text{ in } \W
\\
u &= 0 \text{ on } \partial\W.
\end{split}
\end{equation}
The approximation scheme is then \eqref{eq:discretenewtonsmethod} with
\begin{gather}
\N(\geomat X) :=
\begin{bmatrix}
3\geomat X_{11}^2 + 1 & 0\\
0 & 3\geomat X_{22}^2 +1
\end{bmatrix}
\\
g(\geomat X)
:=
f + 2\qp{\geomat X_{11}^3 + \geomat X_{22}^3}.
\end{gather}
Figure \ref{Fig:nonlin-krylov} details a numerical experiment on
this problem in the case $d=2$ and $\W = [-1,1]^2$ triangulated with
a criss-cross mesh. A similar example is also studied in
\cite[Ex 5.2]{DavydovSaeed:12:techreport} using B\"ohmers method.
\begin{figure}[h]
\caption[]{\label{Fig:nonlin-krylov} Numerical experiments
for Example \ref{eq:fullynl-unsolvable}. Choosing $f$ appropriately such
that $u(\geovec x) = \exp\qp{-10\norm{\geovec x}^2}$. We use an
initial guess $u^0 = 0$ and run the iterative procedure until
$\Norm{U^{n+1} - U^n} \leq 10^{-8}$, setting $U := U^M$
the final Newton iterate of the sequence. Here we are plotting
log--log error plots together with experimental convergence rates
for $\leb{2}(\W), \sobh1(\W)$ error functionals for the problem
variable, $U$, and an $\leb{2}(\W)$ error functional for the
auxiliary variable, $\H[U]$. Notice that there is a
``superconvergence'' of the auxiliary variable for both
approximations.}
\subfigure[][{Taking $\fes$ to be the space of
piecewise linear functions on $\W$ ($p = 1$). Notice that $\Norm{u -
U^M} = \Oh(h^2)$, $\norm{u - U^M}_1 = \Oh(h)$ and $\Norm{\Hess u
- \H[U^M]} \approx \Oh(h^{1.5})$. }]{
\label{Fig:sin-lap-p1}
\includegraphics[scale=\figscale,width=0.47\figwidth]{Figures/fullynonlinear-krylov-exp-linears}
} \hfill \subfigure[][{Taking $\fes$ to be the space of piecewise
quadratic functions on $\W$ ($p = 2$). Notice that $\Norm{u -
U^M} = \Oh(h^3)$, $\norm{u - U^M}_1 = \Oh(h^2)$ and
$\Norm{\Hess u - \H[U^M]} = \Oh(h^{1.3})$}]{
\label{Fig:sin-lap-p2}
\includegraphics[scale=\figscale,width=0.47\figwidth]{Figures/fullynonlinear-krylov-exp-quadratics}
}
\end{figure}
\end{Example}
\section{The \MAD problem}
\label{sec:ma}
In this section we propose a numerical method for the \MAD (MAD) problem
\begin{equation}
\label{eq:ma}
\begin{split}
\det{\Hess u}
&=
f \text{ in }\W
\\
u
&=
g \text{ on }\partial \W.
\end{split}
\end{equation}
Our numerical experiments exhibit robustness of our method when
computing (smooth) classical solutions of the MAD equation. Most
importantly we noted the following facts:
\begin{enumerate}[(i)\ ]
\item
the use of $\poly p$ elements with $p\geq2$ is essential
as $\poly1$ do not work,
\item
the convexity of the Newton iterates is
conserved throughout the computation, in a similar way to the
observations in \cite{loeper-rapetti:05}, where the authors prove this
convexity-conservation property.
\end{enumerate}
Our observations are purely
empirical from computations, which leaves an interesting open problem
of proving this property.
\begin{Rem}[the MAD problem fails to satisfy Assumption \ref{ass:well-posed}]
To clarify Assumption \ref{ass:well-posed} for the MAD problem
(\ref{eq:ma}), in view of the characteristic expansion of
determinant if $\geomat X,\geomat Y \in \symm$
\begin{equation}
\det\qp{\geomat X + \epsilon \geomat Y}
=
\det{\geomat X} + \epsilon \frob{\cof\qp{\geomat X}}{\geomat Y} + \Oh(\epsilon^2),
\end{equation}
where $\cof{\geovec X}$ is the
matrix of cofactors of $\geovec X$. Hence
\begin{equation}
\label{eq:der-of-det-is-cof}
\nlfunk'(\geomat X) = \cof{\geomat X}
.
\end{equation}
This implies that the linearisation of MAD is only well posed if we
restrict the class of functions we consider to those $u$ that
satisfy
\begin{equation}
\label{eqn:MAD-ellipticity-via-derivative}
\Transpose{\geovec \xi} \cof \Hess u\, \geovec \xi
\geq
\lambda \norm{\geovec\xi}^2 \Foreach \geovec\xi\in\reals^d
\end{equation}
for some ($u$-dependent) $\lambda>0$.
Note that (\ref{eqn:MAD-ellipticity-via-derivative}) is equivalent to the
following two conditions as well
\begin{equation}
\begin{gathered}
\Transpose{\geovec \xi} \Hess u\, \geovec \xi
\geq
\lambda \norm{\geovec\xi}^2 \Foreach \geovec\xi\in\reals^d
\\
$u$ \text{ is strictly convex. }
\end{gathered}
\end{equation}
\cite{LoeperRapetti:2005} have shown that for the \emph{continuous}
(infinite dimensional) Newton method described in
\ref{sec:newtons-method}, given an strictly convex initial guess
$u^0$, each iterate $u^n$ will be convex. It is crucial that this
property is preserved at the discrete level, as it guarantees the
solvability of each iteration in the \emph{discretised} Newton
method. For this it the right notion of convexity turns out to be
the \emph{finite element convexity} as developed in
\cite{Aguilera:2008}. In \cite{Pryer:2010}, an intricate method
based on semidefinite programming provided a way to constrain the
solution in the case of $\poly1$ elements. Here we observe that
the finite element convexity is automatically preserved, provided we
use $\poly2$ or higher conforming elements.
\end{Rem}
\subsection{Newton's method applied to \MA}
In view of (\ref{eq:der-of-det-is-cof}) it is clear that
\begin{equation}
\Dnlop{u}v = \frob{\cof{\Hess u}}{\Hess v}.
\end{equation}
Applying the methodology set out in \S\ref{sec:unconstrained-fnl-pde}
we set
\begin{gather}
\label{eq:maN}
\N(\Hess u^n) = \cof{\Hess u^n},
\\
\label{eq:mag}
g(\Hess u^n) = f - \det{\Hess u^n} + \frob{\cof{\Hess
u^n}}{\Hess u^n},
\end{gather}
\begin{Rem}[{relating cofactors to determinants}]
\label{rem:cof-to-det}
For a generic (twice differentiable) function $v$ it holds
that
\begin{equation}
d \det{\Hess v} = \frob{\cof{\Hess v}}{\Hess v}.
\end{equation}
Using this formulation we could construct a simple fixed point
method for the \MA equation.
\end{Rem}
In view of Remark \ref{rem:cof-to-det} $g$ can be further
simplified
\begin{equation}
\begin{split}
g(\Hess u^n) &= f - \det{\Hess u^n} + \frob{\cof{\Hess
u^n}}{\Hess u^n}
\\
&= f + (d-1) \det{\Hess u^n}.
\end{split}
\end{equation}
Newton's method reads: Given $u^0$ for each $n\in\naturals_0$ find
$u^{n+1}$ such that
\begin{equation}
\label{eq:linearised-ma}
\frob{\N(\Hess u^n)}{\Hess u^{n+1}}
=
g(\Hess u^n).
\end{equation}
\subsection{Numerical experiments}
In this section we study the numerical behaviour of the scheme
presented in Definition \ref{The:nlfem} applied to the MAD problem.
We present a set of benchmark problems constructed from the problem
data such that the solution to the \MA equation is known. We fix $\W$
to be the square $S = [-1,1]^2$ or $[0,1]^2$ (specified in the
problem) and test convergence rates of the discrete solution to the
exact solution.
Figures \ref{Fig:EOC-ma1}--\ref{Fig:EOC-ma2} details the various
experiments and shows numerical convergence results for each of the
problems studied as well as solution plots, it is worthy of note that
each of the solutions seems to be convex, however this is not
necessarily the case. They are all though \emph{finite element convex}
\cite{Aguilera:2008}. In each of these cases the Dirichlet boundary
values are not zero. The implementation of nontrivial boundary
conditions is described in \cite[\S 3.6]{LakkisPryer:2011} or in more
detail in \cite[\S 4.4]{Pryer:2010}.
\begin{Rem}[choosing the ``right'' initial guess]
As with any Newton method we require a starting guess, not just for
$U^0$ but also of $\H[U^0]$. Due to the mild nonlinearity with the
previous example an initial guess of $U^0\equiv 0$ and $\H[U^0]
\equiv \geomat 0$ was sufficient. The initial guess to the MAD
problem must be more carefully sought.
Since we restrict our solution to the space of convex functions, it
is prudent for the initial guess to also be convex. Moreover we must
rule out constant and linear functions over $\W$, since the Hessian
of these objects would be identically zero, destroying ellipticity
on the initial Newton step. Hence we specify that the initial guess
to (\ref{eq:linearised-ma}) must be strictly convex. Rather than
postprocessing the finite element Hessian from a initial project
(although this is an option) to initialise the algorithm we solve a
linear problem using the nonvariational finite element
method. Following a trick, described in \cite{DeanGlowinski:2003},
we chose $U^0$ to be the standard $\fes$-finite element
approximation of $u^0$ such that
\begin{gather}
\label{eq:laplace-initial-guess}
\Delta u^0 = 2\sqrt{f} \text{ in } \W
\\
u^0 = g \text{ on } \partial\W.
\end{gather}
\end{Rem}
\begin{Rem}[degree of the FE space]
\label{rem:degree-of-fe-space}
In the previous example the lowest order convergent scheme was found
by taking $\fes$ to be the space of piecewise linear functions
($p=1$). For the MAD problem we require a higher approximation
power, hence we take $\fes$ to be the space of piecewise
quadratic functions, \ie $p=2$.
Although the choice of $p=1$ gives a stable scheme, convergence is
not achieved. This can be characterised by \cite[Thm
3.6]{Aguilera:2008} that roughly says you require more
approximation power than what piecewise linear functions provide to
be able to approximate all convex functions. Compare with Figure \ref{Fig:EOC-ma4}.
\end{Rem}
\begin{figure}[h]
\caption{ \label{Fig:EOC-ma1} Numerical results for the MAD problem
on the square $S = [-1,1]^2$. We choose the problem data $f$ and
$g$ appropriately such that the solution is the radially symmetric
function $u(\geovec x) = \exp\qp{{\norm{\geovec x}^2}/{2}}$. We
plot the finite element solution together with a log--log error
plot for various error functionals as in Figure
\ref{Fig:nonlin-regular-abs-lap}. Note for $p=2$ the $\leb{2}(\W)$
error rate of convergence is suboptimal, this is in agreement with
the numerical examples produced in \cite{BrennerGudiNeilanSung:2011}}
\begin{center}
\subfigure[][{The FE approximation to the function $u(\geovec x) = \exp\qp{\frac{\norm{\geovec
x}^2}{2}}$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{Figures/ma-solution-exp}
} \hfill
\subfigure[][{Log--log error plot for $\poly{2}$ Lagrange FEs.}]{
\label{Fig:ma-radial}
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/EOC-MA-radial}
}
\end{center}
\end{figure}
\begin{figure}[h]
\caption{ \label{Fig:EOC-ma3} A FE--convexity test for the numerical example given in \ref{Fig:EOC-ma1}. We plot $\det{\H[U]}$ together with the principle minor of $\H[U]$.}
\begin{center}
\subfigure[][{The principal minor of $\H[U]$, an approximation to the Hessian of the function $u(\geovec x) = \exp\qp{\frac{\norm{\geovec x}^2}{2}}$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/principle-minor-p2-exp.jpg}
} \hfill
\subfigure[][{The determinant of $\H[U]$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/determinant-p2-exp.jpg}
}
\end{center}
\end{figure}
\begin{figure}[h]
\caption{ \label{Fig:EOC-ma4} Numerical results for the MAD problem
on the square $S = [-1,1]^2$. We choose the problem data $f$ and
$g$ appropriately such that the solution is the radially symmetric
function $u(\geovec x) = \exp\qp{{\norm{\geovec x}^2}/{2}}$.}
\begin{center}
\subfigure[][{The $\poly{1}$ FE approximation to the function $u(\geovec x) = \exp\qp{\frac{\norm{\geovec
x}^2}{2}}$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/p1-test-solution.jpg}
} \hfill
\subfigure[][{The error $u-U$ plotted as a function over $\W$. Note the
FE approximation does not converge in this case, see Remark \ref{rem:degree-of-fe-space}}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/p1-test-error.jpg}
}
\end{center}
\end{figure}
\begin{figure}[h]
\caption{\label{Fig:EOC-ma2} Numerical results for the MAD problem
on the square $S=[-1,1]^2$. Choosing $f$ and $g$ appropriately
such that the solution is $u(\geovec x) = -\qp{2 - x_1^2 -
x_2^2}^{1/2}$. Note the function has singular derivatives on the
corners of $S$. We plot the finite element solution together with
a log--log error plot for various error functionals as in Figure
\ref{Fig:nonlin-regular-abs-lap}.}
\begin{center}
\subfigure[][{The FE approximation to the function $u(\geovec x)
= -\qp{2 - x_1^2 - x_2^2}^{1/2}$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{Figures/ma-solution-ballie}
} \hfill
\subfigure[][{Log--log error plot for $\poly{2}$ Lagrange FEs.}]{
\label{Fig:ma-ballie}
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/EOC-MA-ballie}
}
\end{center}
\end{figure}
\subsection{Nonclassical solutions}
\label{sec:nonclassical-monge}
The numerical examples given in Figures \ref{Fig:EOC-ma1}--\ref{Fig:EOC-ma2} both describe the numerical approximation of classical solutions to the MAD problem. In the case of Figure \ref{Fig:EOC-ma1} $u\in\cont{\infty}(\closure\W)$ whereas in Figure \ref{Fig:EOC-ma2} $u\in\cont{\infty}(\W)\cap\cont{0}(\closure{\W})$. We now take a moment to study less regular solutions, \ie viscocity solutions which are not classical. In this test we the solution
\begin{equation}
u(\geovec x) = \norm{\geovec x}^{2\alpha}
\end{equation}
for $\alpha\in (1/2, 3/4)$. The solution $u(\geovec x) \notin \sobh{2}(\W)$. In Figures \ref{Fig:EOC-alpha-0-55}--\ref{Fig:EOC-alpha-0-7} we vary the value of $\alpha$ and study the convergence properties of the method. We note that the method fails to find a solution for $\alpha \leq 1/2$. Finally in Figure \ref{Fig:monge-adaptive} we conduct an adaptive experiment based on a gradient recovery aposteriori estimator. The recovery estimator we make use of is the Zienkiewicz--Zhu patch recovery technique see
\cite{publication1}, \cite[\S 2.4]{Pryer:2010} or \cite[\S
4]{Ainsworth:2000} for further details.
\begin{figure}[h]
\caption{\label{Fig:EOC-alpha-0-55} Numerical results for the MAD problem
on the square $S=[-1,1]^2$. Choosing $f$ and $g$ appropriately
such that the solution is $u(\geovec x) = \norm{\geovec x}^{2\alpha}$, with $\alpha=0.55$. Note the function is singular at the origin. We plot the finite element solution together with a log--log error plot for various error functionals as in Figure
\ref{Fig:nonlin-regular-abs-lap}.}
\begin{center}
\subfigure[][{The FE approximation to the function $u(\geovec x) =
\norm{\geovec x}^{2\alpha}$, with $\alpha=0.55$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-55.jpg}
} \hfill
\subfigure[][{Log--log error plot for $\poly{2}$ Lagrange FEs.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-55-convergence}
}
\end{center}
\end{figure}
\begin{figure}[h]
\caption{\label{Fig:EOC-alpha-0-6} Numerical results for the MAD problem
on the square $S=[-1,1]^2$. Choosing $f$ and $g$ appropriately
such that the solution is $u(\geovec x) = \norm{\geovec x}^{2\alpha}$, with $\alpha=0.6$. Note the function is singular at the origin. We plot the finite element solution together with a log--log error plot for various error functionals as in Figure
\ref{Fig:nonlin-regular-abs-lap}.}
\begin{center}
\subfigure[][{The FE approximation to the function $u(\geovec x) =
\norm{\geovec x}^{2\alpha}$, with $\alpha=0.6$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-6.jpg}
} \hfill
\subfigure[][{Log--log error plot for $\poly{2}$ Lagrange FEs.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-6-convergence}
}
\end{center}
\end{figure}
\begin{figure}[h]
\caption{\label{Fig:EOC-alpha-0-7} Numerical results for the MAD problem
on the square $S=[-1,1]^2$. Choosing $f$ and $g$ appropriately
such that the solution is $u(\geovec x) = \norm{\geovec x}^{2\alpha}$, with $\alpha=0.7$. Note the function is singular at the origin. We plot the finite element solution together with a log--log error plot for various error functionals as in Figure
\ref{Fig:nonlin-regular-abs-lap}.}
\begin{center}
\subfigure[][{The FE approximation to the function $u(\geovec x) =
\norm{\geovec x}^{2\alpha}$, with $\alpha=0.7$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-7.jpg}
} \hfill
\subfigure[][{Log--log error plot for $\poly{2}$ Lagrange FEs.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-7-convergence}
}
\end{center}
\end{figure}
\begin{figure}[h]
\caption[]{Numerical results for a solution to the \MAD equation with
$f$ and $g$ appropriately
such that the solution is $u(\geovec x) = \norm{\geovec x}^{2\alpha}$, with $\alpha=0.55$.
We choose $p=2$, and use an adaptive
scheme based on Z--Z gradient recovery. The mesh is refined
correctly about the origin. Note that when $\dim{\fes} = 20,420$ the adaptive solution achieves $\Norm{u - U^M} \approx 0.0078$, the uniform solution given in Figure \ref{Fig:EOC-alpha-0-55} satisfies $\Norm{u-U^M} \approx 0.18$ using the same number of degrees of freedom. Using the adaptive strategy both $\Norm{u-U^M}$ and $\norm{u - U^M}_1$ converge like $\Oh(N^{-1})$. {\label{Fig:monge-adaptive} }}
\begin{center}
\subfigure[][{
Adaptive mesh
}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-55-adaptive-mesh.jpg}
}
\hfill
\subfigure[][{Log--log error plot for $\poly{2}$ Lagrange FEs.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/monge-singular-alpha-0-55-adaptive-convergence.pdf}
}
\end{center}
\end{figure}
\clearpage
\section{Pucci's equation}
\label{sec:Pucci}
In this section we look to discretise the nonlinear problem, in this
case Pucci's equation as a system of nonlinear equations. Pucci's
equation arises as a linear combination of Pucci's extremal
operators. It can nevertheless be written in an algebraically
accessible form, without the need to compute the eigenvalues.
\begin{Defn}[Pucci's extremal operators \cite{CaffarelliCabre:1995}]
Let $\geomat N \in \symm$ and $\eig{\geomat N}$ be it's spectrum,
then the extremal operators are
\begin{gather}
\cM(\geomat N) := \sum_{\lambda_i \in \sigma(\geomat N)} \alpha_i \lambda_i = 0,
\end{gather}
with $\alpha_i\in\reals$. The maximal (minimal) operator, commonly
denoted $\cM^+$ ($\cM^-$), has coefficients that satisfy
\begin{equation}
0 < \alpha_1 \leq \dots \leq \alpha_n
\qquad
\qp{ \alpha_1 \geq \dots \geq \alpha_n > 0}
\end{equation}
respectively.
\end{Defn}
\subsection{The planar case and uniform ellipticity}
In the case $d=2$ the normalised Pucci's equation reduces to finding $u$ such that
\begin{equation}
\label{eq:Pucci}
\alpha \lambda_2 + \lambda_1 = 0
\end{equation}
where $\geomat N := \Hess u$. Note that if $\alpha = 1$
(\ref{eq:Pucci}) reduces to the Poisson--Dirichlet problem. This can
be easily seen when reformulating the problem as a second order PDE
\cite{DeanGlowinski:2005}. Making use of the characteristic
polynomial, we see
\begin{equation}
\begin{split}
\lambda_i = \frac{\Delta u \pm \qp{\qp{\Delta u}^2 - 4\det{\Hess u}}^{1/2}}{2} \qquad i = 1,2.
\end{split}
\end{equation}
Thus Pucci's equation can be written as
\begin{equation}
\label{eq:pucci-pde}
0 = \qp{\alpha + 1}\Delta u + \qp{\alpha - 1}\qp{\qp{\Delta u}^2 - 4\det{\Hess u}}^{1/2},
\end{equation}
which is a nonlinear combination of \MA and Poisson problems. However
owing to the Laplacian terms, and unlike the \MAD problem, Pucci's
equation is (unconditionally) uniformly elliptic for
\begin{equation}
\qp{\tr{\geomat X}}^2 - 4\det{\geomat X}
\geq
0
\Foreach
\geomat X \in \reals^{2\times 2}.
\end{equation}
The discrete problem we use is a direct approximation of
(\ref{eq:pucci-pde}), we seek $\qp{U, \H[U]}$ such that
\begin{gather}
\int_\W \bigg(\qp{\alpha + 1}\tr{\H[U]} + \qp{\alpha - 1}\qp{\qp{\tr{\H[U]}}^2 - 4\det{\H[U]}}^{1/2}\bigg)\Phi = 0
\\
\ltwop{\H[U]}{\Psi}
=
-
\int_\W{\nabla U}\otimes{\nabla \Psi}
+
\int_{\partial \W} {\nabla U}\otimes{\geovec{n} \ \Psi}
\Foreach (\Phi,\Psi)\in\feszero \times \fes.
\end{gather}
The result is a nonlinear system of equations which was solved using a
algebraic Newton method.
\subsection{Numerical experiments}
We conduct numerical experiments to be compared with those of
\cite{DeanGlowinski:2005}. The first problem we consider is a
classical solution of Pucci's equation (\ref{eq:Pucci}). Let $\geovec
x = \Transpose{\qp{x, y}}$, then the function
\begin{equation}
\label{eq:Pucci-solution}
u(\geovec x) = -\qp{\qp{\qp{x + 1}^2 + \qp{y+1}^2}^{\qp{1-\alpha}/2}}
\end{equation}
solves Pucci's equation almost everywhere away from $(x, y) = (-1,-1)$
with $g := u|_{\partial\W}$. Let $\T{}$ be an irregular triangulation
of $\W = [-0.95, 1]^2$. In Figure \ref{Fig:pucci-converence} we detail
a numerical experiement considering the case $\alpha \in [2,5]$.
\begin{figure}
\caption[]{Numerical results for a classical solution to Pucci's
equation (\ref{eq:Pucci-solution}). As with the case of the MAD
problem we choose $p=2$. We use a Newton method to solve the
algebraic system until the residual of the problem (see
\cite[c.f.]{Kelley:1995}) is less than $10^{-10}$ (which is
overkill to minimise Newton error effects). We plot log--log error
plots with experimental orders of convergence, for various norms
and values of $\alpha$.{\label{Fig:pucci-converence} }}
\begin{center}
\subfigure[][{
$\alpha = 2$
}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-classical-alpha-2.pdf}
}
\hfill
\subfigure[][{$\alpha =3$}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-classical-alpha-3.pdf}
}
\hfill
\subfigure[][{
$\alpha = 4$
}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-classical-alpha-4.pdf}
}
\hfill
\subfigure[][{$\alpha = 5$}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-classical-alpha-5.pdf}
}
\end{center}
\end{figure}
We also conduct a numerical experiment to be compared with
\cite{Oberman:2008}. In this problem we consider a solution of Pucci's
equation with a piecewise defined boundary. Let $\W = [-1,1]^2$ and
the boundary data be given as
\begin{equation}
\label{eq:pucci-pw-boundary}
g(\geovec x):=
\begin{cases}
1 \qquad \text{ when } \abs{x} \geq \frac{1}{2} \AND \abs{y} \geq \frac{1}{2}
\\
0 \qquad \text{ otherwise.}
\end{cases}
\end{equation}
Figure \ref{Fig:pucci-piecewise} details the numerical experiment on
this problem with various values of $\alpha$.
\begin{figure}[h]
\caption[]{Numerical results for a solution to Pucci's equation with
a piecewise defined boundary condition
(\ref{eq:pucci-pw-boundary}). We choose $p=2$, use a Newton method
to solve the algebraic system until the residual of the problem is
less than $10^{-10}$. We plot the solution for various values of
$\alpha$ as well as a cross section through the coordinate
axis. Notice that the solution becomes extremely badly behaved as
$\alpha$ increases. {\label{Fig:pucci-piecewise} }}
\begin{center}
\subfigure[][{
$\alpha = 2$
}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-pw-alpha-2.jpg}
}
\hfill
\subfigure[][{$\alpha =3$}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-pw-alpha-3.jpg}
}
\hfill
\subfigure[][{
$\alpha = 4$
}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-pw-alpha-4.jpg}
}
\hfill
\subfigure[][{$\alpha = 5$}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-pw-alpha-5.jpg}
}
\hfill
\subfigure[][{Cross section about $x = 0$ for each value of $\alpha$.}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-pw-cross-section.jpg}
}
\end{center}
\end{figure}
Since the solution to the Pucci's equation with piecewise boundary
(\ref{eq:pucci-pw-boundary}) is clearly singular near the
discontinuities we have also conducted an adaptive experiment based on
a gradient recovery aposteriori estimator (as in \S\ref{sec:nonclassical-monge}). As can be seen from Figure
\ref{Fig:pucci-adaptive} we regain qualitively similar results using
far fewer degrees of freedom.
\begin{figure}[h]
\caption[]{Numerical results for a solution to Pucci's equation with
a piecewise defined boundary condition
(\ref{eq:pucci-pw-boundary}). We choose $p=2$, and use an adaptive
scheme based on Z--Z gradient recovery. The mesh is refined
correctly about the jumps on the
boundary. {\label{Fig:pucci-adaptive} }}
\begin{center}
\subfigure[][{
Finite element solution for $\alpha = 3$
}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-pw-adaptive-alpha-3.jpg}
}
\hfill
\subfigure[][{Adaptive mesh}]{
\includegraphics[scale=\figscale,width=0.47\figwidth]{fullynonlinear/Figures/pucci-pw-adaptive-mesh-alpha-3.jpg}
}
\end{center}
\end{figure}
\section{Conclusions and outlook}
In this work we have proposed a novel numerical scheme for fully
nonlinear and generic quasilinear PDEs. The scheme was based on a
previous work for nonvariational PDEs (those given in nondivergence
form) \cite{LakkisPryer:2011}.
We have illustrated the application of the method for a simple, non
physically motivated example, moving on to more interesting problems,
that of the \MA equation and Pucci's equation.
For Pucci's equation we numerically showed convergence and conducted
experiments which may be compared with previous numerical studies.
We demonstrated for classical solutions to the \MA equation the method
is robust again showing numerical convergence. For less regular
viscocity solutions we have found that the method must be augmented
with a penalty term in a similar light to
\cite{BrennerGudiNeilanSung:2011}.
We postulate that the method is better suited to a discontinuous
Galerkin framework which is the subject of ongoing research.
\bibliographystyle{abbrvnat}
|
\section{Introduction}
Recent studies revealed that scattering amplitudes in planar $\mathcal{N}=4$ supersymmetric
Yang-Mills theory (SYM) possess a new hidden symmetry \cite{DrunHenKorSok10,BerMal08}. The
latter appears in addition to the conventional superconformal symmetry of the Lagrangian of the
theory and leads to powerful constraints on the form of the all-loop scattering amplitudes.
A distinguishing feature of $\mathcal{N}=4$ SYM is that all on-shell states (gluons with helicity $\pm 1$,
gluinos with helicity $\pm 1/2$ and scalars) can be encoded in a single superstate $\Phi(p_i,\eta_i)$
\cite{BriMan82,Nair}. States with different helicities appear as coefficients in the expansion of the superstate in powers of the odd
variables $\eta_i^A$ (with $A=1,2,3,4$). As a consequence, all $n-$particle scattering amplitudes can
be combined into a single superamplitude $\mathcal{A}_n$. Supersymmetry restricts the form of the
superamplitude to be
\begin{align}\label{dec}
\mathcal{A}_n = \mathcal{A}_n^{\rm MHV} + \mathcal{A}_n^{\rm NMHV}
+ \ldots + \mathcal{A}_n^{\rm N^{n-4}MHV}\,,
\end{align}
where the $k$th term in the sum, $\mathcal{A}_n^{\rm N^kMHV}$, describes the scattering amplitudes with
total helicity of the particles $(-n+4+2k)$. It has the following general form \cite{Nair}
\begin{align}\label{gen}
\mathcal{A}_n^{\rm N^kMHV}
= i (2\pi)^4
\frac{\delta^{(4)}(p^{\dot\alpha\alpha})\delta^{(8)}(q^A_\alpha)}{\vev{12}\vev{23}\ldots \vev{n1}}
\,
\widehat{\mathcal{A}}_{n;k}(\lambda,\tilde\lambda,\eta;a)\,,
\end{align}
where $p^{\dot\alpha\alpha}=\sum_i p_i^{\dot\alpha\alpha}$ and $q^{\alpha\,A}=\sum_i
\lambda_i^\alpha\eta_i^A$
is the total momentum and chiral supercharge of the $n$ particles. Here we used the
spinor-helicity formalism to parameterize the light-like momenta in terms of two-component commuting spinors
$p_j^{\dot\alpha\alpha} = \widetilde\lambda^{\dot\alpha}_j \lambda_j^\alpha$ and the angular brackets $\vev{jk}$
are defined in Appendix A. The dependence of the scattering amplitude on the 't Hooft coupling constant
$a=g^2 N_c/(4\pi^2)$ is carried by the nontrivial function $\widehat A_{n;k}(\lambda,\tilde\lambda,\eta;a)$
which is given by a homogenous polynomial in the $\eta$'s of
degree $(4k)$. At tree level, the function $\widehat A_{n;k}(\lambda,\tilde\lambda,\eta)$ enjoys the full
$PSU(2,2|4)$ superconformal symmetry of the $\mathcal{N}=4$ SYM, while at loop level
the dilatations, the special conformal boosts and their supersymmetric extension are broken by infrared divergences \cite{Witten:2003nn}.
As we already mentioned, the superamplitude in planar $\mathcal{N}=4$ SYM has another, hidden
symmetry. This symmetry, called ``dual superconformal symmetry" in Ref.~\cite{DrunHenKorSok10}, acts naturally on the so-called
region supermomenta $(x_i,\theta_i)$ defined as
\begin{align}\label{super-mom}
p_i^{\dot\alpha\alpha}=\widetilde\lambda^{\dot\alpha}_i \lambda_i^\alpha
=
(x_i-x_{i+1})^{\dot\alpha\alpha}
\,,\qquad
\lambda_i^\alpha \,\eta_i^A = (\theta_i - \theta_{i+1})^{\alpha\, A}\,.
\end{align}
In spite of the fact that $(x_i,\theta_i)$ have the meaning of momenta, the dual symmetry acts on
them as if they were coordinates in configuration (rather than momentum) space
\cite{B93,DruHenSmiSok06,DrunHenKorSok10}. At tree level, the superamplitude
$\mathcal{A}_n^{(0)}$ enjoys exact dual superconformal symmetry, while at loop level some of the dual
symmetries become anomalous (see below).
Dual superconformal symmetry emerges as a hidden property of the scattering amplitudes in planar
$\mathcal{N}=4$ SYM. For the simplest, maximally helicity violating superamplitude
$\mathcal{A}_n^{\rm MHV}$, the dual conformal symmetry becomes manifest through the conjectured
duality between the functions $\widehat{\mathcal{A}}_{n;0}$ defining the perturbative loop corrections to the MHV
superamplitude, Eq.~\re{gen}, and light-like Wilson loops $W_n$
\footnote{The duality between planar amplitudes and light-like Wilson loops also holds in gauge theories
with less or no supersymmetry, including QCD. However, in distinction with $\mathcal{N}=4$ SYM, there the
relation \re{MHV-dual} holds in the high-energy (Regge) limit only~\cite{Korchemskaya:1996je,KorDruSok08}.}
\begin{align}\label{MHV-dual}
\ln \widehat{\mathcal{A}}_{n;0} = \ln W_n +O(\varepsilon) \,,\qquad
W_n= \frac{1}{N_c} \VEV{ \mathop{\rm tr}\nolimits P\exp\lr{i g\int_{C_n} dx\cdot A(x)}}
\, ,
\end{align}
where $O(\varepsilon)$ stands for terms vanishing in dimensional regularization and the Wilson loop is
evaluated along the polygon light-like contour $C_n=[x_1, x_2] \cup [x_2, x_3] \dots [x_n, x_1]$ formed
by the particle momenta $p_i=x_i-x_{i+1}$.
The power of the duality relation \re{MHV-dual} exhibits itself
through the fact that the dual conformal symmetry of the MHV superamplitude is mapped into the conventional
conformal symmetry of the Wilson loop in $\mathcal{N}=4$ SYM. First hinted at
in Ref.~\cite{AldMal08} by the
strong coupling analysis for $n=4,5$ via the AdS/CFT correspondence,
the duality between the two objects was understood as the invariance of the string sigma model on an
AdS$_5 \times$S$^5$ background under $T$-duality transformations for the bosonic \cite{Kallosh:1998ji} and fermionic variables
\cite{BerMal08}. At weak coupling the proposal (\ref{MHV-dual}) was put forward in Ref.\ \cite{KorDruSok08} and
confirmed at one loop for an arbitrary number of cusps $n$ in Ref.~\cite{BraHesTra07} and at two loops
for $n=4,5-$point Wilson loops \cite{KorDruHenSok08,DruHenKorSok07} by confronting these predictions with the
available results for gluon scattering amplitudes \cite{ABDK03,BDS05}. At the moment, the most thorough test of the
duality \re{MHV-dual} comes from the comparison of the two-loop results for the $n=6$ (hexagon) Wilson loop
\cite{DruHenKorSok08,AnaBraHesKhoSpeTra09,DelDuhSmi10,GonSprVerVol10} with the six-gluon MHV scattering
amplitude computed to the same order in the coupling \cite{BerDixKosRoiSprVerVol08,Cachazo:2008hp}.
The MHV superamplitude is the simplest among all superamplitudes in \re{dec}. In particular, the
corresponding function $\widehat{\mathcal{A}}_{n;0}$ defined in Eq.\ \re{gen} only depends on the bosonic
variables and is independent of the odd $\eta-$variables. At tree level, it is given by
$\widehat{\mathcal{A}}_{n;0}=1+O(a)$ and, according to the duality relation \re{MHV-dual}, its loop corrections
coincide (upon appropriate identification of the regularization parameters) with those of the bosonic light-like
Wilson loop $W_n$. A natural question arises whether
the duality relation \re{MHV-dual} can be extended to the full superamplitude $\mathcal{A}_n$. If
such a duality relation exists, the bosonic Wilson loop in the right-hand side of \re{MHV-dual} should
appear as the first term in the expansion of the dual object $\mathcal{W}_n$ in powers of the
$\eta-$variables analogous to \re{dec}. Viewed as a function of the dual coordinates $(x_i,\theta_i)$, it
should possess the same symmetry as the superamplitude. Namely, the duality between $\mathcal{A}_n$
and $\mathcal{W}_n$ would imply that the dual superconformal symmetry of the former follows from
the conventional superconformal symmetry of the latter in $\mathcal{N}=4$ SYM. Moreover, since
some of the symmetries of the superamplitudes are broken at the quantum level, $\mathcal{W}_n$ should
have the same anomalies.
Regarding the $\mathcal{N}=4$ supersymmetry, the dual object
$\mathcal{W}_n$ should have the following unusual property -- it depends only on the chiral $\theta-$variables
but not on $\bar\theta$. The reason for this can be traced back to the chiral formulation of scattering
amplitudes in the on-shell superspace formalism. In addition, a serious complication is that $\mathcal{N}=4$ SYM
only exists on shell, i.e. the algebra of the supersymmetry transformations closes modulo field equations. This makes
the construction of $\mathcal{W}_n$ somewhat tricky and, as we will see later in the paper, leads to various subtleties.
Recently, two different proposals for $\mathcal{W}_n$ were put forward. In Ref.\ \cite{MasSki10} Mason and Skinner put forward a
formulation of $\mathcal{W}_n$ in twistor space.
Yet another form of $\mathcal{W}_n$, this time in Minkowski space, was suggested by Caron-Huot in Ref.\ \cite{Car10}. It was claimed in
\cite{MasSki10} that all amplitudes in $\mathcal{N}=4$ SYM are described by a supersymmetric Wilson loop in twistor space and the
same statement was apparently implied for its Minkowski transform. The Minkowski version takes the following form \cite{OogRahRobTan00}
\begin{align}\label{super-W}
\mathcal{W}_n = \frac{1}{N_c}\VEV{ \mathop{\rm tr}\nolimits P\exp\lr{ig \int_{\mathcal{C}_n} dx^{\mu} \mathcal{A}_{\mu} (x,\theta)
+
ig \int_{\mathcal{C}_n} d \theta^{\alpha A} \mathcal{F}_{\alpha A}(x,\theta)}}\,,
\end{align}
where the integration goes over a contour ${\mathcal{C}_n}$ in superspace formed by $n$ straight segments
connecting the points $(x_i,\theta_i)$ and defined by the supermomenta of the scattered particles \re{super-mom}.
Here $\mathcal{A}_{\mu}$ and $\mathcal{F}_{\alpha A}$ are the bosonic and fermionic superspace gauge connections,
respectively,
given by a $\theta-$expansion whose coefficients are the various component fields (gluon, gauginos and scalars) of $\mathcal{N}=4$ SYM.
Suppressng the dependence on the $\theta-$variables, the supersymmetrized Wilson loop \re{super-W}
reduces to the bosonic Wilson loop $W_n$ in \re{MHV-dual}. Another proposal for $\mathcal{W}_n$
was presented in \cite{Car10} where certain perturbative data and BCFW-like recursion relations \cite{ArkaniHamed:2010kv,Boels:2010nw}
were used
to advocate a particular supersymmetric generalization of the bosonic light-like Wilson loop. The resulting
supersymmetric Wilson loop was argued to be equivalent to \re{super-W}. Thus, both proposals suggest
the duality ${\mathcal{A}}_n \sim \mathcal{W}_n$.
By construction, the supersymmetric Wilson loop $\mathcal{W}_n$ automatically reproduces the MHV
amplitudes through the
duality relation \re{MHV-dual}. In addition, it yields a definite prediction for non-MHV amplitudes to all loops.
In this paper, we verify the duality between the superamplitudes $\mathcal{A}_n$ and the supersymmetric
Wilson loop $\mathcal{W}_n$ by performing one-loop calculations of \re{super-W} in $\mathcal{N}=4$ SYM.
We show by
an explicit one-loop computation that, contrary to the above expectations, the perturbative corrections to $\mathcal{W}_n$ do
not match those of the non-MHV amplitudes, thus invalidating the duality relation $\mathcal{A}_n\sim\mathcal{W}_n$
beyond the MHV amplitudes. We show that the reason for this has to do with the fact that $\mathcal{W}_n$ is
invariant only under on-shell chiral $\mathcal{N}=4$ supersymmetric transformations, that is up to contributions
proportional to the equations of motion.
At loop level, due to the specific light-cone singularities of the Wilson loops \cite{KK92}, the equations
of motion generate a nontrivial finite effect. As a result, the supersymmetry of $\mathcal{W}_n$ gets
broken by quantum corrections. We compute the corresponding anomaly to one-loop order and demonstrate
that the result obtained for $\mathcal{W}_n$ is in a perfect agreement with the anomalous supersymmetric Ward
identities.
An alternative proposal for the dual description of scattering amplitudes in planar
$\mathcal{N}=4$ SYM via the light-cone limit of correlation functions
of certain half-BPS operators was put forward in Refs.\ \cite{EdeKorSok10,EHKS}. It proves to be immune to the
problems that one encounters in the construction of the super Wilson loops, as we discuss
in the Conclusions.
The paper is organized as follows. After a discussion of the basic symmetries that any
proposal for the dual description of superamplitudes should fulfill, in Sect.~2 we introduce the supersymmetric
generalizations of the bosonic Wilson loop proposed in Refs.~\cite{MasSki10,Car10} and explicitly
demonstrate that the two differ by the
presence of equations of motion when considered off shell. Then, in Sect.\ \ref{SUSYWI}, we derive the
Ward identities resulting from the the transformation properties of the
Wilson loops under Poincar\'e supersymmetry and calculate the corresponding anomalies. In Sect.\
\ref{OneLoopWL} we solve the anomalous Ward identities and reconstruct the explicit form of the one-loop
correction to the rectangular Wilson loop $\mathcal{W}_4$. Sect.\ \ref{Conclusions} contains
concluding remarks. Details on our conventions and normalizations are given in the Appendix A. Finally, the
Ward identities corresponding to the conformal transformations of supersymmetric Wilson loop are discussed
in Appendix B, where we demonstrate their complete consistency with our findings in the main text.
\section{Supersymmetric Wilson loop}
As already reviewed in the Introduction, the symmetries of the scattering amplitudes
extend far beyond the conventional superconformal symmetry of $\mathcal{N}=4$ SYM.
If a generalization of the Wilson loop dual to superamplitudes exists,
both objects will possess the same symmetries. This implies that the dual superconformal
symmetry of the amplitudes should match the conventional superconformal $
\mathcal{N}=4$ symmetry of the supersymmetric Wilson loop.
\subsection{Matching the symmetries of the $\mathcal{N}=4$ amplitudes}
\label{Matching}
Let us assume for a moment that the $\mathcal{N}=4$ superamplitudes are indeed dual to $\mathcal{W}_n$
and let us summarize the constraints imposed by the dual superconformal symmetry of the amplitude
$\mathcal{A}_n$ on the properties of $\mathcal{W}_n$.
By construction, $\mathcal{W}_n$ depends on the Grassmann variables $\theta_i^A$ carrying $SU(4)$
charges. Then by virtue of $SU(4)$ invariance, it admits an expansion in powers $(\theta)^{4k}$:
\begin{equation}\label{W-exp}
\mathcal{W}_n = \mathcal{W}_{n;0}(x_i;a)+ \mathcal{W}_{n;1} (x_i,\theta_i^A;a)
+ \ldots +\mathcal{W}_{n;n}(x_i,\theta_i^A;a)
\, .
\end{equation}
Here $\mathcal{W}_{n;k}$ is a homogeneous $SU(4)$ invariant polynomial of degree $4k$ in the Grassmann
variables $\chi_i^A = \vev{ i \theta_i^A}$ (see Eq.~\re{EtaToChi} below) and the expansion runs up to $k=n$. The lowest term
of the expansion does not depend on the odd variables and, therefore, it coincides with the bosonic Wilson
loop $\mathcal{W}_{n;0}(x_i;a) = W_n$, (see Eq.~\re{MHV-dual}).
The expansion \re{W-exp} is similar to that of the scattering amplitude \re{dec}. In fact, the conjectured duality
between superamplitudes and supersymmetric Wilson loops establishes the correspondence between the two
expansions~\cite{MasSki10,Car10}
\begin{align}\label{dual-par}
\mathcal{W}_{n;k} (x,\theta;a) =a^k \widehat{\mathcal{A}}_{n;k}(\lambda,\tilde\lambda,\eta;a) \,,
\end{align}
where the functions $ \widehat{\mathcal{A}}_{n;k}$ define the perturbative corrections to the N${}^k$MHV
superamplitude,
Eq.~\re{gen}. For $k=0$ the relation \re{dual-par} reproduces the duality between the MHV amplitude and the
bosonic Wilson loop, Eq.~\re{MHV-dual}. Notice the appearance of a power of the coupling constant in the right-hand side of
\re{dual-par}
\footnote{Following \cite{Car10}, the factor $a^k$ in the right-hand side of \re{dual-par} can be eliminated through a redefinition
of the odd variables $\eta\to a^{-1/4}\eta$.}
The reason for it is that the perturbative expansion of
$\mathcal{W}_{n;k}$ starts at order $O(a^k)$. Then, the duality relation \re{dual-par} leads to the following iterative
structure of loop corrections: Having computed $\mathcal{W}_n$ to, say, $k-$loops and expanding the result in
powers of $\theta$'s as in \re{W-exp}, we will be able to determine the $k-$loop corrections to the MHV amplitudes, the
$(k-1)-$loop corrections to the NMHV amplitudes and so on until we reach the tree-level expression for
N${}^k$MHV amplitudes.
We observe that, for a given number of external particles $n$, the expansion in \re{gen}terminates at $k=n-4$ and,
therefore, $\widehat{\mathcal{A}}_{n;k}=0$ for $k\ge n-3$. This property is an immediate consequence of the
conventional on-shell Poincar\'e supersymmetry of the all-loop amplitudes in $\mathcal{N}=4$ SYM. Then, the
duality relation \re{dual-par} implies that the top four components of the expansion \re{W-exp} should vanish
to all loops
\begin{align}\label{W=0}
\mathcal{W}_{n;n-3}(x,\theta;a)
= \mathcal{W}_{n;n-2}(x,\theta;a)
= \mathcal{W}_{n;n-1}(x,\theta;a)
= \mathcal{W}_{n;n} (x,\theta;a)\stackrel{?}{=} 0\,.
\end{align}
We recall that the perturbative expansion of $\mathcal{W}_{n;n-3}(x,\theta;a)$ only starts at $(n-3)-$loops and
accordingly for the remaining functions in this relation. Here we inserted the question
mark (see also Eqs.~\re{W-sym}, \re{super-W4} and \re{KW=0} below) to indicate that these relations
have the status of a conjecture. Below we shall check them by explicit one-loop
calculations.
Tree-level amplitudes are free from infrared divergences and, as a result, they enjoy both conventional and
dual superconformal symmetries. Then, the duality relation \re{dual-par} predicts that the lowest $O(a^k)$
correction to $\mathcal{W}_{n;k} (x,\theta;a)$ should respect the dual superconformal symmetry. At loop
level, the scattering amplitudes suffer from infrared divergences and some of the symmetries become anomalous.
It is paramount for our purposes that the dual $Q-$ and $\bar S-$supersymmetries remain unbroken to all
loops. The reason for this is as follows. As it is evident from its definition, the $Q-$supersymmetry generates
shifts of the odd variables $\delta_Q \theta_i{}_\alpha^{A}=\epsilon_\alpha^{A}$. Since the scattering amplitudes
depend on $\theta$'s only through the differences \re{super-mom}, they stay invariant under the action of
$Q-$supersymmetry. In a similar manner, the dual $\bar S-$supersymmetry coincides with the conventional
$\bar q-$supersymmetry of the $\mathcal{N}=4$ Lagrangian. As long as we employ a regularization that
preserves Poincar\'e supersymmetry, $\bar q-$supersymmetry remains unbroken and the same should be true for the dual $\bar S-$supersymmetry, $\delta_{\bar S} (\theta_i)_\alpha^{A}=
(x_i)_{\alpha\dot\alpha}\xi^{\dot\alpha\, A}$.
Thus, the duality relation \re{dual-par} implies
that the supersymmetric Wilson loop should preserve $Q-$ and $\bar S-$supersymmetries to all loops
\begin{align}\label{W-sym}
Q_A^\alpha \, \mathcal{W}_n(x,\theta;a) =\bar S_{\dot\alpha\, A} \, \mathcal{W}_n(x,\theta;a) \stackrel{?}{=} 0\,.
\end{align}
To discuss the consequence of these relations, it is convenient to introduce new variables \cite{hodges}, the
so-called momentum supertwistors $Z^a=(\lambda_{i}, \mu_i \,, \chi_i)$ with
\begin{align}\label{twistors}
\chi_i^A = \vev{i\theta_i^A}\,,\qquad \mu_{i \,\dot\alpha} = \lambda_i^\alpha(x_i)_{\alpha\dot\alpha} \,.
\end{align}
The dual superconformal symmetry acts on $Z^a$ as linear $GL(4|4)$ transformations. In
particular, the action of the dual $Q-$ and $\bar S-$supersymmetries corresponds to
\begin{align}\label{chi-rot}
\chi_i^A \to \chi_i^A + \vev{ i \epsilon^A} + [ \mu_i\,\bar \xi^{A}]\,,
\end{align}
with $\epsilon_\alpha^{A}$ and $\bar \xi^{\dot\alpha A}$ being the parameters of the corresponding
transformations. Using the invariance of the supersymmetric Wilson loop \re{W-sym}
under the transformations \re{chi-rot}, we can always choose the parameters $\epsilon^A_\alpha$
and $\xi^{\dot\alpha \, A}$ in such a manner
as to set 16 components of the Grassmann variables to zero, e.g.,
$
\chi_1^A=\chi_2^A=\chi_3^A=\chi_4^A=0
$.
This immediately implies that the supersymmetric Wilson loop depends on $(n-4)$ odd variables $\chi_i$
and, therefore, its expansion \re{W-exp} terminates at the term $ \mathcal{W}_{n;n-4}$, in agreement with
\re{W=0}. In particular, in the special case of $n=4$, the supersymmetric Wilson
loop $\mathcal{W}_4$ cannot depend on the odd variables and, therefore, it should coincide with the
bosonic Wilson loop
\begin{align}\label{super-W4}
\mathcal{W}_4(x,\theta;a) \stackrel{?}{=} W_4(x;a)\,.
\end{align}
This is in accord with the well-known fact that for $n=4$ particles the only nonzero amplitude in
$\mathcal{N}=4$ SYM is the MHV one.
Infrared divergences break the dual conformal symmetry of the amplitudes making special
conformal $K-$symmetry anomalous, $K^{\dot\alpha\alpha}\widehat{\mathcal{A}}_{n;k}\neq 0$. Since the
infrared divergences have a universal form independent of the helicity configuration of the scattered particles, they
cancel in the ratio of amplitudes $\widehat{\mathcal{A}}_{n}/\widehat{\mathcal{A}}_{n;0}$. As a consequence, dual conformal symmetry gets restored in the ratio,
$K^{\dot\alpha\alpha}\big({\widehat{\mathcal{A}}_{n}/\widehat{\mathcal{A}}_{n;0}}\big)= 0$
\footnote{This property has been conjectured in Ref.~\cite{DrunHenKorSok10} and
was recently verified in Ref.~\cite{KosRoiVer10} by an explicit two-loop calculation of the
$n=6$ NMHV amplitude in $\mathcal{N}=4$ SYM.} Together with the duality
relation \re{dual-par} this leads to the following constraint on the supersymmetric Wilson loop
\begin{align}
\label{KW=0}
K^{\dot\alpha\alpha}\lr{ {\mathcal{W}}_{n}/ \mathcal{W} _{n;0}}\stackrel{?}{=} 0\,.
\end{align}
In other words, the ratio of the supersymmetric Wilson loop and its lowest bosonic component should be invariant under the special conformal transformations in $\mathcal{N}=4$ SYM.
\subsection{Supersymmetric connections}
\label{SUSYconnection}
Let us start with reviewing the construction of the supersymmetric Wilson loop \re{super-W}
introduced in Refs.\ \cite{MasSki10,Car10}. The super Wilson loop in \re{super-W} is given
by the path-ordered product of supersymmetric Wilson lines calculated along the
segments of the polygon contour $\mathcal{C}_n$
\begin{align}\label{W-prod}
\mathcal{W}_n = \frac1{N_c} \vev{ \mathop{\rm tr}\nolimits \lr{ \mathcal{W}_{[1,2]} \mathcal{W}_{[2,3]}\ldots \mathcal{W}_{[n,1]}} } \,.
\end{align}
Here the Wilson line $\mathcal{W}_{[i,i+1]}$ is evaluated along the straight line
connecting the points $(x_i,\theta_i)$ and $(x_{i+1},\theta_{i+1})$ in the chiral superspace
\begin{align}\label{super-path}
x(t_i) = x_i- t_i x_{i,i+1}\,,\qquad \theta(t_i) = \theta_i -t_i \theta_{i,i+1}\,,\qquad
(0\le t_i \le 1)\,,
\end{align}
with implied periodicity conditions $x_{i+n}\equiv x_i$ and $\theta_{i+n}\equiv \theta_i$.
Then, the supersymmetric Wilson line takes the form
\begin{equation}\label{W-link}
\mathcal{W}_{[i,i+1]} = T\exp \left( ig \int_0^1 dt_i\, \mathcal{B}(t_i) \right)\,,
\end{equation}
where the (M)inkowski (S)uperspace connection $\mathcal{B}(t_i)$ is the sum of the bosonic and fermionic connections
projected on the tangent direction to the trajectory,
\begin{equation}
\label{Bsuperfield}
\mathcal{B}^{\rm MS}(t_i)
=
\frac12 \dot x^{\dot\alpha\alpha}(t_i) {\mathcal A}_{\alpha\dot\alpha}
+
\dot \theta^{\alpha A}(t_i){\mathcal F}_{\alpha A}
\, .
\end{equation}
The supersymmetric Wilson loop \re{W-prod} is invariant under non-Abelian gauge transformations
\begin{align}\label{B-gauge}
& \mathcal{B}(t_i) \to U^\dagger(t_i) \mathcal{B}(t_i) U(t_i) + \frac{i}{g} U^\dagger \partial_{t_i} U\,,
\qquad
\mathcal{W}_{[i,i+1]} \to U^\dagger(1)
\mathcal{W}_{[i,i+1]} U(0)\,,
\end{align}
with $U=U(t_i)$.
For the bosonic and fermionic connections the gauge transformations
look as
\begin{align}\label{gauge}
{\mathcal A}_{\alpha\dot\alpha} &\to U^\dagger {\mathcal A}_{\alpha\dot\alpha} U
+ \frac{i}{g} U^\dagger\partial_{\alpha\dot\alpha} U\,,\qquad
\\[2mm] \notag
{\mathcal F}_{\alpha A} &\to U^\dagger {\mathcal F}_{\alpha A}U + \frac{i}{g} U^\dagger \partial_{\alpha A} U
\,,
\end{align}
where $\partial_{\alpha\dot\alpha}\equiv \partial/\partial {x^{\dot\alpha\alpha}}$ and
$\partial_{\alpha A} \equiv \partial/\partial \theta^{\alpha A}$.
Let us now present the explicit expressions for the bosonic, $\mathcal{A}(x,\theta)$,
and fermionic, $\mathcal{F}(x,\theta)$, connections.
Both of them are given by expansions in the Grassmann variable $\theta^{A}_\alpha$ with
the coefficients of increasing scaling dimension built from the various field components
(gluon $A^{\dot\alpha\alpha}$, gauginos $\psi^A_\alpha, \bar\psi_A^{\dot\alpha}$ and
scalars $\phi^{AB}, \bar\phi_{AB} = \ft12 \epsilon_{ABCD} \phi^{CD}$) and their gauge covariant
derivatives $D_{\alpha\dot\alpha}=\partial_{\alpha\dot\alpha} -ig [A_{\alpha\dot\alpha},\ ]$
\begin{align}\label{Afield}
\mathcal{A}
=& \,
A +
i \ket{ \theta^A} [ \bar\psi_{A}|
+
\frac{i}{2 !} \ket{\theta^A} \bra{\theta^B} D \bar\phi_{AB}
-
\frac{1}{3!} \epsilon_{ABCD} \ket{\theta^A} \bra{\theta^B} D \vev{ \theta^C \psi^{D}}
\\[3mm]
& \hfill +
\frac{i}{4!} \epsilon_{ABCD} \ket{\theta^A} \bra{\theta^B} D \vev{ \theta^C | F | \theta^{D}}
+
\dots \, ,
\nonumber
\\[3mm]
\label{Ffield}
\mathcal{F}_A =& \, \frac{i}2 \bar\phi_{AB} \ket{\theta^B}
-
\frac1{3!!}\epsilon_{ABCD}\ket{\theta^B} \vev{\theta^C \psi^D}
+
\frac{i}{4!!} \epsilon_{ABCD}\ket{\theta^B} \vev{\theta^C | F | \theta^D}+\ldots
\, ,
\end{align}
where we used a compact notation for the quantities carrying spinor indices, e.g.
$\ket{\theta^A} \equiv \theta^A_\alpha$ and $[\bar\psi_A|\equiv \bar\psi_{\dot\alpha A}$ (see Appendix \ref{NotationsConvensions} for explanations).
Notice that in the expression for $\mathcal{A}$, Eq.~\re{Afield},
the covariant derivatives act on the quantum fields only. The expansion of $\mathcal{A}$
starts with the gauge field and the additional factor $1/2$ in front of the bosonic connection in
\re{Bsuperfield} is introduced to get the correct normalization for the lowest component, i.e.,
$\ft{1}{2} dx^{\dot\alpha\alpha} A_{\alpha\dot\alpha} = dx^\mu A_\mu$. The ellipses in the
right-hand sides of \re{Afield} and \re{Ffield} denote higher-order terms in the $\theta$'s up to order
$O(\theta^8)$. A characteristic feature of such terms
is that, in the interaction-free limit (for $g = 0$), they are proportional to the free equations of motion.%
\footnote{This is obvious from the supersymmetric transformations of the holomorphic gluon
field strength
$
\delta_{Q} F^{\alpha\beta} =
\left[
\epsilon^{\alpha A} (\bar\Omega_{\rm f})^{\beta}_{A}
+
\epsilon^{\beta A} (\bar\Omega_{\rm f})^{\alpha}_{A} \right]/4 +O(g)
$ with
$
(\bar\Omega_{\rm f})^{\alpha}_{A} = \partial^\alpha{}_{\dot\alpha} \bar\psi^{\dot\alpha}_A
$.}
Thus, the free-theory expansion of the bosonic and fermionic connections terminates at orders
$O(\theta^4)$ and $O(\theta^3)$, respectively.
The form of the connections and the values of the rational coefficients in \re{Afield} and
\re{Ffield} are fixed from the requirement for the supersymmetric Wilson loop \re{W-prod}
to be invariant under the shift of the odd variables, $\delta \theta_\alpha^A=\epsilon_\alpha^A$, and the simultaneous chiral supersymmetric transformation of fields (see Eq.~\re{Q-susy} in Appendix \ref{NotationsConvensions}). It is a well-known feature of all supersymmetric gauge theories, considered in a non-supersymmetric gauge (Wess-Zumino gauge), that the supersymmetry algebra closes modulo compensating gauge transformations with a field-dependent parameter. In addition, in the absence of auxiliary fields, as is the case of $\mathcal{N}=4$ SYM, the algebra closes on the shell of the field equations. An explicit calculation
yields the following result for the $Q-$variation of both connections
\begin{align}
\label{SUSYtransformAandF}
\delta_Q \mathcal{A}_{\alpha\dot\alpha} & = \partial_{\alpha\dot\alpha} \omega + i g [\omega, {\mathcal A}_{\alpha\dot\alpha}]
+ \Omega_{\alpha\dot\alpha}\,,\qquad
\\[2mm]\notag
\delta_Q{\mathcal F}_{\alpha A} & = \partial_{\alpha A}\omega + i g [\omega, {\mathcal F}_{\alpha A} ]
\, ,
\end{align}
with $\omega$ being the field-dependent gauge transformation parameter
\begin{equation}
\label{Gaugeomega}
\omega
=
\vev{\epsilon^{A} \theta^B}\left[
- \frac{i}{2!} \bar\phi_{AB}
+
\frac{1}{3!} \epsilon_{ABCD} \vev{\theta^C \psi^D}
-
\frac{i}{4!}
\epsilon_{ABCD} \vev{\theta^C | F | \theta^D }
+
\dots
\right]
\, ,
\end{equation}
and $\epsilon^A_\alpha$ being the parameter of the $Q-$transformations. Also, $ \Omega_{\alpha\dot\alpha}$
in the first relation in \re{SUSYtransformAandF} is given by
\begin{eqnarray}
\label{EOMOmega}
\Omega_{\alpha\dot\alpha}
=
-
2
\epsilon_{ABCD}
(\epsilon^{\beta A} \theta_\beta^B) \theta^C_\alpha
\left[
\frac{1}{3!} (\Omega_{\rm f})_{\dot\alpha}^D
-
\frac{i}{4!} \theta_\gamma^D (\Omega_{\rm g})_{\dot\alpha}{}^\gamma
+ \dots
\right]
\, ,
\end{eqnarray}
and it
is proportional to the fermion and gluon equations of motion
\begin{equation}
\label{EOMoperators}
(\Omega_{\rm f})_{\dot\alpha}^A
= D_{\dot\alpha}{}^\gamma \psi^A_\gamma
\, , \qquad
(\Omega_{\rm g})_{\dot\alpha}{}^\alpha
= D_{\dot\alpha}{}^\gamma F_\gamma{}^\alpha
\, .
\end{equation}
Let us now examine the variation of the supersymmetric Wilson loop under the supersymmetric
transformations \re{SUSYtransformAandF}. Comparing the relations \re{SUSYtransformAandF}
and \re{gauge}, we observe that the $\omega-$dependent terms in \re{SUSYtransformAandF}
can be eliminated in $\mathcal{W}_n$ by a compensating (infinitesimal) gauge transformation
\re{gauge} with the parameter $U=1+ig \omega$. Thus, we are left over with the variation of the
bosonic connection proportional to the equations of motion, $\delta_Q \mathcal{A}_{\alpha\dot\alpha}
=\Omega_{\alpha\dot\alpha}$. Although such terms vanish on shell, that is for quantum fields
satisfying their equations of motion, they do provide a nonvanishing contribution to the variation of
$\mathcal{W}_n$ under off-shell supersymmetry transformations. As we will see in a moment,
this subtlety plays a very important r\^ole in testing the duality between supersymmetric Wilson
loops and scattering amplitudes.
\subsection{Equivalent form of the supersymmetric Wilson loop}
As was explained in Sect.\ \ref{Matching}, to discuss the duality between supersymmetric
Wilson loops and superamplitudes it is convenient to switch
from the $(x,\theta)-$coordinates to the momentum supertwistor $(\lambda,\mu,\chi)-$variables
defined in \re{twistors}. Inverting the relations \re{twistors} we find with the help of
\re{super-mom}
\begin{align}\label{EtaToChi}
x_i^{\dot\alpha\alpha}
=
\frac{\mu_{i-1}^{\dot\alpha}\lambda_i^\alpha -\mu_{i}^{\dot\alpha}\lambda_{i-1}^\alpha }{\vev{i-1\, i}}
\,,\qquad
\theta_i^{\alpha A} = \frac{\chi_{i-1}^A \lambda_i^\alpha - \chi_i^A \lambda_{i-1}^\alpha}{
\langle i-1\, i \rangle}\,.
\end{align}
Rewriting the supersymmetric Wilson loop \re{W-prod} and \re{W-link} in terms of these variables,
we obtain the equivalent form of $\mathcal{W}_n$ introduced in Ref.\ \cite{Car10}.
To begin with, we examine the expression for the Wilson line \re{W-link} and
replace the bosonic and fermionic connections in \re{Bsuperfield} by their
explicit expressions \re{Afield} and \re{Ffield}. Then, we use the definition
\re{EtaToChi} to get, after a rather lengthy calculation,
\begin{equation}
\label{MStoCH}
\mathcal{B}^{\rm MS} (t_j)
=
\mathcal{E}_j (t_j)
+ \lr{
\frac{d}{dt_j} \mathcal{V}_j (t_j) + ig [\mathcal{V}_j (t_j), \mathcal{E}_j (t_j) ]}
+
\Delta \Omega_j (t_j)
\, ,
\end{equation}
where the three terms in the right-hand side have the following meaning. The first term
$ \mathcal{E}_j (t_j)$ depends on the single odd variable $\chi_j^A$ and its expansion
runs to order $O(\chi_j^4)$ only
\begin{eqnarray}
\label{CHexponentE}
2 \mathcal{E}_j (t)
&=&
- \langle j | A | j ]
-
i \chi_j^A [ \bar\psi_A | j ]
+\ft{i}{2!}
\chi_j^A \chi_j^B
\frac{\langle j-1 | D | j ]}{\langle j-1 \,j \rangle} \bar\phi_{AB}
\nonumber\\
&-&
\ft{1}{3!}
\epsilon_{ABCD} \chi_j^A \chi_j^B \chi_j^C
\frac{\langle j-1 | D | j ] \langle \psi^D | j - 1\rangle}{\langle j-1\, j \rangle^2}
\nonumber\\
&+&
\ft{i}{4!}
\epsilon_{ABCD} \chi_j^A \chi_j^B \chi_j^C \chi_j^D
\frac{\langle j-1 | D | j ] \langle j-1 | F | j-1 \rangle}{\langle j-1\, j \rangle^3}
\, ,
\end{eqnarray}
where we used the compact spinor notations explained in Appendix \ref{NotationsConvensions}.
The second term in the right-hand side of \re{MStoCH} involves the covariant derivative
of the function
\begin{eqnarray}\label{Vj}
\mathcal{V}_j
\!\!\!&=&\!\!\!
\ft{i}{2!}
\chi_j^A \frac{\langle j-1, \theta^B \rangle}{\langle j-1\, j \rangle} \bar\phi_{AB}
-
\ft{1}{3!}
\epsilon_{ABCD} \chi_j^A
\frac{\langle j-1, \theta^B \rangle}{\langle j-1\, j \rangle^2}
\langle \Lambda_j^C | \psi^D \rangle
\\ \nonumber
&+&\!\!\!
\ft{i}{4!} \epsilon_{ABCD} \chi_j^A
\frac{\langle j-1, \theta^B \rangle}{\langle j-1\, j \rangle^3}
\Big\{
\langle \Lambda_j^C | F | \Lambda_j^D \rangle
+
\chi^C_j \langle j-1 | F | \Lambda_j^D \rangle
+
\chi^C_j \chi^D_j \langle j-1 | F | j-1 \rangle
\Big\}
+ \dots \,,
\end{eqnarray}
where $\theta^A=\theta^A(t_j)$ was defined in \re{super-path} and
a shorthand notation was introduced for the combination
\begin{equation}
\ket{\Lambda_j^{C}} = \ket{j} \langle j-1| \theta^C(t_j) \rangle - 2 \ket{j-1} \chi_j^C
\, .
\end{equation}
The contribution of $\mathcal{V}_j(t_j)$ to the Wilson line \re{W-link}
can be factored out into boundary terms depending on $\mathcal{V}_j(t_j=1)$ and $\mathcal{V}_j(t_j=0)$. In the formulation
of Ref.~\cite{Car10}, such terms get absorbed into the so-called
vertex operators localized at the vertices of the super-polygon $\mathcal{C}_n$.
Finally, the last term in the right-hand side of \re{MStoCH} is proportional to the
equations of motion
\begin{align}\label{EOMinMSminusCH}
2\Delta \Omega_j
=
- \ft{1}{3!}
\epsilon_{ABCD} \chi_j^A \chi_j^B
\frac{\langle j-1 \,\theta^C \rangle}{\langle j-1\, j \rangle}
[ \Omega_{\rm f}^{D} j ]
-
\ft{i}{4!}
\epsilon_{ABCD} \chi_j^A \chi_j^B
\frac{\langle j-1 \,\theta^C \rangle}{\langle j-1\, j \rangle^2}
\langle \Lambda^C_j | \Omega_{\rm g} | j ]
\, ,
\end{align}
where $\Omega_{\rm f}$ and $\Omega_{\rm g}$ were defined in \re{EOMoperators}.
Neglecting the contribution of the latter to \re{MStoCH}, we can define another connection
\begin{align}
\label{B-CH}
\mathcal{B}^{\rm CH}(t_j)
=
\mathcal{E}_j (t_j)
+
\lr{
\frac{d}{dt_j} \mathcal{V}_j (t_j) + ig [\mathcal{V}_j (t_j), \mathcal{E}_j (t_j) ]}
\, .
\end{align}
Then, we can use this function to construct the Wilson line \re{W-link} and
define the corresponding supersymmetric Wilson loop \re{W-prod}. The resulting
expression for $\mathcal{W}_n^{\rm CH}$ coincides with the supersymmetric
Wilson loop introduced in Ref.~\cite{Car10}. However, since by construction
\begin{align}
\label{B-diff}
\mathcal{B}^{\rm MS}(t_j)-\mathcal{B}^{\rm CH}(t_j) = \Delta \Omega_j (t_j)\,,
\end{align}
the two Wilson loops proposed in Refs.~\cite{MasSki10,Car10} are identical only on shell. Thus, the difference between the all-loop expressions for the two Wilson loops can be
attributed to the contribution of the field-equation operators.
As a consequence of our analysis, the natural question arises whether the equations of motion can
produce a nonvanishing contribution to the expectation value of the supersymmetric Wilson
loop \re{super-W}. If they do not, then the two Wilson loops $\mathcal{W}_n^{\rm MS}$ and
$\mathcal{W}_n^{\rm CH}$ respect the $Q-$supersymmetry and they are identical at the quantum
level. However, if the contribution of the equations of motion is different from zero, then
$\mathcal{W}_n^{\rm MS}\neq \mathcal{W}_n^{\rm CH}$ and both Wilson loops are not $Q-$supersymmetry invariants. In the latter case, neither of the Wilson loops can be dual to the
superamplitudes in $\mathcal{N}=4$ SYM since the latter are invariant under the
$Q-$supersymmetry.
\section{Supersymmetric Ward identities}
\label{SUSYWI}
To understand better what happens to supersymmetry at the quantum level, let us derive the Ward identities
for the two supersymmetric Wilson loops introduced in the previous section.
We start with the path integral representation for the vacuum expectation value of the Wilson loop
\begin{equation}\label{path}
\langle \mathcal{W}_n(x_i,\theta_i) \rangle = \int [DX] \, \mathcal{W}_n(x_i,\theta_i) \, {\rm e}^{i S_{\mathcal{N}=4}[X]}
\, ,
\end{equation}
where the integration goes over all fields in $\mathcal{N}=4$ SYM collectively called $X$. Then, we make the shift
$\theta_i\to \theta_i+\epsilon$ in both sides of \re{path} and perform a compensating supersymmetry transformation
of the fields \re{Q-susy} inside the path integral, $X\to X+\delta_Q X$.
Making use of the invariance of the $\mathcal{N}=4$ action, $S_{\mathcal{N}=4}[X]=S_{\mathcal{N}=4}[X+\delta X]$
we arrive at %
\footnote{More precisely, the $Q-$variation of the gauge-fixed action is different from zero, but for the supersymmetry
preserving regularization, i.e., dimensional reduction, $\delta_Q S_{\mathcal{N}=4}$ contains BRST-exact operators
only and thus does not produce nontrivial
contributions when inserted into the correlation function with the gauge-invariant Wilson loop,
$\langle \mathcal{W}_n \left( i \delta_Q S_{\mathcal{N}=4} \right)\rangle = 0$ (see, e.g., \cite{BelMul00}). }
\begin{align}\label{Ward}
(\epsilon\cdot Q) \langle \mathcal{W}_n \rangle \equiv
\sum_{i=1}^n \epsilon^A_\alpha \frac{\partial}{\partial \theta_{i \alpha}^A} \langle \mathcal{W}_n \rangle
=
\langle \delta_Q \mathcal{W}_n \rangle
\, .
\end{align}
To make use of this relation we have to analyze the variation of
the supersymmetric Wilson loops $\delta_Q \mathcal{W}_n$, which we come to do next.
It follows from the definitions \re{W-prod} and \re{W-link} that the
supersymmetry variation of the Wilson loop $\mathcal{W}_n$ amounts to inserting a
local operator $\delta \mathcal{B}(t_j)$ on the super-polygon $\mathcal{C}_n$
\begin{align}\notag
\vev{\delta_Q \mathcal{W}_n}
&= \sum_{j=1}^n \frac1{N_c} \vev{ \mathop{\rm tr}\nolimits \lr{ \mathcal{W}_{[1,2]} \ldots \delta_Q \mathcal{W}_{[j,j+1]} \ldots \mathcal{W}_{[n,1]}} }
\\\label{delta-W}
&= \sum_{j=1}^n\, \frac{1}{N_c} \vev{\mathop{\rm tr}\nolimits \lr{ig \int_0^1 dt_j\, \delta_Q \mathcal{B}(t_j)} \mathcal{W}_n(t_i) }\,.
\end{align}
Here $\mathcal{W}_n(t_i)$ stands for the supersymmetric Wilson line
evaluated along an open contour in superspace that starts at the point
$(x(t_i),\theta(t_i))$, goes along the polygon $\mathcal{C}_n$ and returns to the
starting point.
\subsection{Anomalies}
Let us first compute $\delta_Q \mathcal{B}(t_j)$ for the supersymmetric Wilson loop
$\mathcal{W}_n^{\rm MS}$ defined in \re{Bsuperfield}, \re{Afield} and \re{Ffield}. We apply
the relation \re{SUSYtransformAandF} to get
\begin{align}\notag
\delta_Q \mathcal{B}^{\rm MS}(t_j)
&= \ft12 \dot x^{\dot\alpha\alpha}(t_j)
\delta_Q {\mathcal A}_{\alpha\dot\alpha}+\dot \theta^{\alpha A}(t_j) \delta_Q{\mathcal F}_{\alpha A}
\\ \label{B-con}
&=\frac{d \omega}{dt_j} +i g [\omega, \mathcal{B}^{\rm MS}(t_j)]
-\ft12 \bra{j} \Omega(t_j) |j]\,,
\end{align}
with $\omega$ and $\Omega$ defined in \re{Gaugeomega} and \re{EOMOmega}, respectively. As
was explained in Sect.~\ref{SUSYconnection}, the terms involving $\omega$ do not contribute to the
right-hand side of \re{delta-W} by virtue of gauge invariance. Therefore, computing \re{delta-W} we can
retain only the last term in \re{B-con}
\begin{align}\label{Omega-MS}
\delta_Q \mathcal{B}^{\rm MS}(t_j) = \Omega^{\rm MS}_j
= \epsilon_{ABCD}
\langle \epsilon^A \theta^B \rangle \vev{j\, \theta^C}
\bigg\{
\frac{1}{3!} [ \Omega_{\rm f}^D j]
+
\frac{i}{4!} \bra{\theta^D} \Omega_{\rm g}{}|j] + \dots
\bigg\}\,,
\end{align}
where $\theta^A=\theta^A(t_j)$ was defined in \re{super-path} and the
ellipses denote terms vanishing in the free theory limit (i.e., for $g=0$).
The analysis of the Wilson loop $\mathcal{W}_n^{\rm CH}$ goes along the same lines.
We use the relations \re{CHexponentE} and \re{B-CH} to verify that, modulo gauge transformations,
\begin{align}\nonumber
\delta_Q \mathcal{B}^{\rm CH}(t_j) = \Omega^{\rm CH}_j
& =
\frac{1}{4} \epsilon_{ABCD} \frac{\chi_j^A \chi_j^B \langle j-1 \epsilon^C \rangle}{\langle j-1\, j\rangle}
[ \Omega_{\rm f}^{ D} j ]
\\& \label{Omega-CH}
-
\frac{i}{12} \epsilon_{ABCD}
\frac{\chi_j^A \chi_j^B \chi_j^C \langle j-1 \epsilon^D \rangle}{\langle j-1\, j\rangle^2}
\langle j-1 | \Omega_{\rm g} | j ] + \ldots
\, ,
\end{align}
where $\chi_j^A=\vev{j\theta_j^A}$. The obtained expressions for the anomalies have to satisfy a consistency condition that follows from \re{B-diff}
\begin{align}
\delta_Q \mathcal{B}^{\rm MS}-\delta_Q \mathcal{B}^{\rm CH}=\Omega^{\rm MS}_j-\Omega^{\rm CH}_j=\delta_Q \Delta\Omega_j(t_j) \,.
\end{align}
Replacing $\Omega_j(t_j)$ by its explicit expression \re{EOMinMSminusCH} we find after lengthy
calculations that this relation is indeed satisfied.
Substituting the obtained expressions for $\delta_Q \mathcal{B}$ into \re{delta-W} and \re{Ward} we
deduce the Ward identity for the supersymmetric Wilson loop
\begin{align}
\label{susyWI}
(\epsilon \cdot Q) \langle \mathcal{W}_n \rangle = \sum_{j=1}^n\, ig
\int_0^1 dt_j\, \VEV{\frac{1}{N_c}\mathop{\rm tr}\nolimits \left[{ \Omega_j(t_j)} \mathcal{W}_n(t_j)\right] }\,,
\end{align}
where $\Omega_j(t_j)$ is given by \re{Omega-MS} and \re{Omega-CH} for the two supersymmetric
Wilson loops defined above and the operator $(\epsilon\cdot Q)$ admits two equivalent forms
\begin{align}\label{Q-oper}
(\epsilon \cdot Q) = \sum_{j=1}^n \epsilon^A_\alpha \frac{\partial}{\partial \theta_{j \alpha}^A}
= \sum_{j=1}^n \vev{i \epsilon^A} \frac{\partial}{\partial \chi_{j}^A} \,.
\end{align}
We would like to emphasize that that the right-hand side of \re{susyWI} involves
the correlation function of the equations of motion with the Wilson loop.
For the Wilson loop to be invariant under the $Q-$supersymmetry, the right-hand side
of \re{susyWI} should vanish. Otherwise, the $Q-$supersymmetry will be
broken.
\subsection{One-loop calculation}
Let us now compute the one-loop corrections to the right-hand sides of the Ward
identities \re{susyWI} for both Wilson loops.
We recall that the anomalies \re{Omega-MS} and \re{Omega-CH} are given by
a linear combination of the fermion and gluon equations of motion, $\Omega_{\rm f}$
and $\Omega_{\rm g},$ defined in \re{EOMoperators}.
We start with \re{Omega-MS} and
retain for the moment only the terms involving $\Omega{}_{{\rm f}\, \dot\alpha}^A = \partial_{\dot\alpha}{}^\gamma
\psi^A_\gamma+O(g)$. Their contribution to the first term in the sum, i.e., for $j=1$, in the right-hand side of
\re{susyWI} is given by (the remaining terms can be obtained through a cyclic shift of the indices)
\begin{align}\label{psi-B}
\frac{ig}{3!}\epsilon_{ABCD}\int_0^1dt_1
\langle \epsilon^A \theta^B(t_1) \rangle \vev{1 \theta^C(t_1)}
\VEV{\frac{1}{N_c}\mathop{\rm tr}\nolimits \left\{[ \Omega_{\rm f}^D(x(t_1)) , 1] \lr{ig\int_{\mathcal{C}_n} dt\, \mathcal{B}(t)}\right\} } +O(g^4)\,,
\end{align}
where we replaced the Wilson line $\mathcal{W}_n(t_1)$ by its lowest order expansion. Here the connection $\mathcal{B}(t)$ is
integrated over the contour $\mathcal{C}_n$ and the fermionic operator is inserted at the point
$y=x(t_1)$ on the segment $[x_1,x_2]$. Computing the correlation function in \re{psi-B} to the lowest order in the coupling, we have to Wick contract $\psi^A_\beta(x(t_1))$
from $\Omega_{\rm f}^D$ with the fermion field $\bar\psi_{\dot\beta B}(x(t))$ inside $\mathcal{B}(t)$.
It is easy to see from \re{MStoCH}, \re{CHexponentE} and \re{B-CH} that the corresponding terms look alike for both
connections and for the $k$th segment they are given by
\begin{align}\label{part-B}
\mathcal{B}(t_k) = -\frac{i}2 \chi_k^A\, [ \bar\psi_{A}(x(t_k))\, k] +\ldots
\end{align}
In Eq.~\re{psi-B}, the $\bar\psi-$field can be located at any of
the segments of the contour $\mathcal{C}_n$ including $[x_1,x_2]$. In that case,
the two fields $\psi^A_\beta(x(t_1))$ and $\bar\psi_{\dot\beta B}(x(t_k))$ belong to the
same segment and they become light-like separated (recall that $x_{12}^2=0$).
Due to light-cone singularities, the product of two quantum fields separated by
a light-like interval is not well defined. Therefore, in order to
define the corresponding contribution to \re{psi-B} we have to introduce a regularization.
In what follows we shall use the Four-Dimensional Helicity (FDH) regularization of $\mathcal{N}=4$ \cite{Bern:2002zk}.
The main advantage of this scheme is that it preserves the Poincar\'e supersymmetry of $\mathcal{N}=4$ SYM
(at least to the lowest order in coupling as we do in our analysis) and allows
us to use the spinor decomposition for super-momenta \re{super-mom} without it interfering with the change of dimensionality
of Minkowski space-time. Then, the regularized correlator of the gaugino field with the fermion equation
of motion takes the form
\begin{align}\label{fermi-cor}
\frac1{N_c} \mathop{\rm tr}\nolimits \langle \partial_{\dot\alpha}{}^\beta \psi^A_\beta(y) \bar\psi_{\dot\beta B} (x) \rangle
=
i {C_F} \delta^A_B \,
\frac{\varepsilon \Gamma (2 - \varepsilon)}{\pi^{2 - \varepsilon}}
\frac{\epsilon_{\dot\alpha \dot\beta}}{[- (x-y)^2+i0]^{2 - \varepsilon}}
\, ,
\end{align}
where $C_F=(N_c^2-1)/(2N_c)$ is the quadratic Casimir of the $SU(N_c)$ gauge group.
According to the standard prescription for tadpoles \cite{Col82}, the correlation function \re{fermi-cor} vanishes
for $(x-y)^2=0$ within the framework of dimensional regularization. This implies that
\re{psi-B} receives zero contribution when $\mathcal{B}(t)$ is integrated
along the segment $[x_1,x_2]$.
Notice that the two-point function \re{fermi-cor} is proportional to the parameter
of dimensional regularization $\varepsilon$. Therefore, for the correlation function
in \re{psi-B} to be different from zero as $\varepsilon\to 0$ the integration over
the position of the fields in \re{psi-B} should produce a pole $1/\varepsilon$.
As follows from \re{fermi-cor}, this could only happen if the two-point function
\re{fermi-cor} is integrated through a
region where $(x-y)^2\to 0$. As an example, let us consider the contribution to
\re{psi-B} when the connection $\mathcal{B}(t)$ is integrated along the segment $[x_2,x_3]$, that is for $k=2$ in \re{part-B}
\begin{align} \label{F-exam}
\text{Eq.\,\re{psi-B}} &= -
\frac{1}{3!}\frac{g^2 C_F}{2\pi^{2 - \varepsilon}}\epsilon_{ABCD} \chi_1^C \chi_2^D [12] \int_0^1dt_1
\int_0^1dt_2
\frac{ \varepsilon \Gamma (2 - \varepsilon)\langle \epsilon^A \theta^B(t_1) \rangle}{[- x_{13}^2 (1-t_1) t_2+i0]^{2 - \varepsilon}}\,,
\end{align}
where $x(t_i)=x_i-t_ix_{i,i+1}$, $\theta(t_i) = \theta_i-t_i\theta_{i,i+1}$ and we used of the identities
$\vev{i\,\theta^A(t_i)} =\chi_i^A$ and
\begin{equation}
(x(t_1)-x(t_2))^2= (x_{12}(1-t_1) + x_{23} t_2 )^2 = x_{13}^2 (1-t_1) t_2\,.
\end{equation}
Here it is crucial that the two segments are light-like, $x_{i,i+1}^2=0$.
We observe that the integration in the right-hand side of \re{F-exam} around $t_2=0$, $t_1=1$
produces a pole $1/\varepsilon$. It compensates the factor of $\varepsilon$ coming from \re{fermi-cor}
and produces a finite contribution
\begin{align}
\text{Eq.\,\re{psi-B}}=
\frac{1}{3!}\frac{g^2 C_F}{2\pi^2} \frac{[12]}{(x_{13}^2)^2}\epsilon_{ABCD}
\langle \epsilon^A \theta_{12}^B \rangle \chi_1^C \chi_2^D
+
O(\varepsilon)\,.
\end{align}
This example illustrates the general mechanism which is at work for the fermion equations of motion and which produces a
nonvanishing contribution to the right-hand side of \re{susyWI}.
Repeating the analysis for $k=3,\ldots,n$ in \re{part-B} it is straightforward to
show that the right-hand side of \re{psi-B} only receives a nonzero contribution
from four segments of the contour $\mathcal{C}_n$ adjacent to $[x_1,x_2]$,
that is from those with $k=2,3,n-1,n$.
The analysis of the contribution from the gluon equations of motion to the right-hand
side of \re{susyWI} goes along the same lines. To lowest order in the coupling,
the gluon equation-of-motion operator (\ref{EOMoperators}) entering $\Omega_j(t_j)$
can only interact with the bosonic part of $\mathcal{W}_n$ given by
$\ft{i}2\oint_{\mathcal{C}_n} dx^{\dot\alpha\alpha}
A_{\alpha\dot\alpha} (x)$. The correlation function of the gauge
field with its equation-of-motion operator (\ref{EOMoperators}) reads, in the Feynman
gauge and to lowest order in the coupling,
\begin{equation}\label{A-Omega}
\frac1{N_c} \mathop{\rm tr}\nolimits \langle A_{\alpha\dot\alpha} (x_1) \Omega{}_{{\rm g} \, \dot\beta}{}^\beta (x_0) \rangle
= i \frac{ C_F}{2 \pi^{2 - \varepsilon}} {\Gamma (2 - \varepsilon)}
\bigg[
(1 + \varepsilon)
\frac{\epsilon_{\dot\alpha\dot\beta} \delta_\alpha^\beta}{(- x_{01}^2+i0)^{2 - \varepsilon}}
+
{(1 - \varepsilon)}
\frac{(x_{01})_{\dot\beta\alpha} (x_{01})_{\dot\alpha}{}^\beta}{(-x_{01}^2+i0)^{3 - \varepsilon}}\bigg]
\, .
\end{equation}
Making use of this relation we find the correlation function of the gluon equation
of motion with the bosonic Wilson loop as
\begin{eqnarray}
\frac1{N_c} \mathop{\rm tr}\nolimits \label{GluonEOMpropagator}
\oint_{\mathcal{C}_n} dx^{\dot\alpha\alpha}
\langle A_{\alpha\dot\alpha} (x) \Omega{}_{{\rm g} \, \dot\beta}{}^\beta (x_0) \rangle
&=&
i \frac{C_F}{ \pi^{2 - \varepsilon}}\,
{\varepsilon \Gamma (1 - \varepsilon)}\sum_{j=1}^n
(x_{j,j+1})_{\dot\beta}{}^\beta D_\varepsilon (x_{j+1,0}, x_{j0})\,,
\end{eqnarray}
where the notation was introduced for
\begin{equation}
D_\varepsilon (x_{j+1,0}, x_{j0})
=
\frac{(- x_{j+1,0}^2+i0)^{\varepsilon - 1} - (- x_{j0}^2+i0)^{\varepsilon - 1}}{x_{j+1,0}^2 - x_{j0}^2}
\, .
\end{equation}
In distinction with \re{A-Omega}, the relation \re{GluonEOMpropagator}
is gauge invariant.
Comparing \re{GluonEOMpropagator} with \re{fermi-cor} we observe the same
pattern. Though both correlation functions vanish for $\varepsilon\to 0$, they
produce a nonvanishing contribution to the right-hand side of \re{susyWI} upon
integration over the polygon $\mathcal{C}_n$. We would like to emphasize that one may miss
this contribution if one performs the calculation in $D=4$ dimensions without properly regularizing the light-cone singularities of
the correlation functions.
\subsection{Supersymmetry anomalies}
It becomes straightforward to compute the one-loop correction to the right-hand side of \re{susyWI}
for both Wilson loops, $\mathcal{W}_n^{\rm MS}$ and $\mathcal{W}_n^{\rm CH}$, by making use
of the relations \re{fermi-cor} and \re{GluonEOMpropagator}. For the sake of simplicity we present here the
explicit expressions for the simplest case of $n=4$, that is for the Wilson loop defined over the rectangular
contour $\mathcal{C}_4$ in the superspace. The generalization of our analysis to arbitrary $n$ is straightforward.
For the Wilson loop $ \mathcal{W}^{\rm CH}_4$ involving the superconnection \re{B-CH} the
supersymmetric Ward identity reads
\begin{align}\label{CH-anom}
(\epsilon \cdot Q) \, \mathcal{W}^{\rm CH}_4 =
- \frac{g^2 C_F}{4 \pi^2} \epsilon_{ABCD}
\chi_1^A \chi_1^B \frac{\langle 4\epsilon^C \rangle}{\langle 41 \rangle}
\frac{[13]}{x_{13}^2 x_{24}^2}
\left( \chi_3^D + \frac{1}{3} \frac{\langle 34 \rangle}{\langle 41 \rangle} \chi_1^D \right)
+
{\rm (cyclic)} ,
\end{align}
where `(cyclic)' stands for terms obtained by the cyclic shift of indices $i\to i+1$ subject to the periodicity
condition $i+4\equiv i$. Here the first and second terms inside the braces arise from the fermion
and gauge equations of motion, respectively.
For the Wilson loop $ \mathcal{W}^{\rm MS}_4$ involving the superconnection \re{Bsuperfield} the
analogous supersymmetric Ward identity looks as
\begin{align}
(\epsilon \cdot Q)\, \mathcal{W}^{\rm MS}_4
& =
\frac{g^2 C_F}{2 \pi^2}
\frac{1}{3!} \epsilon_{ABCD}
\bigg\{
\langle \epsilon^A \theta_{12}^B \rangle \langle 1 \theta_1^C \rangle
\left(
\frac{[12]}{(x_{13}^2)^2} \langle 2 \theta_2^D \rangle
+
\frac{[41]}{(x_{24}^2)^2} \langle 4 \theta_4^D \rangle
\right)
\nonumber\\
& \qquad\qquad \qquad
-
\frac{[13]}{x_{13}^2 x_{24}^2}
\left( \langle \epsilon^A \theta_1^B \rangle + \langle \epsilon^A \theta_2^B \rangle \right)
\langle 1 \theta_1^C \rangle \langle 3 \theta_3^D \rangle
\bigg\}
\nonumber\\
&-
\frac{g^2 C_F}{2 \pi^2}
\frac{1}{4!} \epsilon_{ABCD}
\bigg\{
\frac{[12]}{(x_{13}^2)^2}
\left(
\langle \epsilon^A \theta_{12}^B \rangle \langle 1 \theta_2^C \rangle \langle 2 \theta_2^D \rangle
+
\langle \epsilon^A \theta_2^B \rangle \langle 1 \theta_2^C \rangle \langle 2 \theta_{12}^D \rangle
\right)
\nonumber\\
& \qquad\qquad\qquad
-
\frac{[13]}{x_{13}^2 x_{24}^2}
\left(
\langle \epsilon^A \theta_1^B \rangle \langle 1 \theta_1^C \rangle \langle 3 \theta_1^D \rangle
+
\langle \epsilon^A \theta_2^B \rangle \langle 1 \theta_2^C \rangle \langle 3 \theta_2^D \rangle
\right)
\nonumber\\
\label{MS-anom}
& \qquad\qquad\qquad
+
\frac{[41]}{(x_{24}^2)^2}
\left[
\langle \epsilon^A \theta_{12}^B \rangle \langle 1 \theta_1^C \rangle \langle 4 \theta_1^D \rangle
+
\langle \epsilon^A \theta_1^B \rangle \langle 1 \theta_1^C \rangle \langle 4 \theta_{12}^D \rangle
\right]
\bigg\}
+
{\rm (cyclic)}
\, .
\end{align}
Here the two terms in the right-hand side again describe the contribution of the fermion and gluon
equations of motion, respectively.
The following comments are in order.
The very fact that the right-hand sides of the relations \re{CH-anom} and
\re{MS-anom} are different from zero immediately implies that the chiral
$Q-$supersymmetry of both Wilson loops is broken already at one loop.
Moreover, it can be verified that the two anomalies do not respect the
conformal symmetry (see Sect.\ \ref{SolvingWI} below).%
\footnote{In general, for some quantity depending on $\vev{ij}$ and $[ij]$
to be invariant under conformal transformations, the indices should satisfy the
conditions $|i-j|=1$. This is not the case for the one-loop expressions for the anomalies, Eqs.~\re{CH-anom}
and \re{MS-anom}, which involve the square brackets $[13]$.}
We recall that, if the supersymmetric Wilson loop $\mathcal{W}_4$ respected the
supersymmetry and conformal symmetry, it would be independent of the odd variables,
Eq.~\re{super-W4}. In the next section, we will reconstruct the complete one-loop expression
for the Wilson loops $\mathcal{W}_4$ by solving the Ward identities \re{CH-anom} and
\re{MS-anom} and we will demonstrate that relation
\re{super-W4} is invalidated by the anomalies.
We established in this section, that the correlation functions of the fermion
and gluon equation of motion with the Wilson loop
induce a nontrivial contribution to the supersymmetric Ward identity. Let us now show
that the same correlation functions allow us to compute the difference
between the two Wilson loops under consideration, $\mathcal{W}_{4}^{\rm MS} -\mathcal{W}_{4}^{\rm CH}$.
Indeed, it follows from \re{B-diff} that to lowest order in the coupling
\begin{align}\label{W-W}
\mathcal{W}_{4}^{\rm MS} -\mathcal{W}_{4}^{\rm CH} =
ig\int_0^1dt_1 \VEV{\frac1{N_c}\mathop{\rm tr}\nolimits\left[ \Delta \Omega_1 (t_1) \lr{ig\int_{
\mathcal{C}_4} dt\, \mathcal{B}(t)}\right]} +\text{(cyclic)}\,,
\end{align}
where $\Delta \Omega_1 (t_1)$ is defined in \re{EOMinMSminusCH}. The latter is given
by a linear combination of fermion and gluon equations of motion.
As a result, the calculation of the correlation function in the right-hand side
of \re{W-W} can be easily performed along the same lines as in
\re{susyWI} and \re{psi-B}. Going through the derivation we find
\begin{eqnarray}
\mathcal{W}_{4}^{\rm CH}
\!\!\!&-&\!\!\!
\mathcal{W}_{4}^{\rm MS}
\nonumber\\
&=&\!\!\!
\frac{g^2 C_F}{4 \pi^2} \frac{1}{3!}
\epsilon_{ABCD} \chi_1^A \chi_1^B
\bigg\{
\frac{[13] \eta_1^C \chi_3^D}{x_{13}^2 x_{24}^2}
+
\frac{[12] \eta_1^C \chi_2^D}{x_{13}^4}
+
\frac{[41] \eta_1^C \chi_4^D}{x_{24}^4}
-
\frac{2 [13]}{x_{13}^2 x_{24}^2} \frac{\chi_3^D \chi_4^D}{\vev{41}}
\bigg\}
\nonumber\\
&-&\!\!\!
\frac{g^2 C_F}{4 \pi^2} \frac{1}{4!}
\epsilon_{ABCD} \chi_1^A \chi_1^B
\bigg\{
2 \frac{[12]}{x_{13}^4} \eta_1^C \chi_2^D
+
2 \frac{[41]}{x_{24}^4} \eta_1^C \chi_4^D
+
\frac{[13]}{\vev{41}^2 x_{13}^2 x_{24}^2}
\bigg[
\chi_4^C \left( \vev{13} \chi_4^D + 2 \vev{34} \chi_1^D \right)
\nonumber\\ \label{W-W-1loop}
&+&\!\!\!
(\chi_4^C - \vev{41} \eta_1^C)
\left[ \vev{13} (\chi_4^D - \vev{41} \eta_1^D) + 2 \vev{34} \chi_1^D \right]
\bigg]
\bigg\}
+
{\rm (cyclic)}
\, ,
\end{eqnarray}
where $\eta_1$ was introduced in \re{super-mom}. It
can be re-expressed in terms of $\chi$'s using Eq.\ (\ref{EtaToChi}) as follows
\begin{align}
\eta^A_1
=
\frac{\chi_{4}^A}{\langle 41 \rangle}
+
\frac{\chi_{2}^A}{\langle 12 \rangle}
+
\frac{\langle 24 \rangle \chi_1^A}{\langle 41 \rangle \langle 12 \rangle}\,.
\end{align}
Applying the differential operator \re{Q-oper} to both sides of \re{W-W-1loop} we
verified that $(\epsilon \cdot Q) (\mathcal{W}_{4}^{\rm CH} -\mathcal{W}_{4}^{\rm MS})$
coincides with the difference of the expressions in the right-hand sides of Eqs.\ \re{CH-anom} and \re{MS-anom}.
\section{Supersymmetric Wilson loops at one loop}
\label{OneLoopWL}
In this section, we will compute one-loop correction to the supersymmetric Wilson loop $\mathcal{W}_4$.
As follows from its definition \re{W-prod} and \re{W-link}, the Wilson loop is given to this order by the
following expression
\begin{align}\label{W4-double}
\mathcal{W}_4
= 1+ \sum_{1\le j\le k \le 4}(ig)^2\int_0^1 dt_j \int_0^1 dt_k\,\frac1{N_c}\mathop{\rm tr}\nolimits \VEV{\mathcal{B}_j(t_j) \mathcal{B}_k(t_k)}
+O(g^4)\,,
\end{align}
where the superconnection $\mathcal{B}$ is defined in \re{MStoCH} and \re{B-CH}.
The explicit expression for the superconnection involves the sum over all quantum fields in $\mathcal{N}=4$ SYM.
As a consequence, the direct calculation of \re{W4-double} yields a large number of contributing terms
and makes the analysis very cumbersome. There is however another method that allows one to efficiently fix
the form of the one-loop Wilson loop with little effort. Namely, we will solve the supersymmetric Ward identities,
Eqs.~\re{CH-anom} and \re{MS-anom}, and reconstruct the explicit one-loop expression for the Wilson
loop $\mathcal{W}_4$ using some input requiring only a very small number of diagrams being computed
explicitly.
We recall that the
expansion of $\mathcal{W}_4$ in powers of the odd variables has the general form \re{W-exp}. To one-loop
order, this expansion terminates at %
\footnote{To see this we notice that the one-loop correction to $\mathcal{W}_4$
in \re{W4-double} is bilinear
in the superconnecton $\mathcal{B}(t)$ which in turn is a polynomial of degree 4 in the Grassmann variables.}
\begin{align}\label{W4-ansatz}
\mathcal{W}_4 = \mathcal{W}_{4;0} + \mathcal{W}_{4;1} + \mathcal{W}_{4;2} + O(g^4)\,,
\end{align}
where $\mathcal{W}_{4;0}$ is the bosonic light-like Wilson loop and $\mathcal{W}_{4;k=1,2}$
are given by homogenous polynomials of degree $4k$ in the $\chi$'s.
According to \re{CH-anom} and \re{MS-anom}, the anomaly $(\epsilon\cdot Q) \mathcal{W}_4$
is given to one-loop order by a homogenous polynomial in the odd variables
of degree 4. Then, replacing $\mathcal{W}_4$ by its general expression \re{W4-ansatz} and
matching the degree of the odd variables we find that $\mathcal{W}_{4;0}$
and $\mathcal{W}_{4;2}$ should be annihilated by $(\epsilon\cdot Q)$. For the
bosonic component $\mathcal{W}_{4;0}$ this is obvious, while for $\mathcal{W}_
{4;2}$ it leads to a nontrivial constraint $(\epsilon\cdot Q)\, \mathcal{W}_{4;2} =O
(g^4)$. As we will see in a moment, $\mathcal{W}_{4;2}$ takes zero value at one loop.
Thus, to this order, the supersymmetric Ward identities become anomalous for
the component $\mathcal{W}_{4;1}$ only. By definition, $\mathcal{W}_{4;1}$ is a homogenous
polynomial of degree 4 in $\chi_i$ (with $i=1,\ldots,4$). To one-loop
order, we shall use the following ansatz for $\mathcal{W}_{4;1}$
\footnote{Here, to simplify the notation, we stripped the
$SU(4)$ indices off the Grassmann variables and the accompanying Levi-Civita tensor these
are contracted with. As explained in Appendix \ref{NotationsConvensions}, in this form the $\chi$'s
can be treated as commuting variables.}
\begin{align}\label{W4-c}
\mathcal{W}_{4;1} = \frac{g^2 C_F}{4 \pi^2} \sum_{0\le k_i \le 4
\atop
k_1+k_2+k_3+k_4=4} c_{k_1k_2k_3k_4} (\chi_1)^{k_1} (\chi_2)^{k_2} (\chi_3)^{k_3} (\chi_4)^{k_4}
\, ,
\end{align}
where $c_{k_1k_2k_3k_4}$ are bosonic coefficient functions. As follows from Eq.\ \re{W-prod},
the Wilson loop $\mathcal{W}_4$ is invariant under the cyclic shift of indices, thus we have to require
that the right-hand side of \re{W4-c} should be cyclically symmetric as well.
The general solution to the Ward identities \re{CH-anom} and \re{MS-anom}
is defined up to an arbitrary function depending on the invariants
of the $Q-$supersymmetry transformations, $\chi_i^A \to \chi_i^A + \vev{i \epsilon^A}$.
Such invariants depend on three points and have the following form
\footnote{To construct these invariants we use the transformations
$\chi_i^A \to \chi_i^A + \vev{i \epsilon^A}$ and choose $\epsilon^A$ to put to
zero two of the $\chi$'s, e.g. $\chi_j=\chi_k=0$. Then, the remaining $\chi-$variables will be automatically
invariant under $Q-$supersymmetry.}
\begin{align}
\Theta_{ijk}^A = \chi_i^A \vev{jk} + \chi_j^A \vev{ki} + \chi_k^A \vev{ik}\,.
\end{align}
For $n$ points there are $(n-2)$ linear independent invariants. For $n=4$ we can choose
them to be $\Theta_{412}^A$ and $\Theta_{234}^A$.
Then, the solution to the Ward identities are defined modulo the substitution
\begin{align}\label{W4-amb}
\mathcal{W}_4 \to \mathcal{W}_4 + f_0(x) + f_1(x;\Theta_{412},\Theta_{234}) +
f_2(x) (\Theta_{412})^4 (\Theta_{234})^4\,,
\end{align}
where $f_0(x)$ and $f_2(x)$ are arbitrary functions of the bosonic variables
and $f_1(x;\Theta_{412},\Theta_{234})$ is an arbitrary homogenous polynomial in the odd
variables of degree 4.
\subsection{Boundary conditions}
To define a unique solution to the Ward identities, we have to fix the ambiguity
in \re{W4-amb}, or equivalently determine the functions $f_0$, $f_1$ and $f_2$.
The function $f_0(x)$ affects only the bosonic component and, therefore, its form
is fixed by the one-loop correction to the bosonic Wilson loop $W_4$.
To determine the remaining functions $f_1$ and $f_2$, we shall compute one-loop
corrections to $\mathcal{W}_4$ in the gauge $\chi_2^A=\chi_4^A=0$. In this gauge,
two major simplifications occur. Firstly, the expression for the superconnections
significantly reduce (see, e.g., Eqs.~\re{CHexponentE} and \re{Vj}), thus minimizing the
number of one-loop Feynman diagrams contributing to the loop. Secondly, the
$Q-$invariants now depend on a single odd variable, $\Theta_{412}=\vev{24}\chi_1$ and
$\Theta_{234}=-\vev{24}\chi_3$, and their contribution to \re{W4-amb} takes a particularly simple form.
Let us start with the last term in the right-hand side of \re{W4-amb}. In the gauge
$\chi_2=\chi_4=0$, it is proportional to $\eta_1^4\eta_3^4$. Using the expression for
one-loop corrections to the Wilson loop \re{W4-double}, we find that $O(\eta_1^4\eta_3^4)$
term could only appear from the correlation between $O(\eta_1^4)$
and $O(\eta_3^4)$ terms inside $\mathcal{B}_1(t_1)$ and $\mathcal{B}_3(t_3)$. The
corresponding correlation function involves
chiral components of the gauge field strength tensor located at two different segments of
the polygon, $\vev{F_{\alpha\beta}(x(t_1))F_{\alpha'\beta'}(x(t_3))}$. In $D=4$ dimension it vanishes,
while for $D=4-2\varepsilon$ it is proportional to $\varepsilon$. Therefore, for the result to be different
from zero, the integral over $t_1$ and $t_3$ should produce a pole $1/\varepsilon$. A simple calculation
shows that this does not happen and, therefore, the coefficient in front of $O(\eta_1^4\eta_3^4)$ term in
$\mathcal{W}_4$ equals zero. This immediately implies that the one-loop correction to $\mathcal{W}_4$
does not involve terms of degree 8 in odd variables,
\begin{equation}\label{zero}
\mathcal{W}_{4;2}=0+O(g^4)\,.
\end{equation}
We now turn to computing corrections to \re{W4-double} of the Grassmann degree 4.
They have the general form \re{W4-c}. For $\chi_2 = \chi_4 = 0$ we are left with
the terms of the form $\chi_1^{j_1}\chi_3^{4-j_1}$ with $j_1=0,\ldots,4$. As we
will show in the next subsection, to construct a unique solution to the Ward
identity it is sufficient to identify the contribution of two terms only,
$O(\chi_1^2\chi_3^2)$ and $O(\chi_1\chi_3^3)$. Using explicit expressions for
the superconnections, Eqs.~\re{Afield} and \re{Ffield}, it is easy to see that
such terms arise from the correlation between scalar and fermion fields, respectively,
entering into the expansion of $\mathcal{B}(t_1)$ and $\mathcal{B}(t_3)$.
Since the difference between the two superconnections, Eqs.~\re{B-diff} and \re{EOMinMSminusCH},
does not involve scalar operators, the contribution of $O(\chi_1^2\chi_3^2)$ terms to the two loops,
$\mathcal{W}_4^{\rm MS}$ and $\mathcal{W}_4^{\rm CH}$, is the same.
To compute it we retain in the right-hand side of \re{W4-double}
only one term with $j=1$, $k=3$ and replace
\begin{align}
\mathcal{B}(t) =- \ft{i}{4} \dot x^{\dot\alpha\alpha}
\theta_\alpha^A \theta_\beta^B D^\beta{}_{\dot\alpha} \bar\phi_{AB}
+
\ft{i}{2}
\dot\theta^{\alpha A} \theta_\alpha^B\,\bar{\phi}_{AB} +\ldots
\end{align}
A simple calculation shows that the coefficient in front of $\chi_1^2 \chi_3^2$ vanishes, or in
application to \re{W4-c} $c_{2020} = 0$. Then, the cyclic invariance of the Wilson loop \re{W4-c}
implies that
\begin{align}\label{bc1}
c_{2020} = c_{0202} = 0\,.
\end{align}
We remind that these relations hold for both Wilson loops.
Let us now examine $O(\chi_1\chi_3^3)$ contribution to \re{W4-double}. By virtue of \re{B-diff}, it takes
a different form for the two Wilson loops. For $\mathcal{W}_4^{\rm MS}$,
we replace the superconnection in \re{W4-double} by the following expressions (see Eqs.~\re{Afield} and \re{Ffield})
\begin{align}\notag
\mathcal{B}^{\rm MS}(t_1)
&=
\ft{i}{2} \langle 1 \theta^A_1 \rangle [1 | \bar\psi_A (x_1 - t_1 x_{12}) ]+\ldots
\, , \\[2mm]
\mathcal{B}^{\rm MS}(t_3)
&=
\ft{1}{12} \epsilon_{ABCD}
\left(
\langle 3 \theta^A_3 \rangle
[ 3 | \partial | \theta^B \rangle
+
4 \langle \theta_3^A \theta_{4}^B \rangle
\right)
\langle \theta^C | \psi^D (x_3 - t_3 x_{34}) \rangle+\ldots
\, ,
\end{align}
and find that the coefficient in front of $\chi_1\chi_3^3$ vanishes,
$c^{\rm MS}_{1030}=0$. Again, the cyclic invariance of \re{W4-c} leads to
\begin{align}\label{bc2}
c^{\rm MS}_{1030}=c^{\rm MS}_{0103}=c^{\rm MS}_{3010}=c^{\rm MS}_{0301}=0\,.
\end{align}
For the Wilson loop $\mathcal{W}_4^{\rm CH}$, we use the fact that the difference
between the superconnections \re{B-diff} is proportional to the fermion equations
to motion to obtain
\begin{align}\label{c-CH}
c^{\rm CH}_{1030}=
-\frac{1}{6}
\frac{\langle 24 \rangle^2}{\langle 23 \rangle^2 \langle 34 \rangle \langle 41 \rangle x_{13}^2}
=\frac{1}{6} \lr{\frac1{x_{13}^2}+\frac1{x_{24}^2}}\frac{\vev{24}}{\vev{23}\vev{34}\vev{13}}
\,,
\end{align}
where in the last relation we made use of the identities
\begin{align}
\frac{x_{13}^2}{x_{24}^2} = -\frac{\vev{12}\vev{34}}{\vev{23}\vev{41}}\,,\qquad
\frac{x_{13}^2+x_{24}^2}{x_{24}^2} =-\frac{\vev{13}\vev{24}}{\vev{23}\vev{41}}\,.
\end{align}
The same result \re{c-CH} can be deduced from \re{W-W-1loop} by examining the coefficient in front of
$\chi_1\chi_3^3$ in $\mathcal{W}_{4}^{\rm MS} -\mathcal{W}_{4}^{\rm CH}$.
\subsection{Solving the Ward identities}
\label{SolvingWI}
We are now in a position to construct the solution to the Ward identities \re{CH-anom} and \re{MS-anom}.
We start with the former and substitute the ansatz \re{W4-c} into the left-hand side of \re{CH-anom}.
In this way we obtain a relation both sides of which depend on the parameters of the supersymmetric
transformation $\vev{i\epsilon^A}$ (with $i=1,\ldots,4$). Notice that among the four parameters only
two are linearly independent
\begin{align}
\vev{2 \epsilon} = \vev{1 \epsilon}\frac{\vev{23}}{\vev{13}} + \vev{3 \epsilon}\frac{\vev{12}}{\vev{13}}
\, ,
\qquad
\vev{4 \epsilon} = - \vev{1 \epsilon}\frac{\vev{34}}{\vev{13}}- \vev{3 \epsilon}\frac{\vev{41}}{\vev{13}}\,.
\end{align}
Then, comparing the coefficients in front of various terms involving $\vev{1\epsilon}$, $\vev{3\epsilon}$
and different powers of $\chi_i$, we obtain a system of linear inhomogeneous equations for the coefficients
$c_{k_1k_2k_3k_4}$ entering \re{W4-c}. In addition, we impose the boundary conditions \re{bc1} and \re{bc2}.
The resulting system of equations is overdetermined but it has a unique solution leading to the
following expression for the one-loop Wilson loop
\begin{align}\notag
\mathcal{W}_{4;1}^{\rm MS}
&=
\frac{g^2 C_F}{4\pi^2} \left( \frac{1}{x_{13}^2} + \frac{1}{x_{24}^2} \right)
\bigg\{
- \frac{\vev{24}^2}{\vev{12}^2\vev{41}^2} \frac{\chi_1^4}{24}
-
\left(
\frac{\vev{24}}{\vev{12} \vev{41}^2} \chi_4
+
\frac{\vev{24}}{\vev{12}^2 \vev{41}} \chi_2
\right) \frac{\chi_1^3}{6}
\\\label{MSsolutionWI}
&
+
\bigg(
\frac{\chi_2 \chi_3}{\vev{12} \vev{13}}
-
\frac54 \frac{\chi_2 \chi_4}{\vev{41} \vev{13}}
-
\frac{\chi_3 \chi_4}{\vev{41} \vev{13}}
-
\frac{1}{4}
\frac{3 \vev{24} \vev{13} + 2 \vev{34} \vev{12}}{\vev{12}^2 \vev{13} \vev{24}} \chi_2^2
\bigg) \frac{\chi_1^2}{3}\bigg\}
+ {\rm (cyclic)}
\, .
\end{align}
In a similar fashion, solving the Ward identity \re{CH-anom} subject to
the boundary condition \re{bc1} and \re{c-CH} we find
\begin{align}
\label{CHsolutionWI}
\mathcal{W}_{4;1}^{\rm CH}
&=
\frac{g^2 C_F}{4 \pi^2} \left( \frac{1}{x_{13}^2} + \frac{1}{x_{24}^2} \right)
\bigg\{
\frac{\vev{23}\vev{24}}{\vev{12}^2\vev{13}\vev{41}}
\frac{\chi_1^4}{12}
\nonumber\\
&+
\bigg(
\frac{\vev{24}}{\vev{41}\vev{12}\vev{13}} \chi_3
-
\frac{\vev{13}\vev{24} + \vev{12}\vev{34}}{\vev{12}\vev{13}\vev{41}^2} \chi_4
+
\frac{\vev{23}\vev{41} - \vev{13}\vev{24}}{\vev{12}^2\vev{13}\vev{41}} \chi_2
\bigg)
\frac{\chi_1^3}{6}
\nonumber\\
&+
\bigg(
\frac{\chi_2\chi_3}{\vev{12}\vev{13}}
-
\frac{\chi_3\chi_4}{\vev{13}\vev{41}}
-
\frac{\chi_2\chi_4}{\vev{12}\vev{41}}
-
\frac{\chi_2^2}{\vev{12}^2}
\bigg) \frac{\chi_1^2}{2}
\bigg\}
+ {\rm (cyclic)}
\, .
\end{align}
We verified that these relations obey several consistency conditions. By construction, they satisfy
the Ward identities \re{CH-anom} and \re{MS-anom}, respectively, and,
therefore, they do not respect $Q-$supersymmetry
\begin{align}\label{Q-W}
Q^A_\alpha\, \mathcal{W}_{4;1}\neq 0\,.
\end{align}
Also, the difference between the
two Wilson loops is in agreement with \re{W-W-1loop}.
Next, we computed $O(\chi_1\chi_2\chi_3\chi_4)$
and $O(\chi_1^4)$ terms in the right-hand side of \re{W4-double} and checked that they are correctly
reproduced in \re{MSsolutionWI} and \re{CHsolutionWI}. In the first case, the terms involving all four
$\chi-$variables cannot be produced by the correlation function of two superconnections (see
\re{W4-double}) and, therefore, they should be absent in the one-loop approximation. In the second
case, to identify the terms $O(\chi_1^4)$ we can set $\chi_2=\chi_3=\chi_4=0$ thus reducing significantly
the number of contributing diagrams.
We recall that the relations \re{MSsolutionWI} and \re{CHsolutionWI} define (together with
\re{W4-ansatz}) the part of the one-loop correction to the Wilson loop \re{W4-ansatz} depending on
the Grassmann variables. The bosonic component of the Wilson loop, $\mathcal{W}_{4;0}$, has
been previously computed in Ref.~\cite{KorDruSok08}.
To one-loop order, it develops a double pole $1/\varepsilon^2$ due to the presence of
specific ultraviolet (UV) cusp singularities \cite{KK92}.
In distinction with $\mathcal{W}_{4;0}$, the expressions in the right-hand side of \re{CHsolutionWI}
and \re{MSsolutionWI} are free from any divergences and are well-defined for $\varepsilon\to 0$.
This implies that UV divergences cancel in the ratio of the supersymmetric Wilson loop and
its bosonic component,
\begin{align}\label{W/W}
\mathcal{W}_4/\mathcal{W}_{4;0}=1+\mathcal{W}_{4;1}
+O(g^4)\,,
\end{align}
and,
therefore, the scaling (dilatation) invariance is restored in their ratio, $D\,\lr{\mathcal{W}_4/\mathcal{W}_{4;0}}
=0$. But what about special conformal transformation?
The special conformal boosts are given by the superposition of translations
and inversions, $K^{\dot\alpha\alpha}=I P^{\dot\alpha\alpha} I$. Translations shift $x_i$ but they do not affect
$\lambda_i$ and $\chi_i$. At the same time, the inversion acts as
\begin{align}\label{Inv}
I[(x_i)_{\alpha\dot\alpha}] = \frac{x_i^{\dot\alpha\alpha}}{x_i^2}\,,\qquad
I[\lambda_i^\alpha] = (\mu_i)_{\dot\alpha}\,,\qquad I[\chi_i] =\chi_i\,,
\end{align}
with $\mu_i$ defined in \re{twistors}.
A close examination of \re{CHsolutionWI} and \re{MSsolutionWI} shows that both expressions depend
on the angle brackets of the form $\vev{i\,i+1}$ and $\vev{i\,i+2}$. Using \re{Inv} we find that the former
brackets are transformed covariantly under the inversion,
$I[\vev{i\, i+1}] = \vev{i|x_i x_{i+1}|i+1} = x_{i+1}^2 \vev{i\,i+1}$, but this is not the case for the brackets
of the type $\vev{i\,i+2}$. It is then straightforward to verify that both expressions are transformed nontrivially
under inversion and, therefore, do not respect the special conformal invariance
\begin{align}\label{K-W}
K^{\dot\alpha\alpha} \mathcal{W}_{4;1}\neq 0\,.
\end{align}
This conformal anomaly is explicitly calculated in Appendix \ref{conf-anom}, where we show complete consistency
with the solution of the supersymmetric Ward identity that we constructed in Sect. \ref{SUSYWI}.
Combining together the relations \re{Q-W} and \re{K-W} we conclude that the
ratio of the Wilson loops in \re{W/W} does possess neither $Q-$supersymmetry,
nor conformal (super)symmetry already at one loop. As was already explained in Introduction,
this fact is in contradiction with the relations \re{W-sym} and \re{KW=0} which follow in
their turn from the conjectured duality between supersymmetric Wilson loops and scattering
amplitudes in $\mathcal{N}=4$ SYM.
\section{Conclusions}
\label{Conclusions}
The MHV scattering amplitudes in planar $\mathcal{N}=4$ SYM are dual to bosonic Wilson loops. In
this paper, we explored two recent proposals for extending this duality to generic non-MHV amplitudes. The
corresponding dual object should have the same symmetries as the scattering amplitudes and be invariant
to all loops under the chiral half ($Q-$ and $\bar S-$symmetries) of the $\mathcal{N}=4$ superconformal
symmetry. The supersymmetric extensions of the bosonic Wilson loop proposed in Ref.~\cite{Car10}
comply with this condition at the classical level but only up to terms proportional to field equations. We would like to
point out that while our conclusions apply to the supersymmetric Wilson loop in Minkowski space, the twistor space
version from Ref.~\cite{MasSki10} appears to be formulated off-shell and thus the question of whether any pathologies
permeate this formalism as well requires extra studies.
Examining the properties of the supersymmetric Wilson loops at the quantum level,
one encounters the following complication. According to its definition, the perturbative expansion
of the Wilson loop involves products of various fields (scalars, gauginos and gluon)
integrated along a closed contour. Since the integration contour is uniquely determined by
the light-like momenta of the scattered particles, these fields inevitably become light-like
separated. In the classical theory such an object is well defined in $D=4$ dimensions, while at
the quantum level it suffers from light-cone singularities and requires regularization.
This subtlety is not a specific feature of the supersymmetric Wilson loop and it is already
present in the bosonic light-like Wilson loop. In the latter case, due to the light-cone
singularities of the two-point functions of the gauge fields, the quantum corrections to the bosonic
Wilson loop generate ultraviolet divergences which appear as double poles in $\varepsilon$ in the dimensional
regularization scheme with $D=4-2\varepsilon$. By the same
token, the calculation of loop corrections to the supersymmetric Wilson loop demands
the use of a regularization which we chose to be the supersymmetry preserving
dimensional reduction
\footnote{It is important to point out that there is a difference in the use of dimensional reduction
and dimensional regularization already at one loop. Had we used dimensional regularization
instead, the supersymmetry Ward identities would not be fulfilled indicating inconsistencies in the
treatment at the quantum level.}
We would like to emphasize that even though the final expressions for the one-loop correction to
$\mathcal{W}_{4;1}$ (see Eqs.~\re{CHsolutionWI} and \re{MSsolutionWI}) are finite as $\varepsilon\to 0$,
they arise by adding
together the contributions from several Feynman diagrams, each of which develops light-cone
divergences and, therefore, requires a regularization
Since the calculation of the quantum corrections to the supersymmetric Wilson loop involves
going away from the critical dimension $D=4$, one might suspect that some of the
classical symmetries will be broken at the quantum level. Indeed,
we demonstrated in this paper that, for the supersymmetric Wilson loops under consideration,
both the chiral supersymmetry and conformal invariance are already broken at one loop.
The underlying mechanism looks as follows. The variation of the Wilson loop under
supersymmetry $Q-$transformations is proportional to a contribution from the equation of
motion operators. In dimensional regularization with $D=4-2\varepsilon$, the latter scales as
$O(\varepsilon)$ and vanishes when the number of space-time dimensions is set to four.
However, the correlation function of the equations of motion with the light-like Wilson loop
produces a divergent $O(1/\varepsilon)$ contribution.\footnote{This effect is reminiscent to the
so-called $\mu_\varepsilon^2-$terms in the scattering amplitudes. The latter appear as $O(\varepsilon)$
terms in the expression for the {\it integrand} of the dimensionally regularized amplitudes, but they
produce a nontrivial contribution after integration over loop momenta starting from two loops \cite{BerDixKosRoiSprVerVol08,KosRoiVer10}.}
Combining together the two effects, we obtain a nontrivial finite $O(\varepsilon^0)$ contribution to
the corresponding supersymmetric Ward identity. The latter takes the form of a differential equation
in the Grassmann variables and completely fixes the form of the super-Wilson loops with a minimal
perturbative input. The unique solutions that we found were tested against explicit one-loop calculations of
the Wilson loops.
Our analysis explicitly demonstrates that the quantum anomalies break the chiral
supersymmetry of the Wilson loops, thus invalidating the conjectured duality with
the scattering amplitudes in $\mathcal{N}=4$ SYM. We recall that for the amplitudes
the same symmetry (the dual chiral $Q-$supersymmetry) remains exact at the quantum level. In
fact, it is a trivial consequence of the way the dual variables are introduced in Eq.~\re{super-mom}.
Notice that the general solution to the anomalous supersymmetriy Ward identity is defined
up to an arbitrary supersymmetric invariant function which could be compared with the
scattering amplitude. In application to the Wilson loop
this would mean that while the super Wilson loop is affected by nonvanishing contributions of the
equations of motion, it is its supersymmetric invariant part with the properly subtracted
anomaly which is matched with the scattering amplitudes. This would require, however, a detailed knowledge
of the all-loop anomaly as well as a consistent formulation of a scheme for separation between
anomalous and invariant terms inside the Wilson loop.
The fact that the anomaly in the $Q-$supersymmetry is due to the contributions of the field equations to the Wilson loop correlation function is closely related to the necessity to work with an on-shell realization of the $\mathcal{N}=4$ supersymmetry algebra. The problem we encounter can be presented as follows. The bosonic superconnection $\mathcal{A}_{\alpha\dot\alpha}(x,\theta)$ introduced in \re{Afield}, can be viewed as the supersymmetric extension of the gauge field $A_{\alpha\dot\alpha}(x)$ obtained by a finite chiral supersymmetry transformation with generators $Q^\alpha_A$ and with the Grassmann variables $\theta^A_\alpha$ in the role of the parameter:
\begin{align}\label{expo}
\mathcal{A}_{\alpha\dot\alpha}(x,\theta) = e^{(\theta\cdot Q)} A_{\alpha\dot\alpha}(x)\,,
\end{align}
with $(\theta\cdot Q)=\theta^A_\alpha Q^\alpha_A$.
In expanding the exponential we make use of the relation $\{ Q^\alpha_A, Q^\beta_B\}=0$ which follows from the chiral supersymmetry algebra. However, in the $\mathcal{N}=4$ case even this chiral subalgebra of the full Poincar\'e supersymmetry closes only on shell. Then it is clear that the bosonic connection constructed from (\ref{expo}) is invariant under $Q-$transformations (up to a compensating gauge transformation) only modulo field equations.
In supersymmetric theories the well-known remedy consists in adding sets of so-called auxiliary fileds, whose equations of motion are algebraic and whose role is to maintain the closure of the supersymmetry algebra off shell. If we had the relevant auxiliary fields at hand, our super Wilson loop would presumably not suffer from the anomaly we observed. An example, to be presented elsewhere, is provided by $\mathcal{N}=1$ SYM. In this simplest supersymmetric gauge theory one needs to add just a single scalar auxiliary field in order to close the algebra off shell. The $U(1)$ R symmetry of $\mathcal{N}=1$ SYM does not allow us to construct a purely chiral extension of the Wilson loop. Instead, we may consider the fully supersymmetric $\mathcal{N}=1$ super Wilson loop, with and without the auxiliary field in it. This will give us an alternative view on the anomaly mechanism. However, coming back to the $\mathcal{N}=4$ case, we have to recall the very old result of \cite{Siegel:1981dx} on the absence of auxiliary fields for $\mathcal{N}=4$ SYM. In view of this, the $\mathcal{N}=4$ super Wilson loop seems condemned to suffer from supersymmetry anomalies. The only escape might be that the no-go argument of \cite{Siegel:1981dx} applies to the full $\mathcal{N}=4$ supersymmetry, while here we need just its chiral half. So, this old issue needs to be revisited.
Finally, we would like to comment on another approach to the dual description of scattering amplitudes
proposed in Refs.\ \cite{EdeKorSok10}. The starting point of this proposal is the correlation
function of 1/2 BPS bilinear scalar operators $\mathcal{O}(x)=\mathop{\rm tr}\nolimits ( \phi^2) $. It was found that, in
the multiple light-cone limit $x_{i,i+1}^2\to 0$, the leading asymptotic behavior of the correlation
function $\vev{ \mathcal{O}(x_1)\ldots \mathcal{O}(x_n)}$ is given by a product of free
propagators multiplied by a light-like bosonic Wilson loop squared.
This result was used to demonstrate the duality between the correlation function of
bosonic operators in the light-cone limit $x_{i,i+1}^2\to 0$ and the MHV amplitudes at loop
level. Recently, the duality between the correlation functions and amplitudes was further extended
to non-MHV amplitudes in Refs.~\cite{EHKS}. In $\mathcal{N}=4$ SYM, the scalar operator
$\mathcal{O}(x)$ is the lowest component of the so-called stress-tensor supermultiplet described
by the 1/2 BPS short superfields $\mathcal{T}(x,\theta,\bar\theta)$. Then,
a natural supersymmetric generalization of the correlation function is
\begin{align}\label{TTT}
G_n = \vev{\mathcal{T}(x_1,\theta_1,\bar\theta_1)\ldots \mathcal{T}(x_n,\theta_n,\bar\theta_n)} \,.
\end{align}
This correlation function depends on both Grassmann variables, $\theta_i$ and $\bar\theta_i$,
and it enjoys the full $\mathcal{N}=4$ superconformal symmetry to all loops.
Then, the duality relation between the correlation function and the superamplitudes \re{gen} is
established by setting all $\bar\theta_i=0$ and in the limit where the points $(x_i,\theta_i)$
coincide with the vertices of the light-like polygon $\mathcal{C}_n$
\begin{align}\label{new}
\lim_{x_{i,i+1}^2\to 0} G_{n}(x_i,\theta_i,\bar\theta_i=0)
\sim \lr{\sum_{k=0}^{n-4} a^{k} \widehat{\mathcal{A}}_{n;k}(\lambda_i,\tilde\lambda_i,\eta_i;a)}^2\,,
\end{align}
where the two sets of variables $(x_i,\theta_i)$ and $(\lambda_i,\tilde\lambda_i,\eta_i)$
are related to each other through \re{super-mom} and the proportionality factor is given
by the product of $n$ consecutive scalar propagators $1/(x_{12}^2\ldots x_{n1}^2)$.
This relation has been checked in Ref.~\cite{EHKS}
for $n=4,5,6$ amplitudes at tree- and one-loop level, as well as for the NMHV tree-level amplitudes general $n$. Unlike the supersymmetric Wilson loop,
the $Q-$supersymmetry of the
correlation function in the left-hand side of \re{new} is not broken. The reason for
this is that the correlation function, viewed as a function of $x_{i,i+1}^2$, is a
less singular object as compared with the Wilson loop. For $x_{i,i+1}^2\neq 0$ the former is well-defined in $D=4$ dimensions, while the latter suffers from
UV divergences due to the presence of (non light-like) cusps on the integration contour and,
therefore, requires a regularization \cite{KK92}.
The situation here is very much reminiscent of the one with the operator
product expansion (OPE). Let us consider a product of two protected operators
in $\mathcal{N}=4$ SYM like $\mathcal{T}(x,\theta,\bar\theta)$. It is well-defined in $D=4$ and does not require any
regularization as long as the operators are not null separated.
However, expanding the product of operators into the sum of local (Wilson) operators we find that the latter develop anomalous
dimensions whose calculation requires introducing a UV regularization and whose explicit expressions depend
(starting from two loops) on the choice of the regularization scheme. We recall
however that each Wilson operator is accompanied by corresponding coefficient
function. It is this coefficient function that insures independence of the product
of operators on the renormalization scale as well as its scheme independence.
Coming back to the correlation function \re{TTT}, we expect that in the light-cone limit \re{new}, it
reduces, in complete analogy with conventional OPE, to the product of the supersymmetric Wilson
loop and a coefficient function. Along these lines, all anomalies of the supersymmetric Wilson loops
that we identified in this paper should be compensated by the coefficient function
in such a manner that their product is anomaly free.
\section*{Acknowledgments}
G.K. would like to thank Luca Griguolo, Henrik Johansson and Domenico Seminara for interesting discussions. G.K. and E.S. are grateful
to Simon Caron-Huot, Burkhard Eden, Paul Heslop and David Skinner for discussions. This work of A.B. was supported by the National
Science Foundation under Grant No. PHY-0757394
|
\section{Introduction}
We present a comparison between the oxygen abundances derived homogeneously for
ionized gas in \ion{H}{2} regions and planetary nebulae (PNe) of the solar
neighborhood. This approach is the only way in which we can compare the
abundances of the interstellar medium (ISM, represented in this context by
\ion{H}{2} regions) with stellar abundances (represented by PNe) using the same
procedure and the same atomic data. Furthermore, the abundance analysis is based
on optically thin lines, providing a large advantage over abundances derived
directly for stars. We choose oxygen for this
analysis because it is the element for which we can get the most reliable
abundances in ionized gas. Besides, the oxygen abundance is not expected to be
substantially modified by the evolutionary processes that take place in the
progenitors of PNe with near-solar metallicity (although the most massive
progenitors can achieve a small amount of destruction -- see, e.g.,
\citealp{kar10}).
In principle, our approach has one disadvantage. Weak recombination lines (RLs)
of heavy elements imply higher abundances than collisionally excited lines
(CELs), by factors around or above 2, both in \ion{H}{2} regions and PNe (e.g.,
\citealp{gar06,liu00}). To avoid this difficulty, we will consider here the
results implied by both RLs and CELs. However, different explanations of the
discrepancy imply that the best abundance estimates will be close to the ones
implied by either RLs or CELs, or will be intermediate between them (see
\citealp{rod10}, and references therein). Hence, the interpretation of the
results must take into account that the explanation of the discrepancy can be
different for \ion{H}{2} regions and PNe.
Our sample contains five \ion{H}{2} regions (M8, M16, M17, M20, and M42) and
seven PNe (NGC~3132, NGC~3242, NGC~6210, NGC~6543, NGC~6572, NGC~6720, and
NGC~6884) with deep spectra (\citealp{est04,gar06,gar07,liu04,tsa03,wes04}). All
the objects have individual distance determinations locating them at distances
below 2 kpc. This constraint should minimize the effects of the Galactic
abundance gradient. The PNe were selected from the sample of low-ionization
nebulae compiled by \citet{del09}, and are estimated to have
$\mbox{O}^{3+}/\mbox{O}<0.15$. Since no [\ion{O}{4}] lines are observed in the
optical along with all the other lines we will be using, this restriction
reduces the uncertainties introduced by the correction for the presence of
O$^{3+}$.
\section{Results}
All the objects were analyzed in a similar way and using the same atomic data.
An average electron density, $n_e$, was obtained using 2--3 diagnostic line
ratios. Electron temperatures, $T_e$, for the low- and high-ionization regions
inside each nebula were obtained from the usual [\ion{N}{2}] and [\ion{O}{3}]
diagnostic line ratios (see, e.g., \citealp{ost06}). The $\mbox{O}^{+}/\mbox{H}^+$ and
$\mbox{O}^{++}/\mbox{H}^+$ abundance ratios were calculated using the values of
$n_e$ and $T_e$ ($T_e$[\ion{N}{2}] for O$^+$; $T_e$[\ion{O}{3}] for O$^{++}$)
and the line intensity ratios $I([\ion{O}{2}]~\lambda3727)/I(\mbox{H}\beta)$ and
$I([\ion{O}{3}]~\lambda\lambda4959,5007)/I(\mbox{H}\beta)$. The total oxygen
abundance was derived by adding the O$^{+}$ and O$^{++}$ abundances, and using
ionization correction factors for the presence of O$^{3+}$ that go from
$\mbox{O}/(\mbox{O}^{+}+\mbox{O}^{++})=1$ (for
the \ion{H}{2} regions and some PNe) to 1.18. For the results implied by RLs,
the O$^{++}$ abundance was derived using the \ion{O}{2} RLs of multiplet 1.
Then, we estimated the total oxygen abundance implied by RLs by assuming the
same ionization fractions found with CELs.
Further details on the procedure will be provided in Rodr\'\i guez \&
Delgado-Inglada (2011, in preparation).
Figure~\ref{fig:uno} shows the oxygen abundances implied by CELs and RLs as a
function of $\mbox{O}^+/\mbox{O}^{++}$ for all the objects in the sample. The
abundances implied by CELs in \ion{H}{2} regions are similar, with
$12+\log(\mbox{O}/\mbox{H})=8.45\mbox{--}8.54$, suggesting that the local ISM is
very homogeneous. As for the PNe, if we exclude NGC~3242 (where the correction
for O$^{3+}$ is the largest, making its total oxygen abundance more uncertain),
the abundances implied by CELs are in the range
$12+\log(\mbox{O}/\mbox{H})=8.65\mbox{--}8.80$. Since the oxygen abundances of
PNe are expected to reflect the abundances in the ISM from which their
progenitor stars formed several gigayears ago, this result is opposite to what
we would predict using simple chemical evolution models. This overabundance of
oxygen in PNe also holds when we consider the abundances derived with RLs. In
fact, since the abundances implied by RLs in \ion{H}{2} regions are similar to
the abundances implied by CELs in PNe, even results intermediate between those
implied by CELs and RLs (and shifted by different amounts in both kinds of
objects) would indicate the presence of an overabundance in PNe.
\begin{figure}[!t]
\includegraphics[bb=60 270 553 683,clip,width=7.9cm]{fox1color}
\caption{Oxygen abundances implied by CELs (filled symbols) and RLs (open
symbols) as a function of $\mbox{O}^+/\mbox{O}^{++}$ for \ion{H}{2} regions
(squares) and PNe (circles) in the solar neighborhood. The error bars
show the observational uncertainties.}
\label{fig:uno}
\end{figure}
This overabundance of oxygen in PNe could be due to different causes, like
oxygen production in the stellar progenitors, large-scale gas flows in the
Galaxy, recent infall of low-metallicity gas,
or extensive stellar migration from the inner Galaxy.
A discussion of these possibilities and a comparison of the results with those
implied by stars of different ages and by those based on absorption lines in the
diffuse ISM will be presented elsewhere (Rodr\'\i guez \& Delgado-Inglada 2011,
in preparation).
|
\section{Vanilla Technicolor setup for Linear Colliders} \label{sec:EffL}
The Large Hadron Collider (LHC) is producing a wealth of experimental results which are already providing interesting constraints for time-honored extensions of the SM of high energy physics. It is therefore timely to explore, in case a dynamical origin of the electroweak symmetry breaking occurs, the benefits stemming from the construction of future LCs.
Based on recent progress
\cite{Sannino:2004qp,Dietrich:2006cm,Ryttov:2007sr,Ryttov:2007cx,Dietrich:2005jn,Sannino:2008ha} in the
understanding of Walking Technicolor (WT) dynamics \cite{Holdom:1981rm,Holdom:1984sk,Eichten:1979ah,Lane:1989ej} various phenomenologically viable models have been proposed. Primary examples are: i) the $SU(2)$ theory with two techniflavors in the adjoint representation, known as Minimal Walking Technicolor (MWT) \cite{Sannino:2004qp}; ii) the $SU(3)$ theory with
two flavors in the two-index symmetric representation which is called Next to Minimal
Walking Technicolor (NMWT) \cite{Sannino:2004qp} and iii) The $SU(2)$ theory with two techniflavors in the fundamental representation and 1 techniflavor in the adjoint representation known as (UMT) \cite{Ryttov:2008xe}. These gauge theories have remarkable
properties~\cite{Sannino:2004qp,Hong:2004td,Dietrich:2005jn,Dietrich:2006cm,Ryttov:2007sr,Ryttov:2007cx} and alleviate the tension with the LEP precision data when used for Technicolor~\cite{Sannino:2004qp,Dietrich:2005jn,Foadi:2007ue,Foadi:2007se,Foadi:2008xj}. The effects of the extensions of the Technicolor models to be able to account for the SM fermion masses are important and cannot be neglected as shown in \cite{Fukano:2010yv}.
Despite the different envisioned underlying gauge dynamics it is a fact that the SM structure alone requires the extensions to contain, at least, the following chiral symmetry breaking pattern (insisting on keeping the custodial symmetry of the SM):
\begin{eqnarray}
\label{basepattern}
SU(2)_{\rm L}\times SU(2)_{\rm R} \to SU(2)_{\rm V} \ .
\end{eqnarray}
We will call this common sector of any Technicolor extension of the SM, the {\it vanilla} sector. The reason for such a name is that the vanilla sector is common to old models of Technicolor featuring running dynamics and the ones featuring walking dynamics associated to a slow running of the Technicolor gauge coupling constant. It is worth mentioning that the {\it vanilla} sector is common not only to Technicolor extensions but to {\it any} known extension, even of extra-dimensional type, in which the Higgs sector can be viewed as composite. In fact, the effective Lagrangian we are about to introduce can be used for modeling several extensions with a common {vanilla} sector respecting the same constraints spelled out in \cite{Foadi:2007ue}. The natural candidate for a walking technicolor model featuring exactly this global symmetry is NMWT \cite{Sannino:2004qp}.
Based on the {\it vanilla symmetry} breaking pattern we describe the low energy spectrum in terms of the lightest spin-one vector and axial-vector iso-triplets $V^{\pm,0}, A^{\pm,0}$ as well as the lightest iso-singlet scalar resonance $H$. In QCD the equivalent states are the $\rho^{\pm,0}$, $a_1^{\pm}$ and the $f_0(600)$ \cite{Belyaev:2008yj}. The 3 technipions $\Pi^{\pm ,0}$ produced in the symmetry breaking become the longitudinal components of the $W$ and $Z$ bosons.
The composite spin-one and spin-0 states and their interaction with the SM fields are described via the following effective Lagrangian which we developed, first for minimal models of walking technicolor \cite{Foadi:2007ue,Appelquist:1999dq}:
\begin{eqnarray}
{\cal L}_{\rm boson}&=&-\frac{1}{2}{\rm Tr}\left[\widetilde{W}_{\mu\nu}\widetilde{W}^{\mu\nu}\right]
-\frac{1}{4}\widetilde{B}_{\mu\nu}\widetilde{B}^{\mu\nu}
-\frac{1}{2}{\rm Tr}\left[F_{{\rm L}\mu\nu} F_{\rm L}^{\mu\nu}+F_{{\rm R}\mu\nu} F_{\rm R}^{\mu\nu}\right] \nonumber \\
&+& m^2\ {\rm Tr}\left[C_{{\rm L}\mu}^2+C_{{\rm R}\mu}^2\right]
+\frac{1}{2}{\rm Tr}\left[D_\mu M D^\mu M^\dagger\right]
-\tilde{g^2}\ r_2\ {\rm Tr}\left[C_{{\rm L}\mu} M C_{\rm R}^\mu M^\dagger\right] \nonumber \\
&-&\frac{i\ \tilde{g}\ r_3}{4}{\rm Tr}\left[C_{{\rm L}\mu}\left(M D^\mu M^\dagger-D^\mu M M^\dagger\right)
+ C_{{\rm R}\mu}\left(M^\dagger D^\mu M-D^\mu M^\dagger M\right) \right] \nonumber \\
&+&\frac{\tilde{g}^2 s}{4} {\rm Tr}\left[C_{{\rm L}\mu}^2+C_{{\rm R}\mu}^2\right] {\rm Tr}\left[M M^\dagger\right]
+\frac{\mu^2}{2} {\rm Tr}\left[M M^\dagger\right]-\frac{\lambda}{4}{\rm Tr}\left[M M^\dagger\right]^2
\label{eq:boson}
\end{eqnarray}
where $\widetilde{W}_{\mu\nu}$ and $\widetilde{B}_{\mu\nu}$ are the ordinary electroweak field strength tensors, $F_{{\rm L/R}\mu\nu}$ are the field strength tensors associated to the vector meson fields $A_{\rm L/R\mu}$~\footnote{In Ref.~\cite{Foadi:2007ue}, where the chiral symmetry is SU(4), there is an additional term whose coefficient is labeled $r_1$. With an SU($N$)$\times$SU($N$) chiral symmetry this term is just identical to the $s$ term.}, and the $C_{{\rm L}\mu}$ and $C_{{\rm R}\mu}$ fields are
\begin{eqnarray}
C_{{\rm L}\mu}\equiv A_{{\rm L}\mu}^a T^a-\frac{g}{\tilde{g}}\widetilde{W_\mu^a} T^a\ , \quad
C_{{\rm R}\mu}\equiv A_{{\rm R}\mu}^a T^a-\frac{g^\prime}{\tilde{g}}\widetilde{B_\mu} T^3\ ,
\end{eqnarray}
where $T^a=\sigma^a/2$, and $\sigma^a$ are the Pauli matrices.The 2$\times$2 matrix $M$ is
\begin{eqnarray}
M=\frac{1}{\sqrt{2}}\left[v+H+2\ i\ \pi^a\ T^a\right]\ ,\quad\quad a=1,2,3
\end{eqnarray}
where $\pi^a$ are the Goldstone bosons produced in the chiral symmetry breaking, $v=\mu/\sqrt{\lambda}$ is the corresponding VEV, and $H$ is the composite Higgs. The covariant derivative is
\begin{eqnarray}
D_\mu M&=&\partial_\mu M -i\ g\ \widetilde{W}_\mu^a\ T^a M + i\ g^\prime \ M\ \widetilde{B}_\mu\ T^3\ .
\end{eqnarray}
When $M$ acquires its VEV, the Lagrangian of Eq.~(\ref{eq:boson}) contains mixing matrices for the spin-one fields. The mass eigenstates are the ordinary SM bosons, and two triplets of heavy mesons, of which the lighter (heavier) ones are denoted by $R_1^\pm$ ($R_2^\pm$) and $R_1^0$ ($R_2^0$). These heavy mesons are the only new particles, at low energy, relative to the SM.
Now we must couple the SM fermions.
The minimal form for the quark Lagrangian is
\begin{eqnarray}
{\cal L}_{\rm quark}&=&\bar{q}^i_L\ i \slashed{D} q_{iL} + \bar{q}^i_R\ i \slashed{D} q_{iR} \nonumber \\
&-&\left[\bar{q}^i_L\ (Y_u)_i^j\ M\ \frac{1+\tau^3}{2}\ q_{jR}
+\bar{q}^i_L\ (Y_d)_i^j\ M \ \frac{1-\tau^3}{2}\ q_{jR} + {\rm h.c.}\right] \ ,
\label{eq:quark}
\end{eqnarray}
where $i$ and $j$ are generation indices, and $i=1,2,3$, $q_{iL/R}$ are electroweak doublets.
The covariant derivatives are the ordinary
electroweak ones,
\begin{eqnarray}
\slashed{D}q_{iL}&=&\left(\slashed{\partial}-i\ g\ \slashed{\widetilde{W}}^a\ T^a
-i\ g^\prime \slashed{\widetilde{B}} Y_{\rm L}\right)q_{iL} \ ,\nonumber \\
\slashed{D}q_{iR}&=&\left(\slashed{\partial}-i\ g^\prime \slashed{\widetilde{B}} Y_{\rm R}\right)q_{iR} \ ,
\end{eqnarray}
where $Y_{\rm L}=1/6$ and $Y_{\rm R}={\rm diag}(2/3,-1/3)$. As usual, one can exploit the global symmetries of the kinetic terms
to reduce the number of physical parameters in the Yukawa matrices $Y_u$ and $Y_d$. Thus we can take
\begin{eqnarray}
Y_u={\rm diag}(y_u,y_c,y_t) \ , \quad Y_d= V\ {\rm diag}(y_d,y_s,y_b) \ ,
\end{eqnarray}
and
\begin{eqnarray}
q^i_L=\left(\begin{array}{c} u_{iL} \\ V_i^j d_{jL} \end{array}\right) \ , \quad
q^i_R=\left(\begin{array}{c} u_{iR} \\ d_{iR} \end{array}\right) \ ,
\end{eqnarray}
where $V$ is the CKM matrix.
One can also add mixing terms of the fermions with the $C_{\rm L}$ and $C_{\rm R}$ fields \cite{Foadi:2007ue}. We will however neglect them in our analysis, since they affect the tree-level anomalous couplings highly constrained by experiments.
\subsection{Parameter Space of Vanilla Technicolor}
Some of the parameters of the tree-level Lagrangian can be related to the electroweak $S$-parameter and to the masses of the composite particles as shown in \cite{Foadi:2007ue,Belyaev:2008yj}.
{}To be concrete we assume here $S\simeq 0.3$ corresponding approximately to its naive value in the
NMWT model. The remaining parameters are the tree-level mass of the axial spin-one state $M_A$, the technicolor interaction strength $\tilde{g}$, the coupling $s$, and the composite Higgs mass $M_H$. The two parameters $s$ and $M_H$ mostly impact processes involving the composite Higgs.
\begin{figure}
\includegraphics[height=8cm,width=10cm]{figures/boundss0.eps}
\caption{Bounds, for $S=0.3$, in the $(M_A,\tilde{g})$ plane from: (i) CDF
direct searches of $R_1^0$ at Tevatron, in $p\bar{p}\rightarrow e^+e^-$, for
$s=0$ and $M_{H}=200$~GeV.
The forbidden regions is the uniformly shaded one in the left corner.
(ii) Measurement of the electroweak parameters W and Y at
95\% confidence level. The forbidden region is the striped one in
the left corner. (iii) The excluded region in the right-lower corner comes from imposing the modified Weinberg's sum rules. (iv) Consistency of the theory, i.e. reality for the vector and axial decay constants leads to the horizontal stripe in the upper part
of the figure. We stress again that
the shaded regions are excluded.}
\label{fig:bounds}
\end{figure}
Bounds on these parameters come both, from the electroweak precision tests, and direct searches. We shall give here a brief review of the constraints discussed in detail in \cite{Belyaev:2008yj} and not yet improved by the recent LHC experiments.
CDF imposes~\cite{Aaltonen:2008vx} lower bounds on $M_A$ and $\tilde{g}$ from direct searches of
$R_1^0$ in the $p\bar{p}\rightarrow e^+e^-$ process, as shown by {the uniformly shaded region}
in the lower left of Fig~\ref{fig:bounds}.
The measurements of the electroweak parameters $W$ and $Y$ exclude~\cite{Foadi:2007se} the striped region
on the lower left corner shown in Fig~\ref{fig:bounds} at the 95\% confidence level. The upper bound for $\tilde{g}$, shown by the upper horizontal line in Fig~\ref{fig:bounds}, is dictated by the internal consistency of the model. The upper bound for $M_A$ corresponds to the value for
which both Weinberg's sum rules are satisfied in the running and walking regime of Technicolor \cite{Appelquist:1998xf,Foadi:2007ue}. This bound is shown in the lower right corner of Fig~\ref{fig:bounds}.
Recently, additional bounds from unitarity have been studied in \cite{Foadi:2008xj}, and the interesting possibility of using flavor data to constrain directly the technicolor sector has also been considered in \cite{Fukano:2009zm}. We have checked that these constraint do not further reduce the parameter space in our case.
\subsection{Key features of Vanilla Technicolor Phenomenology at LCs}
When turning off the electroweak interactions interesting decay modes of the spin-one massive vector states and composite scalar are: \begin{eqnarray}
V \to \Pi\ \Pi \ , \quad A \to H\ \Pi \ , \quad H \to \Pi\ \Pi \ ,
\label{basedecays}
\end{eqnarray}
with the appropriate charge assignments. We assumed here that the composite Higgs is lighter than the vector states. This is the case for the scalar field in QCD \cite{Sannino:1995ik,Harada:1995dc,Harada:1996wr,Harada:2003em}.
Once the electroweak interactions are turned on the technipions become the longitudinal components of the $W$ and $Z$ bosons. Therefore the processes in \eqref{basedecays} allow to detect the spin-one resonances at LCs. Here we are making the assumption that $2M_Z < M_H < M_{A,V}$. For the neutral vectors:
\begin{eqnarray}
V \to W W \ , \quad vs \quad A \to H\ Z \ .
\end{eqnarray}
This picture is not quite complete since the massive spin-one states mix with the SM gauge bosons. After diagonalizing the spin-one mixing matrices (see \cite{Foadi:2007ue,Belyaev:2008yj}) the lightest and heaviest of the composite spin-one triplets are termed $R_{1}^{\pm, 0}$ and $R_{2}^{\pm, 0}$ respectively. In the region of parameter space where $R_{1}$ is mostly an axial-like vector (for a mass less than or about one TeV) and $R_2$ mostly a vector state has the following qualitative dependence of the couplings to the SM fields as function of the electroweak gauge coupling $g$ and $\tilde{g}$:
\begin{eqnarray} \label{eq:couplbeh}
g_{R_{1,2} f\bar{f}}\sim \frac{g}{\tilde{g}} \ , \quad g_{R_2 W W}\sim \tilde{g} \ , \quad g_{R_1 H Z}\sim \tilde{g} \ .
\end{eqnarray}
Notice that, since the heavy spin-one states do not couple directly to SM fermions, the couplings $g_{R_{1,2} f\bar{f}}$ arises solely from the mixing with $W$ and $Z$. This coupling is roughly proportional to $g/\tilde{g}$.
Using as guidance the picture above together with the mass spectrum inequality $M_H < M_{R_1} < M_{R_2}$ we set up the following collider search strategy for the heavy vectors and the composite Higgs.
For small $\tilde{g}$ (meaning $g \lesssim \tilde{g} \lesssim 2$) the coupling of $R_{1,2}$ to fermions is large and therefore the heavy spin-one states are produced directly via the elementary process $e^+e^- \to R_{1,2}$.
In this regime the couplings $R_{1,2}$ to fermions are roughly equal and therefore it is easy to identify these states as peaks in the di-lepton final state distributions. In the large $\tilde g$ regime ($ \tilde{g} > 2$) from \eqref{eq:couplbeh} it is clear that it is better to consider the direct production of the heavy states followed by decays to SM gauge bosons as well as one SM gauge boson and a composite Higgs, i.e.
$e^+e^- \to R_1 \to HZ $ and $e^+e^- \to R_2 \to W^+W^-$.
\section{Phenomenology}\label{sec:pheno}
To perform the signal and background analysis
we use the CalcHEP~\cite{Pukhov:2004ca} implementation\footnote{The FeynRules implementation can be downloaded here https://feynrules.phys.ucl.ac.be/wiki/TechniColor and here http://cp3-origins.dk/research/tc-tools.} of the above model described in \cite{Belyaev:2008yj}. The LanHEP
package~\cite{Semenov:2008jy} has been used to derive the Feynman rules for
the model. CalcHEP \cite{calchep_man} implements the Jadach, Skrzypek and Ward expressions of Ref.~\cite{ISR} for Initial State Radiation (ISR) and we use the
the parameterisation of Beamstrahlung specified for the
International LC
(ILC) project in \cite{ILC_RDR}:
\begin{eqnarray}
\mbox{Horizontal beam size (nm)} &=& 640, \nonumber \\
\mbox{Vertical beam size (nm)} &=& 5.7, \nonumber \\
\mbox{Bunch length (mm)} &=& 0.300, \nonumber \\
\mbox{Number of particles in the bunch (N)} &=& 2\times10^{10}.
\end{eqnarray}
These parameters were recently employed in a study of dimuon \cite{Basso:2009hf} and Higgs \cite{Basso:2010si} production at LCs in $Z'$ models.
We shall consider two center-of-mass energies, $\sqrt{s} = 1$ and 3~TeV, corresponding to the maximal energies of ILC and CLIC, respectively.
Within the allowed parameter space of vanilla technicolor it is possible to identify two limiting regions \cite{Foadi:2007ue,Foadi:2008xj} which we term here the {\it low-mass} and {\it high-mass} regions of the spin-one vector mesons. These are the regions investigated here.
In the low-mass region, the axial resonance is lighter than the vector one, and both resonance masses are below one TeV. On the other hand, in the heavy mass region the vector resonance is lighter than the axial, and the masses are above $1.5$ to $2$~TeV.
A LC with $\sqrt{s}=1$~TeV, as we shall see, provides a detailed study of vanilla technicolor in the low-mass region, whereas the high-mass region can be accessed with $\sqrt{s}=3$~TeV.
To be concrete we choose as reference values of the tree-level mass scale $M_A$ to be around $750$~GeV and $2250$~GeV corresponding to the low mass and high mass cases respectively. The physical masses of the heavy spin-one states are reported in Table \ref{table:masses}.
The mass difference of the physical eigenstates $R_1$ (for low masses) and $R_2$ (for high masses) with respect to $M_A$ is due mostly to the mixing with the SM gauge bosons.
\begin{table}
\begin{eqnarray}
\begin{array}{c|c||cc|cc}
M_A(\textrm{GeV}) & \quad \ \tilde g \quad \mbox{} &\quad M_{R_1^0}(\textrm{GeV}) &\quad \Gamma_{R_1^0}(\textrm{GeV}) &\quad M_{R_2^0}(\textrm{GeV}) &\quad \Gamma_{R_2^0}(\textrm{GeV}) \\
\hline\hline
750 & \quad \ 2 \quad \mbox{} & \quad 773 &\quad 1.15 &\quad 844 &\quad 2.53 \\
& \quad \ 3 \quad \mbox{} & \quad 760 &\quad 0.813 &\quad 894 &\quad 6.32 \\
& \quad \ 5 \quad \mbox{} & \quad 753 &\quad 1.81 &\quad 1080 &\quad 82.0 \\
& \quad \ 8 \quad \mbox{} & \quad 750 &\quad 16.2 &\quad 1440 &\quad 265 \\
\hline
2250 & \quad \ 2 \quad \mbox{} & \quad 2270 &\quad 29.6 &\quad 2360 &\quad 36.1\\
& \quad \ 3 \quad \mbox{} & \quad 2220 &\quad 60.4 &\quad 2290 &\quad 64.0 \\
& \quad \ 5 \quad \mbox{} & \quad 2090 &\quad 127 &\quad 2260 &\quad 170 \\
& \quad \ 8 \quad \mbox{} & \quad 1770 &\quad 148 &\quad 2250 &\quad 425
\end{array} \nonumber
\end{eqnarray}
\caption{Physical masses and decay widths of the heavy vector mesons.}
\label{table:masses}
\end{table}
We also assume the composite Higgs mass to be around $M_H=200$~GeV unless stated otherwise. Note that the Higgs mass affects the $e^+e^- \to \ell^+\ell^-$ and $e^+e^- \to W^+W^-$ processes via the widths of the heavy technivector states.
\begin{figure}[tbh]
\includegraphics[width=0.45\textwidth]{figures/110216_sigmavsgt_r1_ss1.eps}%
\hskip .5cm
\includegraphics[width=0.45\textwidth]{figures/110216_sigmavsgt_r2_ss1.eps}
\vspace{10mm}
\includegraphics[width=0.45\textwidth]{figures/110216_sigmavsgt_r1_ss3.eps}%
\hskip .5cm
\includegraphics[width=0.45\textwidth]{figures/110216_sigmavsgt_r2_ss3.eps}%
\caption{\label{fig:gtdep}Top row: The signal cross-sections for key processes at $\sqrt{s} = 1$~TeV as a function of $\tilde g$ with keeping the resonance masses $M_{R_1}$ (left) and $M_{R_2}$ (right) fixed at $0.75$~TeV. Bottom row: The same for $\sqrt{s} = 3$~TeV, with masses fixed at $2.25$~TeV. }
\end{figure}
We start by plotting in Fig.~\ref{fig:gtdep} the cross sections associated to the signals for the selected processes as a function of $\tilde g$. For the dimuon, $W^+W^-$ and $HZ$ final states we included only the diagrams with the heavy spin-one resonance $R_1^0$ ($R_2^0$) as intermediate states. The plots on the left (right) correspond to $R_1^0$ ($R_2^0$). We use two reference values for the center-of-mass energy, $\sqrt{s} = 1$ (top row) and $3$~TeV (bottom row). Notice that we keep the physically relevant resonance mass ($M_{R_1}$ on the left and $M_{R_2}$ on the right) fixed.
Albeit the detailed cross section differ for different processes it is clear from Fig.~\ref{fig:gtdep} that the dimuon and diboson cross sections are of the same order of magnitude for sufficiently low $\tilde g$. This is the region where the axial vector and vector spin-one states are heavily mixed. The dimuon process $e^+e^- \to R_{1,2}^0 \to \mu^+\mu^-$ (black thick lines) has the largest cross section for small $\tilde g$ and for $\sqrt{s} = 1$~TeV. For $\sqrt{s}=3$~TeV the cross sections for the diboson channels $e^+e^- \to R_{1,2}^0 \to W^+W^-$ (blue lines) and $e^+e^- \to R_{1,2}^0 \to HZ$ (red lines) are enhanced, but the dimuon channel is still expected to produce the best signal for low $\tilde g$ since it has a cleaner final state. For large $\tilde g$ the $HZ$ final state (for axial vector spin-one resonances) and the $W^+W^-$ final state (for vector spin-one resonances) are dominant as expected.
Fig.~\ref{fig:gtdep} also includes the cross section for the vector boson fusion $e^+e^- \to R_{1,2}^0 +2 \nu$ (thin magenta lines). In principle this process would be interesting at large $\tilde{g}$ because the heavy vector boson is produced via the fusion of two $W$s and therefore the production vertex depends only on $\tilde{g}$. However, due to the dynamical constraints for running and walking technicolor imposed in \cite{Foadi:2007ue} this process turns out to be suppressed with respect to the diboson processes. We will, therefore, analyze in detail the following signatures:
\begin{itemize}
\item[(1)]
$e^+ e^-\to R^{0}_{1,2}\to \ell^+\ell^-$
\item[(2)]
$e^+ e^-\to R^{0}_{1,2}\to W^\pm W^\mp\to
\ell + \nu + 2j$
\item[(3)]
$e^+ e^-\to R^{0}_{1,2}\to ZH\to
\ 2\ell 4 j$
,
\end{itemize}
where $\ell$ denotes a charged lepton (electron or muon) and $j$ denotes a jet. We choose as reference values $\tilde g=2,\ 3, \ 5$, and $8$. As pointed out above, and as we shall see in the detailed analysis below, signature (1) is expected to yield the best signal for $\tilde g=2, 3$, while (2) and (3) are relevant for investigating the values $\tilde g=5,8$.
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/eetomm_MA750.eps}%
\includegraphics[width=0.45\textwidth]{figures/eetomm_MA2250.eps}
\includegraphics[width=0.45\textwidth]{figures/eetomm_ss1_MA750.eps}%
\includegraphics[width=0.45\textwidth]{figures/eetomm_ss03.eps}%
\caption{\label{fig:dimuon}Top
row: Total cross-section for $e^+ e^-\to \mu^+\mu^-$ as a function of $\sqrt{s}$ with $M_A=0.75$~TeV (top left) and $M_A=2.25$~TeV (top right). Bottom row: The differential cross section for $e^+ e^-\to \mu^+\mu^-$ as a function of the dimuon invariant mass. The thicknesses and colors of the solid lines indicate values of $\tilde g$. The used values are $\tilde g =2,3,5$, and 8 form the thick black lines to thin red ones. The dashed black line is the SM prediction. }
\end{figure}
\subsection{Dimuon and $WW$ final states}
The $e^+e^-\to \mu^+\mu^-$ cross section is shown in Fig.~\ref{fig:dimuon} (top) as a function of $\sqrt{s}$ at fixed $M_A$. The SM cross section is shown as dashed black line.
Peaks in the cross section are observed when the center-of-mass energy hits a resonance mass, tabulated in Table~\ref{table:masses}. As explained earlier, the peaks get suppressed with increasing $\tilde g$.
We also plot the dimuon differential cross section as function of the invariant dimuon mass in Fig.~\ref{fig:dimuon} (bottom). It is the initial state radiation which allows for a dependence on the dimuon mass. The composite states $R_1^0$ and $R_2^0$ are clearly identified for $\tilde g=2,3$, whereas for $\tilde g=5,8$ the peaks are hardly visible. The trend mimics the one observed when investigating the Drell-Yan production at the LHC \cite{Foadi:2007ue}.
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/eetoWW_MA750.eps}%
\includegraphics[width=0.45\textwidth]{figures/eetoWW_MA2250.eps}
\includegraphics[width=0.45\textwidth]{figures/eetoww_mdist_ss1.eps}%
\includegraphics[width=0.45\textwidth]{figures/eetoww_mdist_ss3.eps}%
\caption{\label{fig:wwevents}Top row: Total cross-section for $e^+ e^-\to W^+W^-$ as a function of $\sqrt{s}$ with $M_A=0.75$~TeV (top left) and $M_A=2.25$~TeV (top right). Bottom row: Differential cross section for $e^+ e^-\to W^+ W^-$ as a function of the $M_{W^+ W^-}$ invariant mass.}
\end{figure}
The cross section for the $e^+e^-\to W^+W^-$ process as a function of $\sqrt{s}$ at fixed $M_A$ is shown in the top figures of Fig.~\ref{fig:wwevents}, while the differential cross section as function of the invariant mass is shown in the bottom plots of the same figure. The cross section for the signal here is larger than the dimuon one for $\tilde g =5, 8$. This requires also $M_A$ to be large.
\subsection{Reach in dilepton and WW production}
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/mm_le_R1_lumi.eps}%
\includegraphics[width=0.45\textwidth]{figures/WW_le_R1_lumi.eps}
\includegraphics[width=0.45\textwidth]{figures/mm_le_R2_lumi.eps}%
\includegraphics[width=0.45\textwidth]{figures/WW_le_R2_lumi.eps}%
\caption{\label{lumiss1} Estimates for the required luminosity for $3\sigma$ (dashed lines) and $5\sigma$ (solid lines) discoveries of the vanilla technicolor vector resonances in dilepton and $WW$ production with $\sqrt{s} = 1$~TeV. The various lines are for $\tilde g =2,3,$ and 5 form thick black lines to thin magenta lines.}
\end{figure}
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/mm_he_R1_lumi.eps}%
\includegraphics[width=0.45\textwidth]{figures/WW_he_R1_lumi.eps}
\includegraphics[width=0.45\textwidth]{figures/mm_he_R2_lumi.eps}%
\includegraphics[width=0.45\textwidth]{figures/WW_he_R2_lumi.eps}%
\caption{\label{lumiss3} The same as Fig.~\ref{lumiss1} but for higher $\sqrt{s} = 3$~TeV. The various lines are for $\tilde g =2,3,5$, and 8 form thick black lines to thin red lines.}
\end{figure}
Let us now estimate the luminosity needed to discover heavy spin-one states at LCs. We start with signature (1), i.e. $e^+e^- \to \ell^+\ell^-$. We use the following cuts:
\begin{eqnarray} \label{eq:acccuts}
p_\perp^\ell &>& 10~{\rm GeV},\qquad |\cos\theta^\ell| < 0.95
\end{eqnarray}
where $\theta^\ell$ is the angle with respect to the beam direction. The precision of the energy reconstruction is simulated by adding a Gaussian random noise with variance $(0.15 GeV)^2 E/GeV$ to the energy of each lepton.
In addition we perform the following cut for the mass distribution near the resonance peak \cite{Basso:2009hf}
\begin{eqnarray} \label{eq:peakcut}
|M_{\ell\ell}-M_{R}| < \max\left(0.5\Gamma_R,0.15 \sqrt{\frac{M_R}{\rm GeV}} {\rm GeV}\right)
\end{eqnarray}
where $R$ is either of the vector resonances. The two different cuts inside the $\max$ function come respectively from the width of the resonance and the lepton energy resolution. $M_R$ and $\Gamma_R$ are calculated from the effective theory and shown in Table~\ref{table:masses}.
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/mm_v_R1_lumi.eps}%
\includegraphics[width=0.45\textwidth]{figures/WW_v_R1_lumi.eps}
\includegraphics[width=0.45\textwidth]{figures/mm_v_R2_lumi.eps}%
\includegraphics[width=0.45\textwidth]{figures/WW_v_R2_lumi.eps}
\caption{\label{lumissvar} Estimates for the required luminosity for $3\sigma$ (dashed lines) and $5\sigma$ (solid lines) discoveries of the vanilla technicolor vector resonances with $\sqrt{s} = M_R+30$~GeV and with different values of the input parameters.
}
\end{figure}
To a good approximation the background can be taken to be the SM prediction for this process.
The signal is then defined as the excess of the vanilla technicolor over the SM background.
The significance of the signal is then defined as the number of
signal events divided by the square root of the number of
background events when the number of events is large, while a
Poisson distribution is used when the number of events is small.
We have also analyzed the signature (2),
$W^+W^- \to \ell + \nu + 2 j$, by studying the distribution of the transverse mass variable
\begin{eqnarray}
(M^{T}_{jj\ell})^2 &=&
\left[\sqrt{M^2(jj\ell)+p_T^2(jj\ell)}
+|\, {\slash\!\!\!\!\!\:p}_T|\right]^2-|\vec{p}_T^{}(jj\ell)+{\slash\!\!\!\!\!\:\vec{p}}_T|^2 \ .
\end{eqnarray}
The jet energy reconstruction is simulated by smearing the energy $E$ of each jet with a Gaussian random error having variance $(0.5 GeV)^2 E/GeV$. The lepton energies are smeared as explained above for the dilepton final state. As the background, we use again the SM prediction for the same process. We cut the signal and background as follows. We require that both leptons and jets have $|\cos\theta|<0.95$ with respect to the beam axis, the lepton has $|p_\perp|> 10$~GeV and that the jets have $|p_\perp|>20$~GeV. We add the cut
\begin{equation}
|M^T_{jj\ell}-M_{R}| < \max\left(0.5\Gamma_R,0.5 \sqrt{\frac{M_R}{\rm GeV}} {\rm GeV}\right) \ ,
\end{equation}
where the latter expression inside $\max$ estimates the jet energy resolution which is the dominating error source.
Notice that, in particular for high $M_R$ the $W$ decaying to two jets will be highly boosted, and hence the jets easily merge. However, we did not require a jet separation in our analysis, so we effectively count the single jet events in the signal as well. This improves the signal since there is no additional single jet background. The significance is calculated as for the $\ell^+\ell^-$ signature above.
The calculated significances for $e^+e^- \to \ell^+\ell^-$ and for $e^-e^+ \to W^+W^-$ are used to estimate the necessary luminosities to discover the heavy vector states for various parameter values in Figs.~\ref{lumiss1},~\ref{lumiss3}, and~\ref{lumissvar}.
First, let us take the center-of-mass energy to be fixed to the maximum operating energy of a given LC.
Figs.~\ref{lumiss1} and \ref{lumiss3} show the estimates for $3\sigma$ and $5\sigma$ discoveries with $\sqrt{s} = 1$~TeV and 3~TeV, respectively. We used the reference values $\tilde g=2,3,5,$ and 8 and scanned over $M_A$ from 0.5~TeV to 3~TeV, covering roughly the allowed parameter
space of Fig.~\ref{fig:bounds}.
In particular, for $\tilde g=3,5$ the two resonance peaks tend to overlap for some values of $M_A$. In such cases we only show the signal for the dominant resonance peak. This is why some of the lines end abruptly in the plots. For example, the $R_2$ resonance dominates the $W^+W^-$ signal for $\sqrt{s}=1$~TeV (right hande side plots in Fig.~\ref{lumiss1}), $\tilde{g}=2$ and when the resonance masses are near 1~TeV. Therefore our simple estimate for the $R_1$ reach fails, and is not shown. A more sophisticated analysis would be necessary to estimate the size of the weaker signal in the excluded regions.
As expected from the distribution plots of the previous section, the required luminosity for the $\ell^+\ell^-$ signature (left hand plots) increases strongly with $\tilde g$.
However, the $\tilde g$ dependence of the $W^+W^-$ estimates (right hand plots) is weaker, and therefore the required luminosities are lower than for the dilepton final state for high values of $\tilde g$.
It is also interesting to study the capability of the collider to find resonances in the scanning mode (varying $\sqrt{s}$) and its ability to confirm weak signals observed at the LHC. The reach for this kind of situation is presented in
Fig.~\ref{lumissvar}, where the center-of-mass energy is tuned to $M_{R_i}+30$~GeV. The value of the shift $30$~GeV between the resonance mass and $\sqrt{s}$ was chosen to be slightly larger than a typical vector width in our model. Notice that for $\tilde g = 2$ one has that $R_2^0$ appears to be detectable at very low luminosities.
For high $\tilde g$ and masses above 2~TeV the resonances are very broad, and our approach is expected to underestimate the signal.
Notice the peculiar structure in the $W^+W^-$ plots (visible on right hand side in Figs.~\ref{lumiss3} and~\ref{lumissvar}): we have estimates for $R_1^0$ for large $M_{R_1}$, which end suddenly at $M_{R_1}=1.6$~TeV and continue almost smoothly on the $R_2^0$ plot below for $M_{R_2}<1.6$~TeV. This is due to the level crossing between the axial and vector spin-one states in our model, which occurs at $1.6$~TeV. As pointed out above, the $W^+W^-$ signal \cite{Foadi:2007ue} is dominated by the vector spin-one resonance, while the axial one mostly decays to $HZ$. The axial vector signal can be easily separated from the vector one only in the regime of very small $M_{R_1}$.
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/HZ_MA750.eps}%
\includegraphics[width=0.45\textwidth]{figures/HZ_MA2250.eps}
\includegraphics[width=0.45\textwidth]{figures/HZ_mdist_ss1.eps}%
\includegraphics[width=0.45\textwidth]{figures/HZ_mdist_ss3.eps}%
\caption{\label{eehzprod} Top row:
The cross section as a function of $\sqrt{s}$ with $M_A=0.75$~TeV (top left) and $M_A=2.25$~TeV (top right). Bottom row: The differential cross section for $e^+ e^-\to ZH$ as a function of $M_{ZH}$. The various lines are for $\tilde g =2,3,5$, and 8 form thick black lines to thin red lines. }
\end{figure}
\subsection{6 particle final states from $HZ$ production}
The presence of the heavy vectors can lead to an enhancement of the composite Higgs production in association with a SM vector bosons, as pointed out in \cite{Zerwekh:2005wh,Belyaev:2008yj}. Since, in models of minimal walking technicolor, the composite Higgs is expected to be heavy with respect to $M_Z$ but light \cite{Hong:2004td,Dietrich:2005jn,Dietrich:2006cm,Sannino:2009za,Sannino:2008ha} with respect to the scaled up sigma in QCD \cite{Sannino:1995ik,Harada:1995dc,Harada:1996wr,Black:1998wt,Harada:2003em}, the 6 particle final states from the associate Higgs production is therefore an appealing discovery channel \cite{Belyaev:2008yj}.
The Higgs production amplitude is proportional to the coupling strength of the Higgs to the vector states.
At the effective Lagrangian level this coupling is a free parameter with its value depending, at least, on the coupling $s$ \cite{Foadi:2008xj}. We shall use $s=0$ which still allows for a reasonable estimate of the expected order of magnitude for this process.
In Figure~\ref{eehzprod} we present the $HZ$ production
cross section, compared the SM background as a
function of the center-of-mass energy.
The inclusion of the composite vectors leads to an highly enhanced cross section with respect to the SM one in most parts of the parameter space.
We also plot the differential cross section as a function of the $HZ$ mass in Fig.~\ref{eehzprod} (bottom). In contrast to the dimuon production the peaks associated to the heavy vector states $R_{1,2}^0$ are clearly visible also for $\tilde g=5, 8$.
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/2l4j_MA750_MH200_log.eps}%
\includegraphics[width=0.45\textwidth]{figures/2l4j_MA750_MH400_log.eps}
\includegraphics[width=0.45\textwidth]{figures/2l4j_MA950_MH200_log.eps}%
\includegraphics[width=0.45\textwidth]{figures/2l4j_MA950_MH400_log.eps}
\caption{\label{2l4jlm}Distribution of events for the $e^+e^- \to 2\ell+4j$ signature at $\sqrt{s}=1$~TeV and $L=100\ fb^{-1}$. Continuous lines are the distribution for signal events, and dotted lines are the background. The value of $\tilde{g}=2$ corresponds to the black solid line while the magenta corresponds to $\tilde{g}=5$.}
\end{figure}
\begin{figure}[htb!]
\includegraphics[width=0.45\textwidth]{figures/2l4j_MA2250_MH400_log.eps}%
\includegraphics[width=0.45\textwidth]{figures/2l4j_MA2900_MH400_log.eps}%
\caption{\label{2l4jhm} The same as Fig.~\ref{2l4jlm} but for $\sqrt{s}=3$~TeV.}
\end{figure}
Finally, let us consider a full simulation for the signatures $e^+e^- \to ZH \to ZZZ \to 6\ell$ and $e^+e^- \to ZH \to ZW^+W^- \to 2\ell + 4j$. The total cross section for the $6\ell$ signature is very low due to the small $Z \to 2\ell$ branching ratio. Scaling the Higgs production cross section from above we find that $\sigma(e^+e^-\to6\ell)\sim 0.1$~fb at most, so a very high luminosity is required for studying this channel.
Our conclusion differs from that of \cite{Barger:2009xg}, where the $6\ell$ final state in $pp\to HZ$ was suggested as a promising channel to study $Z'$ resonances at the LHC. This happens as in our model the heavy vector states couple to the SM fermions only through mixing with the electroweak gauge bosons, and therefore their fermionic couplings are suppressed.
Hence we concentrate on the $2\ell + 4j$ signature, which has considerably larger cross section.
We adapt as acceptance cuts for charged leptons and jets:
\begin{eqnarray}\label{acceptancecuts}
|p_\perp^\ell| > 10~{\rm GeV},\qquad|p_\perp^j| &>& 20~{\rm GeV},\qquad |\cos\theta^{\ell,j}| < 0.95 \ .
\end{eqnarray}
Here $\theta^{\ell,j}$ is, as before, either the angle between the beam axis and the lepton or the beam axis and the jet. We also require that all jet pairs are well separated:
\begin{equation}
\Delta \eta^2 + \Delta \phi^2 > 0.3^2 \ ,
\end{equation}
where $\Delta \eta$ is the pseudorapidity difference between the total jet momenta and $\Delta \phi$ is the azimuthal separation. Also, the energies of the leptons and jets are smeared as explained above.
Event distributions for the signal $e^+e^- \to HZ \to 2\ell + 4j$ at $\sqrt{s}=1$~TeV is presented in Fig.~\ref{2l4jlm} for the integrated luminosity of $100$~fb$^{-1}$ \footnote{We only included the dominant $H \to WW \to 4j$ decay. The signal cross section will be slightly enhanced if other channels are added.}. The dotted lines are our estimate for the background identified with the SM process $e^+e^- \to ZW^+W^- \to 2\ell + 4j$.
For comparison we also show the SM cross section for the same signature (dashed lines), i.e., $e^+e^- \to HZ \to 2\ell + 4j$ in the absence of heavy composite vectors in the intermediate state.
We also repeated the study for $\sqrt{s}=3$~TeV and for higher mass of the composite states in Fig.~\ref{2l4jhm}.
\begin{table}
\begin{eqnarray}
\begin{array}{c|c|c||c|c}
M_A(\textrm{GeV}) & \quad M_H(\textrm{GeV}) & \quad \ \tilde g \quad \mbox{} &\quad L(\textrm{fb}^{-1})\ \textrm{for}\ 3\sigma &\quad L(\textrm{fb}^{-1})\ \textrm{for}\ 5\sigma \\
\hline\hline
750 & \quad 200 & \quad \ 2 \quad \mbox{} & \quad 18 &\quad 59 \\
750 & \quad 200 & \quad \ 5 \quad \mbox{} & \quad 4.0 &\quad 12 \\
750 & \quad 400 & \quad \ 2 \quad \mbox{} & \quad 80 &\quad 230 \\
750 & \quad 400 & \quad \ 5 \quad \mbox{} & \quad 7.8 &\quad 24 \\
\hline
950 & \quad 200 & \quad \ 2 \quad \mbox{} & \quad 0.28 &\quad 0.86 \\
950 & \quad 200 & \quad \ 5 \quad \mbox{} & \quad 0.48 &\quad 1.5 \\
950 & \quad 400 & \quad \ 2 \quad \mbox{} & \quad 0.61 &\quad 1.9 \\
950 & \quad 400 & \quad \ 5 \quad \mbox{} & \quad 0.71 &\quad 2.2
\end{array} \nonumber
\end{eqnarray}
\caption{Estimates for required luminosities for $3\sigma$ and $5\sigma$ signals of the axial vector resonance in associated Higgs production at $\sqrt{s}=1$~TeV.}
\label{table:ss1est}
\end{table}
\begin{table}
\begin{eqnarray}
\begin{array}{c|c|c||c|c}
M_A(\textrm{GeV}) & \quad M_H(\textrm{GeV}) & \quad \ \tilde g \quad \mbox{} &\quad L(\textrm{fb}^{-1})\ \textrm{for}\ 3\sigma &\quad L(\textrm{fb}^{-1})\ \textrm{for}\ 5\sigma \\
\hline\hline
2250 & \quad 400 & \quad \ 2 \quad \mbox{} & \quad 5.0 &\quad 12 \\
2250 & \quad 400 & \quad \ 5 \quad \mbox{} & \quad 56 &\quad 170 \\
\hline
2900 & \quad 400 & \quad \ 2 \quad \mbox{} & \quad 5.2 &\quad 13 \\
2900 & \quad 400 & \quad \ 5 \quad \mbox{} & \quad 77 &\quad 210
\end{array} \nonumber
\end{eqnarray}
\caption{Estimates for required luminosities for $3\sigma$ and $5\sigma$ signals of the axial vector resonance in associated Higgs production at $\sqrt{s}=3$~TeV.}
\label{table:ss3est}
\end{table}
Let us then estimate the luminosity which is necessary to study this process. We use the following cuts
\begin{eqnarray}\label{lastcuts}
|M_{2\ell4j}-M_{R}| & < & \max\left(0.5\Gamma_{R},0.5 \sqrt{\frac{M_{R}}{\rm GeV}} {\rm GeV}\right) \ ,\nonumber \\
|M_{4j}-M_{H}| & < & \max\left(0.5\Gamma_{H},0.5 \sqrt{\frac{M_{H}}{\rm GeV}} {\rm GeV}\right)
\end{eqnarray}
so that the four jets are produced by the Higgs decay. The significance of the signal is then calculated as detailed above. The luminosities corresponding to the distributions of Figs.~\ref{2l4jlm} and~\ref{2l4jhm} are listed in Tables~\ref{table:ss1est} and~\ref{table:ss3est}. We only considered the most significant axial vector peak which is $R_1$ in the $\sqrt{s}=1$~TeV scenario and $R_2$ for $\sqrt{s}=3$~TeV.
\begin{figure}[thb!]
\includegraphics[width=0.45\textwidth]{figures/combined_lumi_gt2.eps}%
\includegraphics[width=0.45\textwidth]{figures/combined_lumi_gt5.eps}
\caption{\label{reach}Estimated luminosity required for a $5\sigma$ discovery of {\em any} vector resonance in {\em any} of the channels of Figs.~\ref{lumiss1}, \ref{lumiss3}, and~\ref{lumissvar} at LCs as a function of the resonance mass. The solid black lines are the estimates for LHC \cite{Belyaev:2008yj}, dash-dotted magenta (dashed red) lines are for a 1~TeV (3~TeV) LC, and the dotted blue lines are for LC in the scanning mode ($\sqrt{s} = M(R)+30$~GeV). Left: $\tilde g=2$, right: $\tilde g=5$.}
\end{figure}
\section{Comparison to LHC and conclusions}
We combine the reach estimates of the signatures (1) and (2) (Figs.~\ref{lumiss1}, \ref{lumiss3}, and~\ref{lumissvar}) in Fig.~\ref{reach}, and compare them to the LHC reach \cite{Belyaev:2008yj}. We only present the luminosity required for a $5 \sigma$ discovery of the most significant resonance in either of the signatures ($e^+e^- \to R_{1,2}^0 \to \ell^+\ell^-$ and $e^+e^- \to R_{1,2}^0 \to \nu+\ell+2j$)
for the LC curves. For the LHC at $\sqrt{s} = 14$~TeV the included channels are $pp \to R_{1,2}^0 \to \ell^+\ell^-$, $pp \to R_{1,2}^\pm \to \ell^\pm\nu$, and $pp \to R_{1,2}^\pm \to 3\ell+\nu$. Since the luminosity for LC in the scanning mode (the dotted blue curve) is mostly below the LHC one (solid black), the LC can confirm, or actually discover, resonances found or missed at the LHC already at relatively low luminosities. We also observe that the LC working at fixed energy improves the LHC reach when the resonance masses are near the beam energy. In particular, a 3~TeV LC can discover resonances at large values of their masses $\gtrsim 2$~TeV better than the LHC. We recall that heavy spin-one resonances, in the range of energy accessible to LCs, are able to delay the unitarity violation of the WW scattering till several TeVs \cite{Foadi:2008ci}.
On the other hand these plots suggest that in the low mass region (below 1~TeV) for $\tilde g=5$ there is no improvement with respect to the LHC. However, we have not included the $e^+e^- \to R_{1,2} \to HZ$ channel for which the final state is more involved which could also improve the LC reach in this region of parameter space. In fact, the lightest (axial) vector resonance decays dominantly to $HZ$, and as seen from Table~\ref{table:ss1est}, including the signature (3) is expected to decrease the required LC luminosities by a factor of about $10^{2}$ for $\tilde{g} = 5$ and $M_A$ less than a TeV.
We have analyzed the reach of LCs for models of dynamical electroweak symmetry breaking and shown that they can be used to efficiently discover composite vector states with a fairly low luminosity.
\acknowledgments
We thank for discussions A.~Belyaev. MTF acknowledges a VKR Foundation Fellowship. MJ is supported in part by Regional Potential program of the E.U.FP7-REGPOT-2008-1: CreteHEPCosmo-228644 and by Marie Curie contract PIRG06-GA-2009-256487.
|
\section{Introduction}
The multiple quantum (MQ) NMR dynamics in solids \cite{BMGP} is extremely usefull for an investigation of solid structures and dynamical processes therein, for a counting the number of spins in impurity clusters \cite{BGPGR,H}, for a simplification of the standard NMR spectra \cite{WWP}. Usually the MQ NMR experiments deal with samples where the nuclear spin system is initially prepared in the thermal equilibrium state in the strong external magnetic field \cite{BMGP}. However, it is possible to carry out the MQ NMR experiments with samples prepared in different initial states \cite{FG}. In particular, one can prepare a spin system in the dipolar ordered state \cite{G} using either the adiabatic demagnetization
method in the rotating reference frame (RRF) \cite{G,SH} or the two-pulse Broekaert-Jeener
sequence \cite{G,JB}. The MQ NMR dynamics with this initial state in small spin systems has been simulated in refs. \cite{DFKFG,DFKFG2}. Using the dipolar ordered initial state in the MQ NMR experiment one should expect the earlier appearance of multiple spin clusters and correlations in comparison with the MQ NMR experiment with the thermal equilibrium initial state in the strong external magnetic field. In fact, the analysis of the MQ NMR experiments in the six-eight spin systems \cite{DFKFG,DFKFG2} demonstrates that the six-eight order MQ coherences appear earlier in
the experiment with the dipolar ordered initial state.
One of the basic problems for the theoretical description of the MQ NMR experiments is an exponential growth of the density matrix dimensionality with the increase in the number of spins.
Therefore the modern numerical methods have been developed for simulation of the MQ NMR dynamics, which are based either on Chebyshev polynomial expansion \cite{DRKN} or quantum parallelism \cite{ZCAPCDRV}. However, these methods allow one to study the MQ NMR dynamics in systems of no more than several tens of spins. A significant progress in this direction has been achieved in simulation of the MQ NMR dynamics in the system of equivalent spins \cite{BKHWW,FR,DFFZ}, which may be prepared, for instance, filling the closed nanopore with the gas of spin-carring
molecules (or atoms).
The matter is that, due to the special symmetry of the Hamiltonian governing the dynamics in such spin system, it becomes possible to study the MQ NMR dynamics in systems of hundreds of spins and even more \cite{BKHWW,FR}. The nature of the above mentioned symmetry can be clarified as follows. { As far as the characteristic time between two successive collisions with the nanopore walls is} several orders less that the time of mutual flip-flops of any two nuclear spins (which is defined by their dipole-dipole interaction (DDI)) \cite{BKHWW,FR}, it seems to be reasonable to use the averaged DDI, which may be obtained by averaging over the spin positions in the nanopore \cite{BKHWW}. This means that the constant of the averaged DDI remains the same for any pair of spins in the nanopore {\cite{BKHWW}},
so that the nuclear spins become equivalent. For this reason, the Hamiltonian of the nuclear spin DDI in the nanopore commutes with the operator of the square of the
total spin angular momentum $I^2$ \cite{DFFZ,DFFZ2}. Thus, it becomes possible to use the basis of the common eigenfunctions for the operator of the square of the
total spin momentum $I^2$ and of its projection $I_z$ on the direction of the external magnetic field instead of the standard multiplicative basis of the eigenfunctions of $I_z$, which yields an exponential growth of the Hilbert space dimensionality with the increase in the number of spins \cite{DFFZ,DFFZ2}. As a result, { we simplify calculations which allows us to succeed in} both the investigation of the MQ NMR dynamics in systems of 200 -- 600 spin-1/2 particles \cite{DFFZ} and in the study of the dependence of the coherence relaxation time on the MQ NMR coherence order and the number of spins \cite{DFZ}. Emphasize that the nuclear spin system with the thermodynamic equilibrium initial state in the strong external magnetic field is used in refs.\cite{DFFZ,DFFZ2,DFZ}.
The MQ NMR dynamics in the large system of equivalent spins with the dipolar ordered initial state is studied in the present paper.
The theory of MQ NMR dynamics of equivalent spins with this initial state is given in Sec.\ref{Section:theory}. The dependence of the MQ NMR coherence intensities on the coherence orders (the profiles of MQ NMR coherence intensities) for systems of 200-600 spins is represented in Sec.\ref{Section:numerics}. The MQ NMR dynamics in systems with the
dipolar ordered initial state is compared with the dynamics in systems with the thermal equilibrium initial state in the strong external magnetic field in Sec.\ref{Section:comp}. The basic results are collected in Sec.\ref{Section:conclusion}.
\section{The MQ NMR coherence intensities in systems of equivalent spins prepared in the dipolar ordered initial state}
\label{Section:theory}
We consider the system of equivalent spin-1/2 particles with the dipole-dipole interaction (DDI) in the strong external magnetic field. The secular part of the DDI Hamiltonian \cite{G} reads:
\begin{eqnarray}
\label{Hdz}
H_{dz} = \sum_{j<k} D_{jk}(2 I_{jz} I_{kz} - I_{jx} I_{kx} -I_{jy} I_{ky}),
\end{eqnarray}
where $D_{jk}=\frac{\gamma^2 \hbar}{2 r_{jk}^3} ( 1-3 \cos^2\theta_{jk})$ is the
constant of DDI, $\gamma$ is the gyromagnetic ratio, $r_{jk}$ is the distance between
the $j$th and $k$th spins, $\theta_{jk}$ is the angle between the internuclear vector
$\vec r_{jk}$ and the external magnetic field $\vec B_0$, and $I_{j\alpha}$
($\alpha=x,y,z$)
is the $j$th spin projection operator on the axis $\alpha$.
Using either the adiabatic demagnetization in the rotating reference frame \cite{G,SH} or the Broekaert-Jeener two-pulse sequence \cite{G,JB} one can prepare the spin system in the dipolar ordered initial state with the following density matrix:
\begin{eqnarray}\label{HT}
\rho(0) = \frac{1}{Z} \exp(-\beta H_{dz}) \approx \frac{1}{2^N} ( 1 -\beta H_{dz}),
\end{eqnarray}
where $\beta=\frac{\hbar}{kT}$ is the inverse spin temperature ($k$ is the Boltsman constant, $T$ is the temperature), $Z={\mbox{Tr}} \exp(-\beta H_{dz})$ is the partition function and $N$ is the number of spins.
If the system under consideration consists of the spin carring molecules (atoms) in the closed nanopore, then DDI is averaged (incompletely) by the fast molecular diffusion so that the constants of DDI for any spin pair equal each other, i.e. $D_{jk}\equiv D$ \cite{BKHWW,FR}. As a result, the Hamiltonian (\ref{Hdz}) may be written as follows \cite{FR}:
\begin{eqnarray}\label{Hdz:equiv}
\bar H_{dz} =\frac{D}{2} (3 I_z^2 - I^2),
\end{eqnarray}
where $I_\alpha=\sum_{i=1}^N I_{i\alpha}$, ($\alpha=x,y,z$), $I^2=I_x^2+I_y^2+I_z^2$ is the square of the spin angular momentum.
It is important to justify that the high temperature approximation (\ref{HT}) is applicable to the system of equivalent spins.
The simple analisis \cite{D} demonstrates that approximation (\ref{HT}) for the system with the Hamiltonian (\ref{Hdz:equiv}) is valid if
\begin{eqnarray}\label{HTc}
\beta D N \ll 1.
\end{eqnarray}
The systems considered in our work have $N=200-600$, so that the condition (\ref{HTc}) is satisfied.
\begin{figure*}
\epsfig{file=fig1
, scale=0.5,angle=0
}
\caption{The basic scheme of the four period MQ NMR experiment.
}
\label{Fig:NMR}
\end{figure*}
The MQ NMR experiment consists of four basic periods shown in Fig.\ref{Fig:NMR}: the preparation period $\tau$, the evolution period $t$, the mixing period $\tau$ and the detection. The MQ NMR coherences are generated on the preparation period due to the irradiation of the sample by the multiple eight-pulse sequence of the resonance pulses \cite{BMGP}. Let us express MQ NMR coherences in terms of the density matrix of the preparation period.
For this purpose note that the averaged Hamiltonian $\bar H_{MQ}$ (nonsecular two-spin/two-quantum Hamiltonian \cite{BMGP}) describing the MQ dynamics in the system of equivalent spins during the preparation period in the rotating reference frame may be written as follows \cite{DFFZ,DFFZ2}:
\begin{eqnarray}\label{HMQ}
\bar H_{MQ} =- \frac{D}{4} \{ (I^+)^2 + (I^-)^2\},
\end{eqnarray}
where $I^+$ and $I^-$ are the raising and lowering operators ($I^\pm=I_x \pm i I_y$).
In order to investigate the MQ NMR dynamics, one has to find the density matrix $\rho(\tau)$ for the spin system solving the Liouville equation \cite{G}:
\begin{eqnarray}
\label{Liouville}
i \frac{d\rho}{d\tau} = [\bar H_{MQ},\rho(\tau)],
\end{eqnarray}
with the initial condition $\rho(0)=\bar H_{dz}$. This initial condition is obtained from eqs. (\ref{HT}) and (\ref{Hdz:equiv}) by dropping both the unit operator and the factor ($-\beta/2^N$), which are not significant for the MQ NMR dynamics.
Taking into account the Hamiltonians on the different periods of the MQ NMR experiment (these Hamiltonians are shown in Fig.\ref{Fig:NMR}) one can write the expression for the dipolar energy $\langle \bar H_{dz} \rangle(\tau,t)$ after the mixing period of the MQ NMR experiment (Fig.\ref{Fig:NMR})
as follows:
\begin{eqnarray}\label{H_dz}
\langle \bar H_{dz}\rangle(\tau,t)&=&\frac{{\mbox{Tr}}\{U^+(\tau) e^{-i\Delta t I_z} U(\tau)
\bar H_{dz} U^+(\tau)
e^{i\Delta t I_z} U(\tau) \bar H_{dz}\}}{{\mbox{Tr}}\{ \bar H_{dz}^2\}}=\\\nonumber
&&
\frac{{\mbox{Tr}}\{ e^{-i\Delta t I_z} \rho(\tau)
e^{i\Delta t I_z} \rho(\tau)\}}{{\mbox{Tr}}\{ \bar H_{dz}^2\}},
\end{eqnarray}
where { $\rho(\tau)=U(\tau) \bar H_{dz} U^+(\tau)$} is the solution to Eq.(\ref{Liouville}) and $U(\tau)=\exp(-i\bar H_{MQ} \tau)$.
It is convenient to represent the solution to Eq.(\ref{Liouville}) as the series \cite{FL}
\begin{eqnarray}\label{rhok}
\rho(\tau) =\sum_k\rho_k(\tau),
\end{eqnarray}
where { $\rho_k(\tau)$ obeys the relationship $[I_z,\rho_k(\tau)]=k\rho_k(\tau)$ which can be considered as the definition of $\rho_k(\tau)$ (in other words, $\rho_k(\tau)$ collects those entries of the density matrix $\rho(\tau)$ which are responsible for the $k$th order MQ NMR coherence)}. Then eq.(\ref{H_dz}) reads
\begin{eqnarray}\label{H_dz2}
\langle \bar H_{dz}\rangle(\tau)=\sum_k e^{-ik\Delta t}\frac{{\mbox{Tr}}\{
\rho_k(\tau)\rho_{-k}(\tau)
\}}{
{\mbox{Tr}} \{ \bar H_{dz}^2\}}= \sum_k e^{-ik\Delta t} J_k(\tau),
\end{eqnarray}
where the $k$th order MQ NMR coherence intensity $J_k$ is defined as follows \cite{FL}:
\begin{eqnarray}\label{Jk}
J_k(\tau)=\frac{
{\mbox{Tr}} \{ \rho_k(\tau)\rho_{-k}(\tau)\}}{
{\mbox{Tr}} \{ \bar H_{dz}^2\}
}.
\end{eqnarray}
This formula will be used in the numerical analysis of the MQ NMR coherences in Sec.\ref{Section:numerics}.
It is easy to write the explicit expression for ${\mbox{Tr}} \{ \bar H_{dz}^2\}$.
First of all, using eq. (\ref{Hdz:equiv}) we may write
\begin{eqnarray}\label{tr}
&&
{\mbox{Tr}} \{ \bar H_{dz}^2\} =\frac{D^2}{4} {\mbox{Tr}}(3 I_z^2 -I^2)^2 =\\\nonumber
&&
\frac{D^2}{4} {\mbox{Tr}}\{
9 I_z^4 - 3 I_z^2 I^2 + I^2 ( I^2 -3 I_z^2)
\}=\\\nonumber
&&
\frac{3D^2}{4} {\mbox{Tr}}\{
3 I_z^4 - I_z^2 I^2
\}=\frac{3D^2}{2} {\mbox{Tr}}\{
I_z^4 - I_x^2 I_z^2
\}.
\end{eqnarray}
Here we take into account that
\begin{eqnarray}\label{tr2}
{\mbox{Tr}}\{I^2 ( I^2 -3 I_z^2)\}=0, \;\; {\mbox{Tr}}\{I_x^2 I_z^2\} = {\mbox{Tr}}\{I_y^2 I_z^2\}.
\end{eqnarray}
Next, it is simple to obtain the explicit expressions for ${\mbox{Tr}} \{I_z^4\}$ and
${\mbox{Tr}}\{ I_x^2 I_z^2 \}$:
\begin{eqnarray}\label{tr3}
{\mbox{Tr}}\{ I_z^4\} = 2^{N-4} N (3 N -2),\;\;\;
{\mbox{Tr}} \{I_x^2 I_z^2\} = 2^{N-4} N^2.
\end{eqnarray}
Finally, eqs.(\ref{tr}-\ref{tr3}) yield the following result:
\begin{eqnarray}
{\mbox{Tr}} \{ \bar H_{dz}^2\} =3 N(N-1) 2^{N-4} D^2.
\end{eqnarray}
Taking into account the structure of the MQ Hamiltonian $\bar H_{MQ}$, it may be readily shown that only the even order MQ NMR coherences appear in our numerical experiment and the coherence order may not exceed the number of spins $N$ \cite{BMGP,FL}.
It is obvious that the MQ Hamiltonian (\ref{HMQ}) commutes with the square of the total spin momentum operator $I^2$ \cite{DFFZ,DFFZ2,DFZ}. As far as $[I^2,I_z]=0$, it is possible to use the basis of the common eigenfunctions of operators $I^2$ and $I_z$ for the description of the MQ NMR dynamics.
Namely this fact allows one to { avoid} the problem of the exponential growth of the matrix dimensionality with the increase in the number
of spins, which appears in the
traditional multiplicative basis \cite{BMGP} of the eigenfunctions of the operator $I_z$. In the new basis, the MQ Hamiltonian $\bar H_{MQ}$ consists of blocks $\bar H_{MQ}^S$ corresponding to the different values of the total spin momentum $S$ ($\hat S^2 = S(S+1)$, $S=\frac{N}{2}, \frac{N}{2}-1, \frac{N}{2}-2,\cdots, \frac{N}{2}-\left[\frac{N}{2}\right]$, $[i]$ is an integer part of $i$):
\begin{eqnarray}
\bar H_{MQ}={\mbox{diag}}\{ \bar H^{\frac{N}{2}}_{MQ} ,\bar H^{\frac{N}{2}-1}_{MQ},
\dots ,\bar H^{\frac{N}{2}-\left[\frac{N}{2}\right]}_{MQ}\}.
\end{eqnarray}
As far as the Hamiltonian $\bar H_{dz}$ (\ref{Hdz:equiv}) is diagonal in the basis of the common eigenfunctions of the operators $I^2$ and $I_z$, then the density matrix $\rho(0)$ may be also splitted into the diagonal blocks $\rho^S(0)$ with $S=\frac{N}{2}, \frac{N}{2}-1, \frac{N}{2}-2,\cdots, \frac{N}{2}-\left[\frac{N}{2}\right]$. Consequently, the matrix $\rho(\tau)$ has also the diagonal block structure with blocks $\rho^S(\tau)$. Thus, the problem becomes separated into the set of independent problems for each $(2S+1)\times (2S+1)$-dimensional block $\rho^S(\tau)$ which is a solution to the Liouville equation (\ref{Liouville}) with the Hamiltonian $\bar H_{MQ}^S$. Of course, expansion (\ref{rhok}) may be applied to each block $\rho^S$. The contribution $J_{k,S}$ from the block $\rho^S$ to the intensity $J_k$ of the $k$th
order coherence is defined by the obvious formula \cite{DFFZ}:
\begin{eqnarray}\label{JkS}
J_{k,S}(\tau)=\frac{
{\mbox{Tr}}\{
\rho^S_k(\tau) \rho^S_{-k}(\tau)
\}}{{\mbox{Tr}} \bar H_{dz}^2},
\end{eqnarray}
where $\rho^S_k(\tau)$ is the contribution from the matrix $\rho^S$ to the $k$th order coherence. One has to take into account that each block $\rho^S(\tau)$ is degenerated with the multiplicity $n_N(S)$ \cite{LL,DFFZ}:
\begin{eqnarray}
n_N(S)=\frac{N!(2 S+1)}{(\frac{N}{2} + S +1)!(\frac{N}{2} -S)!},\;\;0\le S\le \frac{N}{2}.
\end{eqnarray}
As a result, the observable intensities $J_k(\tau)$ ($-N\le k\le N$) are following \cite{DFFZ,DFFZ2}:
\begin{eqnarray}
J_k(\tau) =\sum_Sn_N(S) J_{k,S}(\tau).
\end{eqnarray}
Remember that the dimensionality of each block $\bar H_{MQ}^S$ of the MQ NMR
Hamiltonian $\bar H_{MQ}$ is $(2S+1)$. Taking into account the block degeneration we
obtain the correct value for the matrix dimensionality of both the Hamiltonina $\bar H_{MQ}$ and the density matrix $\rho$ \cite{DFFZ}:
\begin{eqnarray}
\sum_S n_N(S) (2S+1)= 2^N,
\end{eqnarray}
which is valid for the system of $N$ interacting spin-1/2 particles.
The numerical algorithms describing the MQ NMR dynamics in the systems of equivalent spins with the thermal equilibrium initial state in the strong external magnetic field
have been developed in \cite{DFFZ,DFFZ2,DFZ}. With minor corrections, these algorithms may be used for the simulation of the dynamics of the MQ NMR coherences in the spin system with the dipolar ordered initial state. In particular, the integral of motion related with the MQ NMR Hamiltonian invariance with respect to the rotation over the angle $\pi$ around $z$-axis \cite{DFGM} is also present. Thus, for the odd $N$, it is enough to solve the problem for the MQ NMR Hamiltonians with two times lower matrix dimensionality and then double the resulting intensities \cite{DFGM}. We use the spin systems with odd $N$ in all numerical calculations in the next section.
\section{The numerical analysis of the MQ NMR profiles}
\label{Section:numerics}
Using the method developed in the previous section we investigate the profiles of MQ NMR coherences. As far as all spins are ''nearest neighbors'' in the system of equivalent spins, $N$ spin cluster appears already after the time interval $\tau\sim 1/D$. However, some reorganization of this cluster is required for the MQ coherence formation \cite{DFFZ}. The analysis of the MQ NMR coherence dynamics demonstrates that the quasistationary profile of MQ NMR coherences is created during {$\bar \tau\sim 2$} ($\bar\tau=D\tau$ is the dimensionless time hereafter) and remains fast oscillating for $\bar\tau>2$. Because of these oscillations it is convenient to use the averaged intensities $\bar J_k$ \cite{DFFZ} instead of intensities $J_k$ itselves. We estimate the dimensionless averaging time interval as $T\sim 2\pi /|\lambda^{min}_{3/2}|\approx 7.255 $, where $\lambda^{min}_{3/2}=\sqrt{3}/2$ is the minimal eigenvalue of the Hamiltonian \cite{DFFZ,DFFZ2}. For our convenience, we take $T=8$ so that
\begin{eqnarray}\label{barI}
\bar J_k =\frac{1}{T}\int_{2}^{2+T} J_k(\bar\tau) d\bar\tau.
\end{eqnarray}
It is observed that $\bar J_k$ do not significantly vary with increase in $T$, so that the definition of the averaged intensities given by eq. (\ref{barI}) is valid. Although the dynamics of all coherences has been found in the numerical simulations, the intensities of the high order coherences are negligible, so that we represent the intensities of the MQ NMR coherences up to the 50th order in all figures below. The profiles of MQ NMR coherence intensities for systems of 201, 401 and 601 spins with dipolar ordered initial state are shown in Fig.\ref{Fig:prof}. These profiles are similar to those which have been found for the systems with the thermal equilibrium initial state in the strong external magnetic field \cite{DFFZ,DFFZ2}. Similar to refs.\cite{DFFZ,DFFZ2}, the averaged intensities of MQ NMR coherences are separated into two families:
\begin{eqnarray}\label{fam}
&&
\Gamma_1=\{\bar J_{4 k-2},\;\;k=0,\pm 1,\pm 2,\dots\}, \\\nonumber
&&
\Gamma_2=\{\bar J_{4 k},\;\;k=\pm 1,\pm 2,\dots\},
\end{eqnarray}
with the zero order coherence intensity $\bar J_0$ does not corresponds to any of these families.
Each family may be approximated by the smooth distribution function as follows:
\begin{eqnarray}\label{distr}
\bar J_{2k}\approx \left\{
\begin{array}{ll}
\displaystyle A_1\Big(1+\sum_{i=1}^4 (-1)^i a_{1i} (2 |k|)^i\Big) e^{-2\alpha_1 |k|}, &k=\pm 1,\pm 3,\dots\cr\displaystyle
A_2\Big(1+\sum_{i=1}^6(-1)^i a_{2i} (2 |k|)^i\Big) e^{-2\alpha_2 |k|}, &k=\pm 2,\pm 4,\dots\cr
\end{array}
\right.,
\end{eqnarray}
where parameters $A_i$, $a_{ij}$ and $\alpha_i$ for spin systems with $N=201$, 401 and 601 are collected in Table 1.
The algorithm defining the approximation parameters is similar to that suggested in ref.\cite{DFFZ}.
\begin{table}[!htb]
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline
N &$A_1$ &$a_{11}$ & $a_{12}$ & $a_{13}$ & $a_{14}$ &$\alpha_1$\\\hline
201&$8.498\times 10^{-2}$&$3.133\times 10^{-1}$&$8.025\times 10^{-2}$&$6.109\times 10^{-3}$&$2.739\times 10^{-4}$&$2.437\times 10^{-1}$\\
401&$4.546\times 10^{-2}$&$1.694\times 10^{-1}$&$3.357\times 10^{-2}$&$1.649\times 10^{-3}$&$5.739\times 10^{-5}$&$1.638\times 10^{-1}$\\
601&$3.546\times 10^{-2}$&$1.456\times 10^{-1}$&$2.471\times 10^{-2}$&$1.028\times 10^{-3}$&$2.855\times 10^{-5}$&$1.342\times 10^{-1}$\\\hline
\end{tabular}
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline
N &$A_2$ &$a_{21}$ & $a_{22}$ & $a_{23}$ & $a_{24}$ & $a_{25}$ \\\hline
201&$2.217$\hspace*{1.3cm}&$4.029\times 10^{-1}$&$7.522\times 10^{-2}$&$7.638\times 10^{-3}$&$4.624\times 10^{-4}$&$1.562\times 10^{-5}$ \\
401&$1.411$\hspace*{1.3cm}&$3.587\times 10^{-1}$&$5.648\times 10^{-2}$&$4.616\times 10^{-3}$&$2.116\times 10^{-4}$&$5.136\times 10^{-6}$ \\
601&$1.210$\hspace*{1.3cm}&$3.441\times 10^{-1}$&$5.123\times 10^{-2}$&$3.924\times 10^{-3}$&$1.664\times 10^{-4}$&$3.686\times 10^{-6}$ \\\hline
\end{tabular}
\begin{tabular}{|c|c|c|}
\hline
N & $a_{26}$ &$\alpha_2$\\\hline
201&$2.510\times 10^{-7}$&$3.039\times 10^{-1}$\\
401&$5.533\times 10^{-8}$ &$2.183\times 10^{-1}$\\
601&$3.512\times 10^{-8}$ &$1.940\times 10^{-1}$\\\hline
\end{tabular}
\label{Table:HPST2}
\caption{ The parameters $A_i$, $a_{ij}$ and $\alpha_i$ of the distribution function given by eq.(\ref{distr}) for $N=201$, 401 and 601}
\end{table}
Similar to the profiles of the MQ NMR coherence intensities obtained for the systems of equivalent spin-1/2 particles with the thermal equilibrium initial state in the strong external magnetic field \cite{DFFZ},
profiles for the systems with the dipolar ordered initial state seemed out to be exponential. This conclusion agrees with results obtained during an elaboration of numerous MQ NMR spectra \cite{LHG} and contradicts the phenomenological theory \cite{BMGP} predicting the Gauss profiles of the MQ NMR coherence intensities.
\begin{figure*}
\epsfig{file=fig2
, scale=0.5,angle=270
}
\caption{Coherence intensities profiles for spin systems with $N=201,401,601$. The inset demonstrates that zero order coherence does not belong neither to $\Gamma_1$ nor to $\Gamma_2$. Only intensities of positive order coherences are presented. }
\label{Fig:prof}
\end{figure*}
Liouville equation (\ref{Liouville}) and formulas for the MQ NMR coherence intensities (\ref{Jk}) yield the conservation law of the sum of the coherence intensities \cite{LHGl}:
\begin{eqnarray}
\sum_k \bar J_k(\tau)=1.
\end{eqnarray}
This law together with the approximating formula (\ref{distr}) allows one to find a good approximation to the zero order coherence intensity. We compare the calculated values of $\bar J_0$ (see eq.(\ref{Jk})) with the values
$\bar J_0^{appr}=1-2 \sum_{k=1}^\infty J_{2k}$ found using the above conservation law and distribution function (\ref{distr}):
\begin{eqnarray}
\bar J_0=\left\{\begin{array}{ll}
2.519 \times 10^{-1},& N=201\cr
2.417 \times 10^{-1},& N=401\cr
2.361 \times 10^{-1},& N=601
\end{array}\right.,\;\;\;\bar J_0^{appr}=\left\{\begin{array}{ll}
2.501 \times 10^{-1},& N=201\cr
2.381 \times 10^{-1},& N=401\cr
2.387 \times 10^{-1},& N=601
\end{array}\right. .
\end{eqnarray}
Some discrepancy between $\bar J_0$ and $\bar J_0^{appr}$ appears because we take into account only coherences up to the 50th order in constructing the distribution function (\ref{distr}), while contributions from the higher order coherences are missed.
\iffalse
\begin{figure*}
\epsfig{file=fig3
, scale=0.5,angle=270
}
\caption{Evolution of the coherence cluster in a spin system with $N=201$ and $N=401$; cluster collects coherences with $J_k>0.003$. The formation time $\approx 1.5$, which agrees with \cite{DFZ}.}
\label{Fig:cluster}
\end{figure*}
\fi
\section{Comparison of the MQ NMR dynamics in the spin systems with
two different initial states}
\label{Section:comp}
The preparation of the system in the dipolar ordered initial state means that the two-spin correlations appear already at the initial time, unlike the standard MQ NMR experiment, where the thermal equilibrium initial state is defined by the one-spin Zeemann interaction with the external magnetic field \cite{DFKFG,DFKFG2}. This statement is justified by Fig.\ref{Fig:comp}, where the formation times {$\bar\tau_f(n)$} of different order coherences for both initial states are represented. Here the term formation time $\bar\tau_f(n)$ of the $n$th coherence means the time moment when the $n$th coherence intensity $J_n(\bar\tau)$ reaches the value $\bar J_n$ for the first time. We see that MQ NMR coherences in the system with the dipolar ordered initial state (the lower solid line) appear much earlier. This result agrees with that obtained in \cite{DFKFG} for the MQ NMR in systems with small number of spin-1/2 particles.
\begin{figure*}
\epsfig{file=fig3
, scale=0.5,angle=270
}
\caption{Coherence formation time $\bar\tau_f$ versus the coherence number for the dynamics of equivalent spins with the dipolar ordered and the thermal equilibrium in the strong external field initial states.}
\label{Fig:comp}
\end{figure*}
\begin{figure*}
\epsfig{file=fig4
, scale=0.6,angle=270
}
\caption{The MQ NMR profiles for the systems of equivalent spins with the dipolar ordered and thermal equilibrium in the strong external magnetic field initial states, $N=201$; the inset demonstrates that $\bar J_0$ and $\bar J_4$ are essentially bigger in the case of the dipolar ordered initial state.}
\label{Fig:comp_prof}
\end{figure*}
The profiles of the MQ NMR coherence intensities for the system of $N=201$ spins with both the dipolar ordered and the thermal equilibrium in the strong magnetic field initial states are compared in Fig.\ref{Fig:comp_prof}. This figure demonstrates that the discrepancy between two families of MQ NMR coherences $\Gamma_1$ and $\Gamma_2$ (see Eq.(\ref{fam})) is bigger for $n<10$ and smaller for $n>10$ in the case of the dipolar ordered initial state. The inset shows that $\bar J_0$ and $\bar J_4$ are
essentially bigger in the case of the dipolar ordered initial state.
Thus, similar to the usual NMR experiments in solids \cite{G}, the MQ NMR experiment in the systems of equivalent spins with the dipolar ordered initial state may be usefull to supplement the MQ NMR experiment with the thermal equilibrium initial state in the strong external magnetic field.
\section{Conclusions}
\label{Section:conclusion}
We have studied the MQ NMR dynamics in the systems of equivalent spin-1/2 particles with the dipolar ordered initial state. For this purpose we modify the method developed in ref.\cite{DFFZ} for the system of equivalent spin-1/2 particles with the thermal equilibrium initial state in the strong external magnetic field. Similar to ref.\cite{DFFZ}, the high symmetry of such systems allows one to investigate the dynamics in large spin systems containing hundreds of interacting spins.
We obtain dependence of the MQ NMR coherence intensities on their order (the profiles of the MQ NMR coherence intensities) in systems of 200-600 spins and demonstrate that these profiles may be well approximated by the exponential distribution functions. As far as the analogues result has been obtained in refs.\cite{DFFZ,DFFZ2} for the systems with the thermal equilibrium initial state in the strong external magnetic field, we may suppose that the exponential profiles of MQ NMR coherence intensities are fundamental fact in MQ NMR dynamics.
The theoretical results obtained in refs.\cite{ZL,ZL2} confirm this conclusion.
We demonstrate that the MQ NMR coherences appear faster in the spin systems with the dipolar ordered initial state. The MQ NMR experiments with the dipolar ordered initial states expand possibilities of the MQ NMR spectroscopy in study of the structures of solids and the dynamical processes therein.
All numerical simulations have been performed using the resources of the Joint Supercomputer Center (JSCC) of
the Russian Academy of Sciences. The work was
supported by the Program of the Presidium of Russian Academy of Sciences No.21 " Foundations of fundamental
investigations of nanotechnologies and nanomaterials".
|
\section*{Preamble}
In the past few years two papers appeared on the arXiv claiming to prove the non-existence of the Zeeman topology. The oldest one \cite{sainz1} has been withdrawn and replaced by a new one \cite{sainz2} which shares techniques and conclusions with the oldest. A review of these papers is beyond the scope of this short communication.
Suffice to say that the existence of Zeeman's topology is made obvious by a reading of the original and beautiful paper \cite{zeeman}. For the sake of clarity a constructive proof of its existence is spelt out below, essentially rephrasing an argument of Zeeman.
\section*{Constructive proof of the existence of Zeeman's topology}
The Zeeman topology on Minkowski space $M$ is defined to be the finest topology on $M$ inducing\footnote{Given a topological space $(X,T)$ and a subset $Y\subset X$ with a topology $T_Y$, we say $(X,T)$ induces $T_Y$ on $Y$ if $(Y,T_Y)$ is a subspace of $(X,T)$. In other words, each open set in $(Y,T_Y)$ is the intersection of $Y$ with an open set in $(X,T)$.} the Euclidean topology on each timelike and spacelike affine subspace (from now on called axes). To prove its existence, let us consider the collection $C$ of all subsets of $M$ that meet each axis in an open\footnote{Open w.r.t. the Euclidean topology of the axis, of course.} set. This collection satisfies the axioms for a topology (straightforward exercise, see Lemma III.3 in \cite{dossena}) and induces the Euclidean topology on each axis by construction. Given a topology $T$ on $M$ inducing the Euclidean topology on each axis, any element of $T$ meets each axis in an open set, thereby $T$ is coarser than $C$ and the proof is complete.
\section*{Side note about Larson's theorem}
In \cite{sainz2}, Larson's theorem about maximum and minimum topological spaces is quoted. Let us review this interesting theorem (see \cite{larson}). Fix a topological property $P$. A topology $T$ on some set $X$ is called \emph{maximum P} [\emph{minimum P}] when any topology on $X$ with property $P$ is coarser [finer] than $T$. Larson's theorem states that a topology on $X$ is maximum $P$ or minimum $P$ for some topological property $P$ if and only if each bijection of $X$ onto itself is a homeomorphism. Contrary to claims in \cite{sainz2}, this theorem does not apply to Zeeman's topology since the property of inducing the Euclidean topology on each axis is not a topological property (this is obvious since its very formulation requires the concept of affine subspace, which is not topologically invariant).
|
\section*{Introduction}
In this article we study the {\it motivic Donaldson-Thomas} (DT in short) {\it invariants} introduced in \cite{ks,behrend_bryan_szendroi}.
The DT invariant for a Calabi-Yau $3$-fold $Y$ is a counting invariant of coherent sheaves on $Y$,
which it is introduced in \cite{thomas-dt} as a holomorphic analogue of the Casson invariant on a real $3$-manifold.
The moduli space involves a symmetric obstruction theory and a virtual fundamental cycle \cite{behrend-fantechi-intrinsic,behrend-fantechi}.
The invariant is defined as a integration of the constant function $1$ over the virtual fundamental cycle.
The DT invariant has the other description : it coincides with the weighted Euler characteristic weighted by the Behrend function.
It is known that the moduli space of coherent sheaves on $Y$ can be locally described as the critical locus of a function which is called a {\it holomorphic Chern-Simons functional} (see \cite{joyce-song}).
The value of the Behrend function is given by the Euler characteristic of the {\it Milnor fiber} of the Chern-Simons functional \cite{behrend-dt}.
Following these results of Behrend, it is proposed in \cite{ks,behrend_bryan_szendroi} to study {\it motivic Milnor fiber} as a motivic version of the DT invariant so that we can get a refinement of the ordinary DT invariant by applying a suitable cohomology functor for the motivic one .
Such a refinement has been expected in string theory \cite{RTV,dimofte-gukov,dimofte_gukov_soibelman}.
In \cite{ks}, Kontsevich and Soibelman provided a wall-crossing formula for motivic DT invariants up to a certain identity for motivic Milnor fibers (\cite[Conjecture 4]{ks}).
The aim of this article is to give an alternative proof of the wall-crossing formula for (\cite{behrend_bryan_szendroi}'s) motivic DT invariants in a spacial setting.
\section*{Main result}
Let $(Q,W)$ be a quiver with a potential (QP in short).
In this paper, we assume that $W$ is finite, i.e. a finite linear combination of oriented cycles.
Let $\mathcal{D}_{Q,W}$ be the derived category of dg modules with finite dimensional cohomologies over the (non-complete) Ginzburg's dg algebra and
$\mathrm{mod}J(Q,W)$ be the category of finite dimensional modules over the (non-complete) Jacobi algebra which is the core of the natural bounded t-structure of $\mathcal{D}_{Q,W}$.
The moduli stack of objects is canonically described as the critical locus of a function $f_W$ on a smooth stack $\mathfrak{M}_Q$. We call the function as the {\it Chern-Simons functional}.
We define the motivic DT invariant by the virtual motive $[\mathrm{crit}(f_W)]_{\mathrm{vir}}$ (Definition \ref{defn_vir}) of the critical locus of the Chern-Simons functional \footnote{The smooth stack $\mca{M}_Q$ is described as a quotient stack divided by a special algebraic group $G$. Actually we do not work on critical loci for stacks but only for varieties and define motivic invariants by the quotient of the virtual motives by $G$.}.
For a vertex $k$ without loops, let $\mu_k(Q,W)=(Q',W')$ be the mutation in the sense of \cite{quiver-with-potentials}.
We assume that $Q'$ is the quiver mutation in the sense of Fomin-Zelevinsky and $W'$ is finite \footnote{In \cite{quiver-with-potentials}, it is shown that if the potential $W$ is generic then $Q'$ is the Fomin-Zelevinsky mutation of $Q$. Finiteness of $W$ is stronger assumption.}.
Keller-Yang showed that $\mathcal{D}_{Q,W}$ and $\mathcal{D}_{Q',W'}$ are equivalent \cite{dong-keller,keller-completion}.
We want to describe the relation between the motivic DT invariant for $(Q,W)$ and the one for $(Q',W')$.
The main theorem in this paper is the following :
\begin{thm}\label{thm_main}
Assume that
\begin{itemize}
\item $(Q,W)$ has a cut (Definition \ref{defn_cut}),
\item $k$ is a strict source of $C$ (Definition \ref{defn_strict}).
\end{itemize}
Then we have
\[
\mca{A}_{Q',W'}
\ "="\
\mathbb{E}(s_k[1])\times \mca{A}_{Q,W}\times \mathbb{E}(s_k)^{-1}
\]
where $\mca{A}_{Q,W}$ is the generating series of the motivic Donaldson-Thomas invariants and $\mathbb{E}(s_k)$ is the ``motivic dilogarithm'' (Example \ref{ex_dilog}). This is an equation in the ``{\it motivic torus}'' (Definition \ref{defn_mt}). \footnote{This is an equation of infinite power series. Since we have to make it clear in which completion we work, I use the equal sign with quotation mark $"="$.}
By taking the weight polynomial we get the following :
\[
{A}_{Q',W'}
\ "="\
\mathbb{E}_q(s_k[1])\times {A}_{Q,W}\times \mathbb{E}_q(s_k)^{-1}
\]
where ${A}_{Q,W}$ is the generating series of the ``refined Donaldson-Thomas invariants'' (Definition \ref{defn_rdt}) and $\mathbb{E}_q(s_k)$ is the quantum dilogarithm.
This is an equation in the ``{\it quantum torus}'' (Definition \ref{defn_qt}).
\end{thm}
\section*{Sketch of the proof}
\noindent
{\it \underline{First step}}
\smallskip
The first step is to show the {\it factorization property}.
Take a stability condition, then each object $\mathrm{mod}J(Q,W)$ has the unique Harder-Narashimhan filtration.
Types of the Harder-Narashimhan filtrations induce a filtration by open sets on the moduli stack (see \S \ref{subsec_ks52}).
In particular, each stratum is smooth.
Using this filtration we want a formula which describes the generating function of the motivic DT invariants as the product of the generating functions of the motivic invariants of the moduli stacks of semi-stable objects.
To get the formula, we need the following ;
\begin{quote}
Let ${X}$ be a smooth stack, $f$ be a function on ${X}$ and ${Y}\subset {X}$ be a smooth substack of codimension $d$. Then,
\begin{equation}\label{eq_motivic}
\bigl[\mathrm{crit}(f)\bigr]_{\mathrm{vir}}\overset{?}{=}
\bigl[\mathrm{crit}(f|_{{X}\backslash {Y}})\bigr]_{\mathrm{vir}}+
\mathbb{L}^{-\frac{d}{2}}\cdot \bigl[\mathrm{crit}(f|_{Y})\bigr]_{\mathrm{vir}}.
\end{equation}
\end{quote}
In \S \ref{sec_FP}, we assume that we have a cut $C$ of the QP $(Q,W)$, that is, a nonnegative grading
\[
g_C\colon Q_1\to \mathbb{Z}_{\geq 0}
\]
such that $W$ is homogeneous of degree $1$.
Then the moduli stack involves a $\mathbb{C}^*$-action so that we can apply \cite[Theorem B.1]{behrend_bryan_szendroi} (Theorem \ref{thm_BBS}).
The equation \eqref{eq_motivic} directly follows \cite[Theorem B.1]{behrend_bryan_szendroi} (see Proposition \ref{prop_31}).
\begin{rem}
In \cite{COHA}, Kontsevich-Soibelman introduce the {\it cohomological Hall algebra} (COHA in short) which provide another realization of a refinement of the DT invariant.
The factorization property for the COHA is shown in \cite[\S 5]{COHA}.
For the COHA, the Thom isomorphism is the counterpart of the equation \eqref{eq_motivic}.
\end{rem}
Applying the factorization property in our setting,
we can describe the generating function of the motivic DT invariants as the product of the generating functions of the motivic invariants of
the moduli stacks of objects in
\[
\mathcal{S}:=\{s_k^{\oplus n}\mid n\geq 0\}
\]
and
\[
{}^\bot \mathcal{S}:=\{X\in \mathrm{mod}J(Q,W)\mid \Hom (X,s_k)=0\}
\]
where $s_k$ is the simple $J(Q,W)$-module corresponding to the vertex $k$.
The generating function for $\mathcal{S}$ is given by the {\it quantum dilogarithm}.
\smallskip
\smallskip
\noindent
{\it \underline{Second step}}
\smallskip
In the same way,
we can describe the generating function for $(Q',W')$ as the product of the generating functions of the motivic invariants of
the moduli stacks of objects in
\[
(\mathcal{S}')^\bot:=\{X\in \mathrm{mod}J(Q',W')\mid \Hom (s'_k,X)=0\}.
\]
and
\[
\mathcal{S'}:=\{(s'_k)^{\oplus n}\mid n\geq 0\}
\]
where $s'_k$ is the simple $J(Q',W')$-module.
It is shown in \cite{dong-keller} that the derived equivalence is given by {\it tilting} with respect to the simple module $s_k$, that is, in the derived category we have
\[
\mathcal{S}'=\mathcal{S}[1],\quad
(\mathcal{S}')^\bot={}^\bot \mathcal{S} \quad (\text{see Figure \ref{fig:tilting}}).
\]
\begin{figure}[h]
\centering
\input{tilting.tpc}
\caption{$\mathrm{mod}J(Q,W)$ and $\mathrm{mod}J(Q',W')$}
\label{fig:tilting}
\end{figure}
Now, we get two Chern-Simons functionals which realize the moduli stack of objects in $(\mathcal{S}')^\bot={}^\bot \mathcal{S}$ as the critical loci ;
one is the restriction of $f_W$ and the other is the restriction of $f_{\mu_k W}$.
A priori, the virtual motive depends not only on the scheme structure of the critical locus but also on the choice of the Chern-Simons functional.
So we need to show the following :
\begin{align}
&\text{The virtual motives of the moduli stack of objects in $(\mathcal{S}')^\bot={}^\bot \mathcal{S}$ defined}\notag \\
&\text{by $f_{W}$ and $f_{W'}$ coincide.}\label{coincide}
\end{align}
Combine \eqref{coincide} with the arguments above, we can describe the relation between the motivic DT invariants for $(Q,W)$ and for $(Q',W')$ in terms of the quantum dilogarithm.
We prove \eqref{coincide} under the assumption in Theorem \ref{thm_main} (Proposition \ref{prop_44}).
The proof consists of the following two steps :
\smallskip
\noindent (A)
By taking the torus fixed part of the Jacobi algebra $J(Q,W)$ we can define the {\it truncated Jacobi algebra} $J(Q,W)_C$ (\S \ref{subsec_21}) and we have the following identity (Theorem \ref{prop_44}) :
\begin{align}
&\text{The virtual motive of moduli stack of $J(Q,W)$-modules }\notag \\
&=\text{ the motive of the moduli stack of $J(Q,W)_C$-modules.}\label{reduction}
\end{align}
This is a generalization of \cite[Equation (2.4)]{behrend_bryan_szendroi} and \cite[Theorem 9.5]{hua}.
\smallskip
\noindent (B) If $k$ is a {\it strict source} (Definition \ref{defn_strict}), we can take a cut $C'$ of $(Q',W')$ and show an identity between the moduli stack of $J(Q,W)_C$-modules and the one of $J(Q',W')_{C'}$-modules (Proposition \ref{prop_44}).
This makes us possible to compare the virtual motive of the moduli stack of $J(Q,W)$-modules and the one of $J(Q',W')$-modules.
\section*{Comments}
Let us itemize some applications, related topics and further directions.
Some of them will appear in the forthcoming paper.
\begin{itemize}
\item[(a)]
Applying Theorem \ref{thm_FP} for a product of two simply laced Dynkin quivers, we can show a quantized version of dilogarithm identity in conformal field theory \cite{nakanishi_dilog}.
More general identities has been already shown by B. Keller \cite{keller-q-dilog}
\item[(b)]
In \cite{cluster-via-DT}, the author studied cluster algebras by using the ideas in Donaldson-Thomas theory.
It is expected that we can study quantum cluster algebras \cite{quantum_cluster} using motivic Donaldson-Thomas theory.
\item[(c)]
In this paper, the result of Behrend-Bryan-Szendroi \cite[Theorem B.1]{behrend_bryan_szendroi} plays a crucial role and existence of desirable torus action is indispensable.
We want to show the same results for any generic QP in the future.
Once we get \eqref{eq_motivic} and \eqref{coincide}, then we can prove the same results immediately.
\item[(d)]
During preparing this paper, the author was informed by Sergey Mozgovoy of his related work.
In \cite{mozgovoy} he shows a similar result to Theorem \ref{thm_FP} over finite fields.
He uses the result of M. Reineke \cite{reineke_homomorphism}.
\item[(e)]
During preparing this paper, the author was informed also by Balazs Szendroi of his related work.
In \cite{szendroi_morrison}, Szendroi and A. Morrison provide a motivic version of the result of \cite{nagao-nakajima}.
As a result they realize the refined topological vertex of the generating function of motivic invariants, which has already discussed in physics (\cite{dimofte-gukov, dimofte_gukov_soibelman}).
We can apply Theorem \ref{thm_FP} to study the wall-crossing phenomenon. \footnote{The author was informed by Andrew Morrison that he has a generalization for $\mathbb{C}\times \mathbb{C}^2/(\mathbb{Z}/n\mathbb{Z})$. }
\end{itemize}
\section*{Acknowledgement}
{\ }
\vspace{-5mm}
This paper is strongly influenced by ideas in \cite{ks, COHA}.
This paper has also benefited from lectures by Zheng Hua in January 2011 at Nagoya.
He explained to the author his usage of \cite[Theorem B.1]{behrend_bryan_szendroi} in his preprint \cite{hua}.
The author greatly appreciates him.
The author thanks Osamu Iyama and Perre-Guy Plamondon for the useful discussions and comments.
The author also thanks Z. Hua, S. Mozgovoy and B. Szendroi for pointing out some mistakes in the preliminary version.
The author is supported by the Grant-in-Aid for Research Activity Start-up (No. 22840023) and for Scientific Research (S) (No. 22224001).
\section{Motivic Donaldson-Thomas invariants}
\subsection{Motivic ring}
Let $K_0(\mathrm{Var}/\mathbb{C})$ denote
the free abelian group on isomorphism classes of complex varieties,
modulo relations
\[
[X] = [Z] + [U]
\]
for $Z\subset X$ a closed subvariety with complementary open subvariety $U$.
We can equip $K_0(\mathrm{Var}/\mathbb{C})$ with the structure of a commutative
ring by setting
\[
[X]\cdot[Y]=[X\times Y].
\]
We write
\[
\mathbb{L} = [\mathbb{A}^1]\in K_0(\mathrm{Var}/\mathbb{C})
\]
for the class of the affine line.
We define the motivic ring
\[
\mathcal{M}_\mathbb{C}:=K_0(\mathrm{Var}/\mathbb{C})[\mathbb{L}^{-1/2}]
\]
and its localization
\[
\widetilde{\mathcal{M}}_\mathbb{C}:=
\mathcal{M}_\mathbb{C}[(1-\mathbb{L}^n)^{-1}:n\geq 1].
\]
The following lemma is a consequence of \cite[Lemma 3.8]{bridgeland-hall}:
\begin{lem}\label{lem_comporison}
Let $X$ (resp. $Y$) be a variety with an action of a special algebraic group $G$ (resp. $H$).
Assume we have an isomorphism of stacks between $[X/G]$ and $[Y/H]$, then we have
\[
\frac{[X]}{[G]}=\frac{[Y]}{[H]}\in \widetilde{\mathcal{M}}_\mathbb{C}.
\]
\end{lem}
Let $\hat{\mu}:=\lim_{\leftarrow}\mathbb{Z}/n\mathbb{Z}$ be the group of roots of unity.
We define the Grothendieck group $K_0^{\hat{\mu}}(\mathrm{Var}/\mathbb{C})$
of varieties with good $\hat{\mu}$-actions as in \cite[\S 1.4]{behrend_bryan_szendroi}.
We define $\widetilde{\mathcal{M}}^{\hat{\mu}}_\mathbb{C}$ in the same way.
The additive group $\widetilde{\mathcal{M}}^{\hat{\mu}}_\mathbb{C}$
can be endowed with an associative multiplication $\star$
using convolution involving the classes of Fermat curves \cite{denef-loeser-igusa,looijenga}.
This product agrees with the ordinary product on
the subalgebra $\widetilde{\mathcal{M}}_\mathbb{C} \subset \widetilde{\mathcal{M}}^{\hat{\mu}}_\mathbb{C}$ of classes with trivial $\hat{\mu}$-actions, but not in general.
\subsection{Homomorphisms from the motivic ring}
Deligne's mixed Hodge structure on compactly supported cohomology of a variety $X$ gives rise to the $E$-polynomial homomorphism
\[
E
\colon
K_0(\mr{Var}_\mathbb{C})
\to
\mathbb{Z}[x, y]
\]
defined on generators by
\[
E([X];x,y)=
\sum_{p,q}
x^py^q
\sum_i
(-1)^i \dim H_{p,q}(H^i_c(X,\mathbb{Q})).
\]
This extends to a ring homomorphism
\[
E\colon
\widetilde{\mathcal{M}}_\mathbb{C}
\to
\mathbb{Q}(x^{1/2},y^{1/2})
\]
By the specialization
\[
x = y = (xy)^{1/2}= q^{1/2},
\]
we get
\[
\mathbb{W}\colon
\widetilde{\mathcal{M}}_\mathbb{C}
\to
\mathbb{Q}(q^{1/2})
\]
\begin{ex}\label{ex_dilog}
We put
\[
\mca{T}:=\prod_{n\geq 0}\widetilde{\mca{M}}_{\mathbb{C}}\cdot e_n
\]
where $e_n$ is a formal variable satisfying $e_n\cdot e_m=e_{n+m}$.
We put
\[
\sum_{n\geq 0}\frac{[\mr{pt}]}{[\mr{GL}_n]\cdot \mathbb{L}^{-\frac{\dim \mr{GL}_n}{2}}}\cdot{e}_n\in \mca{T}.
\]
We call this \textup{motivic dilogarithm}.
We extend the homomorphism $\mathbb{W}$ to
\[
\mca{T}\to T:=
\prod_{n\geq 0}
\mathbb{Q}(q^{1/2})
\cdot
e_n.
\]
Then the image of the motivic dilogarithm under $\mathbb{W}$ is the \textup{quantum dilogarithm (\cite{quantum-dilogarithm}) :}
\[
\sum_{n\geq 0}\frac{q^{n^2/2}}{(q^n-1)\cdots(q^n-q^{n-1})}{e}_n\in {T}.
\]
\end{ex}
\subsection{Motivic nearby and vanishing cycles}
Let $f\colon X \to \mathbb{C}$ be a regular function on a smooth variety X and let $X_0 := f^{-1}(0)$ be the central fiber.
Using arc spaces, Denef and Loeser \cite{denef-loeser,looijenga} define
the motivic nearby cycle $[\varphi_f]\in \mathcal{M}_\mathbb{C}^{\hat{\mu}}$ of $f$ and
the motivic vanishing cycle
\[
[\varphi_f]:=[\psi_f]-[X_0]\in \mathcal{M}_{\mathbb{C}}^{\hat{\mu}}
\]
of $f$.
Note that if $f=0$, then $[\psi_0]=-[X]$.
\begin{NB}
The following lemma, which is a special case of the motivic Thom-Sebastiani theorem (Theorem \ref{thm_thos_sebastiani}), directly follows the definition :
\begin{lem}
Let $X$ and $Y$ be smooth varieties and $f$ be a regular function on $X$.
Then we have
\[
[\psi_{\pi_X^*f}]=[\psi_{f}\times Y]
\]
where $\pi_X\colon X\times Y \to X$ is the natural projection.
\end{lem}
\end{NB}
\begin{NB}
\begin{lem}\label{lem_13}
Let $G$ be an algebraic group, $H$ be a subgroup of $G$, $X$ be a smooth variety with an action $H$ and $f$ be an $H$-invariant regular function on $X$.
We put $X_G:=X\times_H G$ and $f_G$ be the pull back of $f$ on $X_G$.
Then we have
\[
\frac{[\varphi_{f_G}]}{[G]}=\frac{[\varphi_f]}{[H]}\in
\widetilde{\mathcal{M}}_\mathbb{C}.
\]
\end{lem}
\begin{proof}
We use the description of the motivic nearby cycle in terms of any embedded resolution (\cite{denef-loeser,looijenga}).
We can take an $H$-equivariant embedded resolution $Y\to X$ of $f^{-1}(0)$.
Then $Y_G:=Y\times_H G$ gives a $G$-equivariant embedded resolution of $f_G^{-1}(0)$.
The claim directly follows the descriptions of $[\varphi_{f}]$ and $[\varphi_{f_G}]$ in terms of $Y$ and $Y_G$.
We omit the details, since in the setting of this paper we can prove the claim using Theorem \ref{thm_BBS}.
\end{proof}
\end{NB}
\begin{thm}[\protect{{\bf Motivic Thom-Sebastiani Theorem} \cite{denef-loeser_ts,looijenga}}]\label{thm_thom_sebastiani}
Let $f$, $g$ be regular functions on smooth varieties $X$, $Y$. Then we have
\[
[-\varphi_{f\oplus g}]
=
[-\varphi_{f}]
\star
[-\varphi_{g}]
\]
\end{thm}
\begin{NB}
\begin{cor}\label{lem_12}
Let $X$ and $Y$ be smooth varieties and $f$ be a regular function on $X$.
Then we have
\[
[\psi_{\pi_X^*f}]=[\psi_{f}]\star [Y]
\]
where $\pi_X\colon X\times Y \to X$ is the natural projection.
\end{cor}
\end{NB}
We say that a $\mathbb{C}^*$-action on a variety $X$ is weakly circle compact
if, for all $x\in X$, the limit $\lim_{\lambda \to 0}\lambda\cdot x$ exists.
\begin{thm}[\protect{\cite[Theorem B.1]{behrend_bryan_szendroi}}]\label{thm_BBS}
Let $f\colon X\to \mathbb{C}$ be a regular morphism on a smooth
quasi-projective complex variety.
Assume that there exists an action of a connected complex torus $T$
on $X$ so that $f$ is $T$-equivariant with respect to a primitive character
$\chi\colon T\to \mathbb{C}^*$, namely $f(t\cdot x)=\chi(t)f(x)$ for all $x\in X$ and $t\in T$.
We further assume that there exists a one parameter subgroup $\mathbb{C}^*\subset T$ such that the induced action is weakly circle compact.
Then the motivic nearby cycle class $[\psi_f]$ is in $\mathcal{M}_\mathbb{C}\subset\mathcal{M}_\mathbb{C}^{\hat{\mu}}$ and is equal to $[X_1] = [f^{-1}(1)]$.
Consequently the motivic vanishing cycle class $[\varphi_f]$ is given by
$[\varphi_f] = [f^{-1}(1)] - [f^{-1}(0)]$.
\end{thm}
\begin{rem}
In \textup{\cite[Theorem B.1]{behrend_bryan_szendroi}} they assume that, moreover, the fixed point set $X^{\mathbb{C}^*}$ is compact.
As they themselves mention in \textup{\cite[pp15 l9-10]{behrend_bryan_szendroi}}, this assumption is not necessary.
\end{rem}
\subsection{Virtual motives of critical loci}
Let $f\colon X\to \mathbb{C}$ be a regular function on a smooth variety $X$,
and let $\mathrm{crit}(f) = \{df = 0\}\subset X$ be its degeneracy locus.
\begin{defn}\label{defn_vir}
We define the virtual motive of $\mathrm{crit}(f)$ to be
\[
[\mathrm{crit}(f)]_\mathrm{vir} := -\mathbb{L}^{-\frac{\dim X}{2}}
[\varphi_f]\in \mathcal{M}^\mu_\mathbb{C}.
\]
\end{defn}
\begin{rem}
The virtual motive may depend not only on the scheme structure of the critical locus but also on the presentation as a critical locus.
\end{rem}
We a smooth variety $X$, we use the following notation :
\[
[X]_\mr{vir}:=[\mathrm{crit}(0\colon X\to \mathbb{C})]_\mathrm{vir}
=\mathbb{L}^{-\frac{\dim X}{2}}[X].
\]
\subsection{Motivic Donaldson-Thomas invariants}
Throughout this paper we assume that a quiver is finite and has no loops and oriented $2$-cycles.
Let $Q$ be a quiver,
$Q_0$ denote the set of vertices of $Q$ and
$Q_1$ denote the set of arrows of $Q$.
For an arrow $e\in Q_1$, we denote by $t(e)\in Q_0$ (resp. $h(e)\in Q_0$) the vertex at which $e$ starts (resp. ends).
Take a dimension vector $\mathbf{v}=(v_i)\in (\mathbb{Z}_{\geq 0})^{Q_0}$ and put $V_i=\mathbb{C}^{v_i}$.
We define
\[
\mathrm{M}(Q;\mathbf{v}):=\bigoplus_{e\in {Q_1}}
\mathrm{Hom}(V_{t(e)},V_{h(e)})
\]
and
\[
G(\mathbf{v}):=\prod_{i\in Q_0}\mathrm{GL}(V_i).
\]
Note that $G(\mathbf{v})$ naturally acts on $\mathrm{M}(Q;\mathbf{v})$ and the quotient gives the moduli stack of representations of $Q$ with dimension vectors $\mathbf{v}$.
Let $\chi_Q \colon \mathbb{Z}^{Q_0} \times \mathbb{Z}^{Q_0} \to \mathbb{Z}$ be the bilinear form \footnote{This is the Euler form on the Grothendieck group of the category of finite-dimensional representations
of $Q$.} given by
\[
\chi_Q(\mathbf{v},\mathbf{v}'):=
-\sum_{i,j\in Q_0}Q_{ij}v_iv'_j+
\sum_{i\in Q_0}v_iv'_i.
\]
Then we have
\[
\dim\mathrm{M}(Q;\mathbf{v})
-
\dim G(\mathbf{v})
=
-\chi_Q(\mathbf{v},\mathbf{v}).
\]
Let $W$ be a potential, that is, a finite linear combination of cyclic paths in $Q$.
Let $f_{W,\mathbf{v}}$ be the $G(\mathbf{v})$-invariant function on $\mathrm{M}(Q;\mathbf{v})$ defined by taking the trace of the map associated to the potential $W$.
A point in the critical locus $\mathrm{crit}(f_{W,\mathbf{v}})$ gives a $J(Q,W)$-module and the quotient stack
\[
\bigl[\mathrm{crit}(f_{W,\mathbf{v}})/G(\mathbf{v})\bigr]
\]
gives the moduli stack of $J(Q,W)$-modules with dimension vectors $\mathbf{v}$.
\footnote{In this sense, the function $f_{W,\mathbf{v}}$ is called a Chern-Simons functional.}
\begin{defn}
For $(Q,W)$ and $\mathbf{v}$, we define {\it motivic Donaldson-Thomas invariant} by
\[
\mathfrak{M}_{\mathrm{vir}}(Q,W;\mathbf{v}):=
\frac{[\mathrm{crit}(f_{W,\mathbf{v}})]_\mathrm{vir}}{[G(\mathbf{v})]_\mr{vir}}
\in
\widetilde{\mathcal{M}}^{\hat{\mu}}_\mathbb{C}.
\]
\end{defn}
\section{Cut of a QP and truncated Jacobian}
\subsection{Cut of a QP}\label{subsec_21}
Let $(Q,W)$ be a QP.
To each subset $C\subset Q_1$ we associate a grading $g_C$ on $Q$ by
\[
g_C(a) =
\begin{cases}
1 & a \in C,\\
0 & a \notin C.
\end{cases}
\]
Denote by $Q_C$, the subquiver of $Q$ with the vertex set $Q_0$ and the arrow set $Q_1\backslash C$.
\begin{defn}[\protect{\cite[\S 3]{herschend-iyama}}]\label{defn_cut}
A subset $C\subset Q_1$ is called a \textup{cut} if $W$ is homogeneous of degree $1$ with respect to $g_C$.
\end{defn}
If $C$ is a cut, then $g_C$ induces a grading on $J(Q,W)$ as well.
The degree $0$ part of $J(Q,W)$ is denoted by $J(Q,W)_C$
and called the {\it truncated Jacobian algebra}.
We have
\begin{align*}\label{eq_t_Jacobian}
J(Q,W)_C
&=
J(Q,W)\big/\langle C\rangle\\
&=
\mathbb{C}Q_C\big/\langle \partial_aW\mid a\in C\rangle.
\end{align*}
Here we will show two examples of cuts.
\subsection{Example(1) : bipartite graph and perfect matching}\label{subsec_ex1}
The following example is studied by \cite{Ishii_Ueda2}.
Let $\Sigma$ be a real $2$-dimensional oriented manifold and $\Gamma$ be a bipartite graph on $\Sigma$, that is, $\Gamma$ is a triple $(B,R,E)$ of disjoint finite subsets $B$ (the set of blue vertices) and $R$ (the set of red vertices) of $\Sigma$ and a set of $1$-cells $E$ such that
\begin{itemize}
\item any two elements of $E$ do not intersect in their interiors.
\item each element of $E$ connects one element in $B$ and another element in $R$.
\end{itemize}
We take the dual graph of $\Gamma$.
For each element $e$ in $E$, we define the orientation of the dual edge $\hat{e}$ so that $\hat{e}$ crosses with $e$ keeping the blue boundary of $e$ on the right hand side. Let $Q_\Gamma$ denote the resulting quiver.
For $b\in B$ (resp. $r\in R$), let $w_b$ (resp. $w_r$) be the minimal cyclic path in $Q_\Gamma$ which goes around $b$ (resp. $r$) clockwise (resp. anti-clockwise).
We put
\[
w_\Gamma:=\sum_{b\in B}w_b-\sum_{r\in R}w_r.
\]
\begin{ex}\label{ex1}
Let $\Gamma$ be the bipartite graph on a torus in the left of Figure \ref{fig1}.
The corresponding quiver $Q_\Gamma$ is given in the right of Figure \ref{fig1}.
The potential $W_\Gamma$ is given by
\[
W_\Gamma=a_1b_1c_1d_1-a_1b_2c_1d_2-a_2b_1c_2d_1+a_2b_2c_2d_2.
\]
The quiver with potential is known to be derived equivalent to the quotient stack $[(\mathcal{O}_{\mathbb{P}^1}(-1)\oplus\mathcal{O}_{\mathbb{P}^1}(-1))/(\mathbb{Z}/2\mathbb{Z})]$ where $\mathbb{Z}/2\mathbb{Z}\subset \mr{SL}(\mathbb{C},2)$ acts fiberwise.
\begin{figure}[htbp]
\centering
\input{dimer.tpc}
\caption{an example of a bipartite graph and the quiver}
\label{fig1}
\end{figure}
\end{ex}
A {\it perfect matching} $P$ is a subset of $E$ such that each element $v\in B\cup R$ there exists exactly one element in $P$ which has $v$ as its boundary.
It is easy to check that any perfect matching $P$, as a subset of $(Q_\Gamma)_1=E$, gives a cut of the QP $(Q_\Gamma,W_\Gamma)$.
\begin{ex}
Let $P$ be the perfect matching in Figure \ref{fig2}.
The corresponding cut is $\{a_1,a_2\}$.
\begin{figure}[htbp]
\centering
\input{perfect-matching.tpc}
\caption{an example of a perfect matching}
\label{fig2}
\end{figure}
\end{ex}
\subsection{Example(2) Geometric helices on Del Pezzo surfaces}\label{subsec_ex2}
The following example is studied by \cite{bridgeland-stern}.
Let $Y$ be a Del-Pezzo surface and let $(E_i)_{i=1,...,n}$ be a full exceptional collection on $Y$.
Put $\mathbb{E} = \bigoplus_{i=1}^n E_i$ and define $A(\mathbb{E}) := \mr{End}(\mathbb{E})$.
We put
\[
\mathbb{H}(\mathbb{E})=(E_i)_{i\in \mathbb{Z}}
:=(\ldots, \omega_Y\otimes E_n,E_1,\ldots,E_n,\omega_Y^{-1}\otimes E_1,\ldots).
\]
and assume that
\begin{itemize}
\item $(E_i,\ldots,E_{i+N-1})$ is an exceptional collection on $Y$ for any $i$, and
\item $\mathrm{Hom}^k(E_i,E_j)=0$ for any $k\neq 0$ and any $i<j$.
\end{itemize}
Such a sequence is $(E_i)_{i\in \mathbb{Z}}$ is called a {\it geometric helix} (\cite{bondal-polishchuk}).
The {\it rolled up helix algebra} is the $\mathbb{Z}$-graded algebra
\[
{B}(\mathbb{H}) =
\bigoplus_{k\in\mathbb{Z}}
\mathrm{Hom}(E,\omega_Y^{-k}\otimes E)
\]
with the obvious multiplication.
The following theorem is proved in \cite[\S 6.9]{keller-completion} and \cite[Appendix A]{VdB-example}.
\begin{thm}\label{thm-helix}
There is a QP $(Q,W)$ and a cut $C$ such that
\[
B(\mathbb{H})\simeq J(Q,W)
,\quad
A(\mathbb{E})\simeq J(Q,W)_C.
\]
\end{thm}
\begin{ex}\label{ex3}
We take $Y:=\mathbb{P}_1\times \mathbb{P}_1$
and a geometric helix
\[
\ldots,
\mathcal{O},
\mathcal{O}(1,0),
\mathcal{O}(0,1),
\mathcal{O}(1,1),
\mathcal{O}(2,2)=\omega^{-1}_Y\otimes \mathcal{O},\ldots
\]
where we put
$\mathcal{O}(a,b)=
\pi_1^*(\mathcal{O}_{\mathbb{P}_1}(a))
\otimes
\pi_2^*(\mathcal{O}_{\mathbb{P}_1}(b))$.
We take the quiver $Q$ in Figure \ref{fig3} and the potential
\[
W:= \sum_{i,j\in\{1,2\}}U_{ij}(t_is_j-S_jT_i),
\]
then $(Q,W)$ satisfies the conditions in Theorem \ref{thm-helix}.
\begin{figure}[htbp]
\centering
\input{helix2.tpc}
\caption{quiver for $\mathbb{P}_1\times \mathbb{P}_1$}
\label{fig3}
\end{figure}
Note that $\mathbb{P}_1\times \mathbb{P}_1$ gives a crepant resolution of the quotient singularity $(\mathcal{O}_{\mathbb{P}^1}(-1)\oplus\mathcal{O}_{\mathbb{P}^1}(-1))/(\mathbb{Z}/2\mathbb{Z})$.
In fact, the QP in this example is obtained by mutating the one in Example \ref{ex1} at the vertex $D$ and they are derived equivalent.
\end{ex}
\subsection{Mutation of QP and cut}
Extending Fomin and Zelevinsky mutations of quivers \cite{fomin-zelevinsky1}, Derksen, Weyman, and Zelevinsky have introduced the notion of mutation of QPs in \cite{quiver-with-potentials}.
As pointed out in \cite{graded_mutation}, we can extend the definition to the graded setting.
Let $(Q,W,d)$ be a $\mathbb{Z}$-graded quiver with a homogeneous potential of degree $r$ and $k$ be vertex of $Q$.
We define $\widetilde{\mu}^L_k(Q,W,d) = (\widetilde{Q},\widetilde{W},\widetilde{d})$ the left mutation of $(Q,W,d)$ at vertex $k$ as follows :
\begin{itemize}
\item[(1)]
the new quiver $\widetilde{Q}$ is defined as follows :
\begin{itemize}
\item[(a)]
for any subquiver $u \overset{a}{\longrightarrow} k \overset{b}{\longrightarrow} v$ with $k$, $u$ and $v$
pairwise different vertices, we
add an arrow $[ba]\colon u\to v$;
\item[(b)]
we replace all arrows $a$ incident with $k$ by an arrow $a^*$ in the opposite direction.
\end{itemize}
\item[(2)]
The new potential $\widetilde{W}$ is defined by the sum $[W] + \Delta$ where $[W]$ is
formed from the potential $W$ replacing all compositions $ba$ through the vertex $k$ by the new arrows $[ba]$, and where $\Delta$ is the sum $\sum a^*b^*[ba]$.
\item[(3)]
The new degree $\widetilde{d}$ is defined as follows :
\begin{itemize}
\item[(a)]$\widetilde{d}(a) = d(a)$ for $a$ not incident to $k$ ;
\item[(b)]$\widetilde{d}([ba]) = d(b) + d(a)$ for a composition $ba$ passing through $k$ ;
\item[(c)]$\widetilde{d}(a^*)=-d(a)+r$ if $t(a)=k$;
\item[(d)]$\widetilde{d}(b^*)=-d(b)$ if the source of $s(b)=k$.
\end{itemize}
\end{itemize}
By the graded version (\cite[Theorem 6.4]{graded_mutation}) of the splitting theorem \cite[Theorem 4.6]{quiver-with-potentials}, any graded QP $(Q,W,d)$ has a direct sum decomposition
\[
(Q,W,d)=(Q,W,d)^{\mathrm{red}}\oplus (Q,W,d)^{\mathrm{triv}}
\]
into a reduced graded QP and a trivial graded QP.
The decomposition is unique up to graded right equivalence.
Assume that the potential $W$ is generic in the sense of \cite{quiver-with-potentials}.
Then the underlying quiver of the reduction $\mu^L_k(Q,W,d):=(\widetilde{Q},\widetilde{W},\widetilde{d})^\mr{red}$ of the left mutation $\widetilde{\mu}^L_k(Q,W,d) = (\widetilde{Q},\widetilde{W},\widetilde{d})$ coincides with Fomin-Zelevinsky's mutation.
\begin{defn}[\protect{\cite[Definition 6.12]{herschend-iyama}}]\label{defn_strict}
Let $C$ be a cut of a QP $(Q,W)$.
We say that a vertex $k$ of $Q$ is a strict source (resp. sink) of $(Q,C)$ if all arrows ending (resp. starting) at $x$ belong to $C$ and all arrows starting (ending) at $x$ do not belong to $C$.
\end{defn}
\begin{ex}
\noindent \textup{(1)} Let $(Q,W)$ and $C$ be given as in \S \ref{subsec_ex1}.
A vertex is a strict source or a strict sink if and only if the vertices of the corresponding face of the bipartite graph is perfectly matched by the perfect matching.
\noindent \textup{(2)} Let $(Q,W)$ and $C$ be given as in \S \ref{subsec_ex2}. Then the vertex corresponding to the exceptional object $E_1$ is a strict source.\end{ex}
The underlying graded quiver of $\mu^L_k(Q,W,d_C)$ is given as follows:
\begin{itemize}
\item[(a)]
add degree $1$ arrows $[ba]$ ;
\item[(b)]
replace $a$ with a degree $0$ arrow $a^*$ ;
\item[(c)]
cancel $2$-cycles\footnote{Since we assume $W$ is generic, we can see any $2$-cycle has degree $1$. So this step has no ambiguity even in the graded sense.}.
\end{itemize}
The new degree gives a cut of the mutated QP $\mu_k(Q,W)$.
Let $\mu_kC$ denote this cut.
\begin{rem}
Given a strict source $k$, a new cut $C_k$ of $(Q,W)$ is defined \textup{(\cite[Definition 6.10]{herschend-iyama})} \footnote{They call $C_k$ the {\it cut mutation}.}.
If $(Q,W)$ is Calabi-Yau, then $J(\mu_kQ,\mu_kW)_{\mu_kC}$ is isomorphic to $J(Q,W)_{C_k}$.
\end{rem}
\begin{NB}
\subsection{Category equivalences}
\begin{NB2}
I'm sure that we can replace this subsection with the arguments in \cite[\S 10]{quiver-with-potentials}.
\end{NB2
Let $\Gamma_{Q,W}$ be the Ginzburg dg algebra and
$\mathcal{D}_{Q,W}$ be the derived category of dg modules with finite dimensional cohomologies over $\Gamma_{Q,W}$.
Let
\[
\Phi=\Phi_k \colon \mathcal{D}_{Q,W} \overset{\sim}{\longrightarrow} \mathcal{D}_{Q',W'}
\]
be the derived equivalence (\cite{dong-keller,keller-completion}) which induces the isomorphism $\phi=\phi_k$ of the Grothendieck group given by
\[
\phi([s_i])=
\begin{cases}
[s'_i]+Q(i,k)[s'_k] & i\neq k.\\
-[s'_k] & i=k.
\end{cases}
\]
Assume that $(Q,W)$ is Calabi-Yau, that is, $\Gamma_{Q,W}$ is concentrated on degree $0$.
In such a case
\begin{itemize}
\item[(1)] $\mathcal{D}_{Q,W}\simeq D^b(\mr{mod}J(Q,W))$,
\item[(2)] $\mu_k(Q,W)$ is also Calabi-Yau (\cite{keller-completion}).
\end{itemize}
Let $\Phi$ be the derived equivalence between $D^b(\mr{mod}J(Q,W))$ and $D^b(\mr{mod}J(\mu_kQ,\mu_kW))$. For any $V\in D^b(\mr{mod}J(Q,W))$, we have
\[
H^j(\Phi(V))_i=0\ (j\neq 0),\quad H^0(\Phi(V))_i=V_i
\]
for $i\neq k$ and the map
\[
[ba] \colon H^0(\Phi(V))_{s(a)} \to H^0(\Phi(V))_{t(b)}
\]
is given by the composition $b\circ a$.
We define the following full subcategories
\begin{align*}
\mr{mod}(J(Q,W))_k
&:=
\{V\in \mr{mod}(J(Q,W))\mid \Hom(s_k,V)=0\},\\
\mr{mod}(J(Q,W)_C)_k
&:=
\mr{mod}(J(Q,W))_k
\cap
\mr{mod}(J(Q,W)_C),\\
\mr{mod}(J(Q',W'))^k
&:=
\{V\in \mr{mod}(J(Q',W'))\mid \Hom(V,s'_k)=0\},\\
\mr{mod}(J(Q',W')_{C'})^k
&:=
\mr{mod}(J(Q',W'))^k
\cap
\mr{mod}(J(Q',W')_{C'}).
\end{align*}
Note that for $V\in \mr{mod}(J(Q,W))$ the condition $\Hom(s_k,V)=0$ is equivalent to the injectiveness of
\[
\sum_{s(b)=k}V_k\to \biggl(\,\bigoplus_{s(b)=k} V_{t(b)}\biggr)
\]
and for $V\in \mr{mod}(J(Q',W'))$ the condition $\Hom(V,s'_k)=0$ is equivalent to the surjectiveness of
\[
\sum_{t(a)=k}\biggl(\,\bigoplus_{t(a)=k} V_{s(a)}\biggr)\to V_k.
\]
It is shown in \cite{keller-completion}, the derived equivalence $\Phi$ induces
\[
\mr{mod}(J(Q,W))_k
\simeq
\mr{mod}(J(Q',W'))^k.
\]
\begin{prop}\label{prop_tilting}
The derived equivalence induces
\[
\mr{mod}(J(Q,W)_C)_k
\simeq
\mr{mod}(J(Q',W')_{C'})^k.
\]
\end{prop}
\begin{proof}
First, take $E\in \mr{mod}(J(Q,W)_C)_k$.
For a composition $ba$ through $k$, $[bc]$ vanishes on $\Phi(E)$ since $a$ vanishes on $E$.
Hence we have $\Phi(E)\in \mr{mod}(J(Q',W')_{C'})^k$.
Next, assume that $\Phi(E)\in \mr{mod}(J(Q',W')_{C'})^k$ for $E\in \mr{mod}(J(Q,W))_k$.
For any $a$ with $t(a)=k$, $a$ vanishes on $E$ since
$\sum_{s(b)=k}[ba]$ vanished and $\sum_{s(b)=k}b$ is injective.
Hence we have $E\in \mr{mod}(J(Q,W)_{C})_k$.
\end{proof}
\begin{rem}
This gives a generalization of a part of the results of \cite{bridgeland-stern}.
\end{rem}
\end{NB}
\section{Factorization property}\label{sec_FP}
Let $(Q,W)$ be a QP and $C$ be a cut.
The grading $g_C$ gives a $\mathbb{C}^*$-action on $\mathrm{M}(Q;\mathbf{v})$ and the action satisfies the assumption of Theorem \ref{thm_BBS} if we put
\[
X=\mathrm{M}(Q;\mathbf{v}), \quad
T=\mathbb{C}^*,\quad
f=f_{W,\mathbf{v}}.
\]
Hence we have
\begin{equation}\label{eq_no_monodoromy}
\mathfrak{M}_{\mathrm{vir}}(Q,W;\mathbf{v})
\in
\widetilde{\mathcal{M}}_\mathbb{C}
\end{equation}
and
\begin{align*}
\mathfrak{M}_{\mathrm{vir}}(Q,W;\mathbf{v})
&=
\frac{\mathbb{L}^{-\dim \mathrm{M}(Q;\mathbf{v})/2}\cdot (f_{W,\mathbf{v}}^{-1}(1)-f_{W,\mathbf{v}}^{-1}(0))}
{[G(\mathbf{v})]_\mr{vir}}\\
&=
\mathbb{L}^{\chi_Q(\mathbf{v},\mathbf{v})/2}\times \frac{f_{W,\mathbf{v}}^{-1}(1)-f_{W,\mathbf{v}}^{-1}(0)}
{[G(\mathbf{v})]}
.
\end{align*}
We define the {\it refined DT invariant} by
\[
m_{\mr{ref}}(Q,W;\mathbf{v})
:=
\mathbb{W}\bigl(
\mathfrak{M}_{\mathrm{vir}}(Q,W;\mathbf{v})
\bigr)
\in \mathbb{Q}(q^{1/2}).
\]
Throughout this section, we will use simplified notations such as
$\mathrm{M}(\mathbf{v})$ and $f_{\mathbf{v}}$
instead of
$\mathrm{M}(Q;\mathbf{v})$ and $f_{W,\mathbf{v}}$
omitting $Q$ and $W$.
\subsection{Filtration and motivic invariants}
\begin{NB}
Let $X$, $T$, $f$ be as in Theorem \ref{thm_BBS} and
\[
0=U_0\subset U_1\subset \cdots \subset U_n =X
\]
be a filtration of $X$ by $T$-invariant open subsets.
We put $X_\alpha:=U_\alpha\backslash U_{\alpha-1}$ and $f_\alpha:=f|_{X_\alpha}$.
Assume that $X_\alpha$, $f_\alpha$ and the $T$-action satisfies the conditions in Theorem \ref{thm_BBS} too.
The next proposition directly follows Theorem \ref{thm_BBS} and Definition \ref{defn_vir}.
\end{NB}
The next proposition directly follows Theorem \ref{thm_BBS} and Definition \ref{defn_vir}.
\begin{prop}\label{prop_31}
Let $X$, $T$, $f$ be as in Theorem \ref{thm_BBS} and $Y$ be a smooth $T$-invariant closed subset $X$ with dimension $d$.
We assume that the $T$-action on $Y$ and $f|_Y$ satisfies the conditions in Theorem \ref{thm_BBS} as well.
Then we have
\[
[\mathrm{crit}(f)]_\mathrm{vir}
=
[\mathrm{crit}(f|_{X\backslash Y})]_\mathrm{vir}
+
\mathbb{L}^{-\frac{d}{2}}
[\mathrm{crit}(f|_{Y})]_\mathrm{vir}.
\]
\end{prop}
\begin{cor}
Let
\[
0=U_0\subset U_1\subset \cdots \subset U_n =X
\]
be a filtration of $X$ by $T$-invariant open subsets.
We put $X_\alpha:=U_\alpha\backslash U_{\alpha-1}$ and $f_\alpha:=f|_{X_\alpha}$.
Assume that $X_\alpha$, $f_\alpha$ and the $T$-action satisfies the conditions in Theorem \ref{thm_BBS}. Then we have
\[
[\mathrm{crit}(f)]_\mathrm{vir}
=
\sum_{\alpha=1}^n \mathbb{L}^{\frac{-\mr{dim}X+\mr{dim}X_\alpha}{2}}
[\mathrm{crit}(f_\alpha)]_\mathrm{vir}.
\]
\end{cor}
\subsection{Filtration by HN property}\label{subsec_ks52}
In this subsection, we repeat \cite[\S 5.2]{COHA} to fix the notations.
We put
\[
\mathbb{H}:=\{z\in \mathbb{C}\mid \mathrm{im} z >0\text{ or } z\in \mathbb{R}_{>0}\}
\]
and define the total order $\succ$ on $\mathbb{H}$ by
\[
z_1 \succ z_2
\overset{\mathrm{def}}{\iff}
\mathrm{Arg}(z_1) > \mathrm{Arg}(z_2)
\text{ or }
\bigl\{ \mathrm{Arg}(z_1) = \mathrm{Arg}(z_2), |z_1|>|z_2|\bigr\}.
\]
We identify the Grothendieck group $K_0(\mathrm{mod}(J(Q,W)))$ with $\mathbb{Z}^{Q_0}$ and put $N:=(\mathbb{Z}_{\geq 0})^{Q_0}$.
Let
\[
Z\colon \mathbb{Z}^{Q_0}\to \mathbb{C}
\]
be a central charge, that is, a group homomorphism such that
\[
Z(N\backslash\{{0}\})\subset\mathbb{H}.
\]
Let $\mathrm{M}_{Z\text{-ss}}(\mathbf{v})$ denote the open subset of $\mathrm{M}(\mathbf{v})$
consisting of $Z$-semistable $Q$-modules.
\begin{defn}
For $\mathbf{v}\in N$,
we define the finite set $P_Z(\mathbf{v})$ by
\[
\Bigl\{
\mathbf{v}_\bullet=(\mathbf{v}_i)\in N^n\,\Big |\, n\geq 1,\ \sum \mathbf{v}_i=\mathbf{v},\
\mathrm{Arg}Z(\mathbf{v}_1)>\cdots >\mathrm{Arg}Z(\mathbf{v}_n)
\Bigr\}.
\]
We introduce a partial order on $P_Z(\mathbf{v})$ by
\begin{align*}
(\mathbf{v}_1,\ldots,\mathbf{v}_n)
&\underset{{\small Z}}{<}
(\mathbf{v}'_1,\ldots,\mathbf{v}'_{n'})\\
\overset{\mathrm{def}}{\iff}
&
\mathbf{v}_1=\mathbf{v}'_1,\ldots, \mathbf{v}_{i-1}=\mathbf{v}'_{i-1} \text{ and } Z(\mathbf{v}_i)\prec Z(\mathbf{v}'_{i'})\text{ for some $i$}.
\end{align*}
\end{defn}
\begin{defn}
Let denote by
$\mathrm{M}(\mathbf{v};\mathbf{v}_\bullet)$ the subset of $\mathrm{M}(\mathbf{v})$ consisting of $Q$-modules which
admit increasing filtrations
\[
0 = E_0\subset E_1 \subset \cdots \subset E_n=E
\]
such that
\[
\underline{\mathrm{dim}}(E_i/E_{i-1})=\mathbf{v}_i
\]
for any $i=1,\ldots,n$.
\end{defn}
\begin{lem}
The subset $\mathrm{M}(\mathbf{v};\mathbf{v}_\bullet)\subset \mathrm{M}(\mathbf{v})$ is closed.
\end{lem}
\begin{proof}
We put
\[
\mathrm{Fl}(\mathbf{v}_\bullet)
:=
\prod_{i\in Q_0}
\mathrm{Fl}(v_{1,i},\ldots,v_{n,i})
\]
where $\mathrm{Fl}(v_{1,i},\ldots,v_{n,i})$ is the flag varieties of all flags in $V_i$ of type $(v_{1,i},\ldots,v_{n,i})$.
Note that the following subset of $\mathrm{M}(\mathbf{v})\times \mathrm{Fl}(\mathbf{v}_\bullet)$ is closed :
\[
\bigl\{(X,F)\in \mathrm{M}(\mathbf{v})\times \mathrm{Fl}(\mathbf{v}_\bullet)\,\big|\,\text{$F$ is $X$-stable}\bigr\}.
\]
Then $\mathrm{M}(\mathbf{v};\mathbf{v}_\bullet)$ is closed since it is the image of the closed set above under the projection
\[
\mathrm{M}(\mathbf{v})\times \mathrm{Fl}(\mathbf{v}_\bullet)\to
\mathrm{M}(\mathbf{v})
\]
which is proper.
\end{proof}
\begin{defn}
We define the locally closed subset $\mathrm{M}_{Z\textup{-HN}}(\mathbf{v},\mathbf{v}_\bullet)$ of
$\mathrm{M}(\mathbf{v})$ by
\[
\mathrm{M}_{Z\textup{-HN}}(\mathbf{v},\mathbf{v}_\bullet)
:=
\mathrm{M}(\mathbf{v};\mathbf{v}_\bullet)
-
\bigcup_{\mathbf{v}'_\bullet\underset{{\small Z}}{<}\mathbf{v}_\bullet}
\mathrm{M}(\mathbf{v};\mathbf{v}'_\bullet).
\]
\end{defn}
\begin{lem}\label{lem_36}
\begin{itemize}
\item[\textup{(1)}]
A $\mathbb{C}$-point in $\mathrm{M}_{Z\textup{-HN}}(\mathbf{v},\mathbf{v}_\bullet)$ represents a $Q$-module whose HN filtration is of type $\mathbf{v}_\bullet$.
\item[\textup{(2)}]
$\mathrm{M}_{Z\textup{-HN}}(\mathbf{v},\mathbf{v}_\bullet)$ is smooth and
\[
\mathrm{codim}\,\mathrm{M}_{Z\textup{-HN}}(\mathbf{v},\mathbf{v}_\bullet)
=
-\sum_{a<b}\chi_Q(\mathbf{v}_a,\mathbf{v}_b).
\]
\end{itemize}
\end{lem}
\begin{proof}
(1) See \cite[pp49, Lemma 2]{COHA}.
\noindent (2)
Fix a direct sum decompositions $V_i=\oplus_{a=1}^n V_{a,i}$ with $V_{a,i}\simeq \mathbb{C}^{v_{a,i}}$.
We define the subspace $\mathrm{M}(\mathbf{v}_\bullet)$ of $\mathrm{M}(\mathbf{v})$ by
\[
\mathrm{M}(\mathbf{v}_\bullet):=\bigoplus_{a\geq b,\;e\in {Q_1}}
\mathrm{Hom}(V_{a,t(e)},V_{b,h(e)}).
\]
We put
\[
\mathrm{M}_{Z\text{-ss}}(\mathbf{v}_\bullet)
:=
\pi^{-1}
\bigl(
\mathrm{M}_{Z\text{-ss}}(\mathbf{v}_1)
\times\cdots\times \mathrm{M}_{Z\text{-ss}}(\mathbf{v}_n)
\bigr)
\]
where
\[
\pi \colon \mathrm{M}(\mathbf{v}_\bullet)\to
\mathrm{M}(\mathbf{v}_1)
\times\cdots\times
\mathrm{M}(\mathbf{v}_n)
\]
is the natural projection.
Note that $\pi$ is a trivial vector bundle
and so $\mathrm{M}_{Z\text{-ss}}(\mathbf{v}_\bullet)$ is smooth.
\begin{NB}
We can see that
\[
\mathrm{M}_{Z\text{-ss}}(Q;\mathbf{v}_\bullet)\times_{G(\mathbf{v}_\bullet)}G(\mathbf{v})
\simeq
\mathrm{M}_{Z\text{-HN}}(Q;\mathbf{v},\mathbf{v}_\bullet)
\]
and so $\mathrm{M}_{Z\text{-HN}}(Q;\mathbf{v},\mathbf{v}_\bullet)$ is smooth.
\end{NB}
Let
\[
\mr{HN}\colon
\mathrm{M}_{Z\text{-HN}}(\mathbf{v},\mathbf{v}_\bullet)
\to
\mr{FL}(\mathbf{v}_\bullet)
\]
be the map defined by taking the Harder-Narashimhan filtration.
Then $\mr{HN}$ is a Zariski locally trivial fibration whose fibers are isomorphic to $\mathrm{M}_{Z\text{-ss}}(Q;\mathbf{v}_\bullet)$.
So $\mathrm{M}_{Z\text{-HN}}(Q;\mathbf{v},\mathbf{v}_\bullet)$ is smooth.
The computation of the codimension is straightforward.
\end{proof}
\begin{NB}
\[
\mathrm{codim}\,\mathrm{M}^{\mathrm{HN}}(\mathbf{v},\mathbf{v}_\bullet)
=
-\sum_{a<b}\chi_Q(\mathbf{v}_a,\mathbf{v}_b).
\]
A $\mathbb{C}$-valued point of the subset
\[
\mathrm{M}_{Z\text{-ss}}(Q;\mathbf{v}_\bullet)\times_{G(\mathbf{v}_\bullet)}G(\mathbf{v})
\subset
\mathrm{M}_{Z\text{-HN}}(Q;\mathbf{v},\mathbf{v}_\bullet)
\]
represents a $Q$-module whose HN filtration is of type $\mathbf{v}_\bullet$.
Let $\mathrm{M}_{Z\text{-ss}}(Q;\mathbf{v}_\bullet)$ denote this smooth subvariety.
\begin{defn}
\begin{itemize}
\item[(1)]
\item[(2)]
\item[(3)]
We denote by
$\mathrm{M}(\mathbf{v};\mathbf{v}_\bullet)$ the subset of $\mathrm{M}(\mathbf{v})$ whose $\mathbb{C}$-points are objects $E\in \mathrm{M}(\mathbf{v};\mathbf{v}_\bullet)$ which
admit an increasing filtration
\[
0 = E_0\subset E_1 \subset \cdots \subset E_n
\]
such that
\[
\underline{\mathrm{dim}}(E_i/E_{i-1})=\mathbf{v}_i
\]
for any $i=1,\ldots,n$.
\end{itemize}
\end{defn}
\begin{lem}\label{lem_HNlocus}
\begin{itemize}\textup{(}See \textup{\cite[pp49, Lemma 2 and pp51]{COHA}}.\textup{)}
\item[\textup{(1)}] The subset $\mathrm{M}(\mathbf{v};\mathbf{v}_\bullet)\subset \mathrm{M}(\mathbf{v})$ is closed.
\item[\textup{(2)}] For any $\mathbf{v}\in C$ and $\mathbf{v}_\bullet\in P(\mathbf{v})$, the set of $\mathbb{C}$-points of the locally closed subset
\[
\mathrm{M}(\mathbf{v},\mathbf{v}_\bullet)-
\bigcup_{\mathbf{v}'_\bullet<\mathbf{v}_\bullet}\mathrm{M}(\mathbf{v},\mathbf{v}'_\bullet)
\]
is the set of representations whose HN-filtrations are of type $\mathbf{v}_\bullet$.
We denote the locally closed subset by $\mathrm{M}^{\mathrm{HN}}(\mathbf{v},\mathbf{v}_\bullet)$.
\item[\textup{(3)}]
The locally closed subset $\mathrm{M}^{\mathrm{HN}}(\mathbf{v},\mathbf{v}_\bullet)\subset \mathrm{M}(\mathbf{v})$ is smooth and its codimension is given by
\[
\mathrm{codim}\,\mathrm{M}^{\mathrm{HN}}(\mathbf{v},\mathbf{v}_\bullet)
=
-\sum_{a<b}\chi_Q(\mathbf{v}_a,\mathbf{v}_b).
\]
\end{itemize}
\end{lem}
\end{NB}
Let $f^{Z\textup{-ss}}_{\mathbf{v}}$ (resp. $f^{Z\textup{-ss}}_{\mathbf{v}_\bullet}$, $f^{Z\textup{-HN}}_{\mathbf{v}_\bullet}$) denote the restriction of the Chern-Simons functional $f_{\mathbf{v}}=f_{W,\mathbf{v}}$ on $\mathrm{M}_{Z\textup{-ss}}(\mathbf{v})$ (resp. $\mathrm{M}_{Z\textup{-ss}}(\mathbf{v}_\bullet)$, $\mathrm{M}_{Z\textup{-HN}}(\mathbf{v},\mathbf{v}_\bullet)$).
\begin{prop}[\protect{see \cite[pp51 Theorem 5]{COHA}}]\label{prop_factorization}
Assume that the QP has a cut, then we have
\[
\frac{\bigl[\mr{crit}\bigl(f^{Z\textup{-HN}}_{\mathbf{v}_\bullet}\bigr)\bigr]_{\mr{vir}}}{[\mr{G}(\mathbf{v})]_\mr{vir}}
=
\mathbb{L}^{-\sum_{a>b}\chi_Q(\mathbf{v}_a,\mathbf{v}_b)}
\times \prod_{a}
\frac{\bigl[\mr{crit}\bigl(f^{Z\textup{-ss}}_{\mathbf{v}_a}\bigr)\bigr]_{\mr{vir}}}{[\mr{G}(\mathbf{v}_a)]_\mr{vir}}.
\]
\end{prop}
\begin{proof}
The claim is a consequence of the following two identities, which is obtained from the descriptions of in the proof of Lemma \ref{lem_36} :
\begin{align*}
\bigl[\mr{crit}\bigl(f^{Z\textup{-HN}}_{\mathbf{v}_\bullet}\bigr)\bigr]_{\mr{vir}}
&=
\bigl[\mr{crit}\bigl(f^{Z\textup{-ss}}_{\mathbf{v}_\bullet}\bigr)\bigr]_{\mr{vir}}
\times
[\mr{FL}(\mathbf{v}_\bullet)]_\mr{vir}\\
&= \bigl[\mr{crit}\bigl(f^{Z\textup{-ss}}_{\mathbf{v}_\bullet}\bigr)\bigr]_{\mr{vir}}
\times
\frac{[\mr{G}(\mathbf{v})]_\mr{vir}}{\prod_a [\mr{G}(\mathbf{v_a})]_\mr{vir}\times \bigl[\mathbb{L}^{\sum_{i,a>b}v_{a,i}\cdot v_{b,i}}\bigr]_\mr{vir}}
\end{align*}
and
\begin{align*}
\bigl[\mr{crit}\bigl(f^{Z\textup{-ss}}_{\mathbf{v}_\bullet}\bigr)\bigr]_{\mr{vir}}
&=
\Bigl[\mathbb{L}^{\sum_{i,a>b} Q_{ij}v_{a,i}\cdot v_{b,i}}\Bigr]_\mr{vir}\times \prod_a
\bigl[\mr{crit}\bigl(f^{Z\textup{-ss}}_{\mathbf{v}_a}\bigr)\bigr]_{\mr{vir}}.
\end{align*}
For the second identity, we use the motivic Thom-Sebastiani theorem (Theorem \ref{thm_thom_sebastiani}).
\end{proof}
Let $\langle\bullet,\bullet\rangle \colon \mathbb{Z}^{Q_0} \times \mathbb{Z}^{Q_0} \to \mathbb{Z}$ be the skewsymmetric bilinear form given by
\[
\langle\mathbf{v},\mathbf{v}'\rangle:=
\chi_Q(\mathbf{v},\mathbf{v}')-\chi_Q(\mathbf{v}',\mathbf{v}).
\]
Combining the results in this subsection, we get the following theorem :
\begin{thm}\label{thm_39}
Assume that the QP has a cut, then we have
\[
\mathfrak{M}_{\mr{vir}}(Q,W,\mathbf{v})
:=
\sum_{\mathbf{v}_\bullet\in P_Z(\mathbf{v})}
\biggl(\mathbb{L}^{\frac{1}{2}\sum_{a<b} \langle\mathbf{v}_a,\mathbf{v}_b\rangle}
\times
\prod_a \mathfrak{M}_{\mr{vir}}(Q,W,\mathbf{v}_a)\biggr)
\]
\end{thm}
\subsection{Factorization property}
We assume that the QP has a cut.
\begin{defn}\label{defn_mt}
The {\it motivic torus} associated to $Q$ is
\[
\hat{\mathcal{T}}_Q:=
\prod_{\mathbf{v}\in N} \widetilde{\mathcal{M}}_\mathbb{C}\cdot y_\mathbf{v}
\]
where $y_\mathbf{v}$'s are formal variables which satisfy the relation
\[
y_{\mathbf{v}_1}\cdot y_{\mathbf{v}_2}
=
\mathbb{L}^{\frac{\langle\mathbf{v},\mathbf{v}'\rangle}{2}}y_{\mathbf{v}_1+\mathbf{v}_2}.
\]
\end{defn}
\begin{defn}
We define the generating series of the motivic Donaldson-Thomas invariants of $(Q,W)$ by
\[
\mca{A}=\mca{A}_{Q,W}:=
1+\sum_{\mathbf{v}\in N}
\mathfrak{M}_{\mr{vir}}(Q,W,\mathbf{v})\cdot y_\mathbf{v}
\in
\hat{\mathcal{T}}_Q.
\]
\end{defn}
\begin{defn}
Let $l\subset \mathbb{H}$ be a ray and $Z$ be a central charge.
We put
\[
\mca{A}^{Z,l}:=
1+\sum_{Z(\mathbf{v})\in l}
\frac{[\mr{crit}(f^{Z\textup{-ss}}_{\mathbf{v}})]_{\mr{vir}}}{[G(\mathbf{v})]}\cdot y_\mathbf{v}
\in
\hat{\mathcal{T}}_Q.
\]
\end{defn}
Theorem \ref{thm_39} implies the following {\it factorization formula} :
\begin{thm}\label{thm_FP}
Assume that the QP has a cut, then we have
\[
\mca{A}=\prod_{l}^{\curvearrowright}\mca{A}^{Z,l}
\]
where the product is taken in the clockwise order over all rays.
\end{thm}
\begin{defn}\label{defn_qt}
The {\it quantum torus associated} to $Q$ is
\[
\hat{{T}}_Q:=
\prod_{\mathbf{v}\in N} \mathbb{Q}(q^{1/2})\cdot \mr{y}_\mathbf{v}
\]
where $\mr{y}_\mathbf{v}$'s are formal variables which satisfy the relation
\[
\mr{y}_{\mathbf{v}_1}\cdot \mr{y}_{\mathbf{v}_2}
=
q^{\frac{\langle\mathbf{v},\mathbf{v}'\rangle}{2}}\mathrm{y}_{\mathbf{v}_1+\mathbf{v}_2}.
\]
\end{defn}
\begin{defn}\label{defn_rdt}
We define the generating series of the refined Donaldson-Thomas invariants of $(Q,W)$ by
\[
{A}={A}_{Q,W}:=
m_{\mr{ref}}(Q,W,\mathbf{v})\cdot \mr{y}_\mathbf{v}
\in
\hat{{T}}_Q.
\]
\end{defn}
In the same way, we define ${A}^{Z,l}$.
Since $\mathbb{W}$ is a ring homomorphism, we get the following
factorization formula for refined DT invariants :
\begin{cor}\label{cor_FP}
Assume that the QP has a cut, then we have
\[
{A}=\prod_{l}^{\curvearrowright}{A}^{Z,l}
\]
where the product is taken in the clockwise order over all rays.
\end{cor}
\section{Wall-crossing formula}\label{sec_WC}
\subsection{Motives for $J(Q,W)_C$}
We put
\begin{align*}
\chi_C(\mathbf{v},\mathbf{v'})&:=\sum_{c\in C}v_{t(c)}v'_{h(c)},\\
\chi_{Q_C}(\mathbf{v},\mathbf{v'})&:=
\chi_Q(\mathbf{v},\mathbf{v'})-\chi_C(\mathbf{v},\mathbf{v'}).
\end{align*}
We put
\[
d:=\dim \mr{M}(Q;\mathbf{v}) - \dim \mr{M}(Q_C;\mathbf{v})
=\chi_C(\mathbf{v},\mathbf{v}).
\]
Let $\mathrm{M}(Q,W,C;\mathbf{v})$ be the subset of $\mathrm{M}(Q_C;\mathbf{v})$ consisting of $J(Q,W)_C$-modules.
The following theorem is a generalization of \cite[Equation (2.4)]{behrend_bryan_szendroi} and \cite[Theorem 7.3]{hua}.
\begin{thm}\label{thm_cut_id}
\[
[\varphi_{f_{W,\mathbf{v}}}]
=
\mathbb{L}^d\cdot [\mathrm{M}(Q,W,C;\mathbf{v})].
\]
\end{thm}
\begin{proof}
Note that we have
\begin{align}
[\varphi_{f_{W,\mathbf{v}}}]
&=
f_{W,\mathbf{v}}^{-1}(1)
-
f_{W,\mathbf{v}}^{-1}(0)\notag\\
&=
\frac{[\mr{M}(Q;\mathbf{v})]-f_{W,\mathbf{v}}^{-1}(0)}{\mathbb{L}-1}
-
f_{W,\mathbf{v}}^{-1}(0)\notag\\
&=
\frac{[\mr{M}(Q;\mathbf{v})]-\mathbb{L}\cdot f_{W,\mathbf{v}}^{-1}(0)}{\mathbb{L}-1}.\label{eq_1}
\end{align}
Let $\pi\colon \mr{M}(Q;\mathbf{v})\to \mr{M}(Q_C;\mathbf{v})$ be the natural projection.
This is a trivial vector bundle of rank $d$.
Since we have
\[
W=\sum_{e\in C}\partial_eW,
\]
the restriction of $f_{W,\mathbf{v}}$ to the fiber $\pi^{-1}(x)$ is zero if $x\in \mathrm{M}(Q,W,C;\mathbf{v})$ and is a non-zero linear function if $x\notin \mathrm{M}(Q,W,C;\mathbf{v})$.
Hence we have
\begin{equation}\label{eq_2}
f_{W,\mathbf{v}}^{-1}(0)=
\mathbb{L}^d\cdot [\mathrm{M}(Q,W,C;\mathbf{v})]
+
\mathbb{L}^{d-1}\bigl([\mathrm{M}(Q_C;\mathbf{v})]-[\mathrm{M}(Q,W,C;\mathbf{v})]\bigr).
\end{equation}
Substitute \eqref{eq_2} to \eqref{eq_1}, then the claim follows.
\end{proof}
\begin{rem}
For the cohomological Hall algebra, a similar statement is proved in \cite[Proposition 6]{COHA}.
\end{rem}
\subsection{Mutation and bilinear forms}
Let $(Q',W',C')$ be the new QP with the cut given by the mutation at a strict source $k$ of $(Q,W,C)$.
We identify the Grothendieck group $K_0(\mathrm{mod}(J(Q,W)))$ with $\mathbb{Z}^{Q_0}$ as before.
We define
\[
\phi_k\colon
\mathbb{Z}^{Q_0} \overset{\sim}{\longrightarrow} \mathbb{Z}^{Q'_0}
\]
by
\[
\phi_k([s_i])=
\begin{cases}
[s'_i] & i\neq k,\\
-[s'_k]+\sum_{t(b)=k}[s'_{h(b)}] & i=k.
\end{cases}
\]
Then we can verify the following :
\begin{lem}\label{lem_43}
\begin{align*}
\chi_Q(\mathbf{v},\mathbf{w})
&=
\chi_{Q'}(\phi_k(\mathbf{v}),\phi_k(\mathbf{w})),\\
\chi_C(\mathbf{v},\mathbf{w})
&=
\chi_{C'}(\phi_k(\mathbf{v}),\phi_k(\mathbf{w})),\\
\chi_{Q_C}(\mathbf{v},\mathbf{w})
&=
\chi_{Q_{C'}}(\phi_k(\mathbf{v}),\phi_k(\mathbf{w})).
\end{align*}
\end{lem}
\begin{rem}
The bilinear form $\chi_Q$ is the Euler form of the derived category of the Ginzburg's dg algebra.
The map $\phi_k$ is induced from the Keller-Yang's derived equivalence (\cite{dong-keller,keller-completion}).
This is the origin of the first equation.
If the Ginzburg's dg algebra is concentrated on degree $0$, then the bilinear form $\chi_{Q_C}$ is the Euler form of the derived category of $J(Q,W)_C$. In such a cases, we have the derived equivalence between $J(Q,W)_C$ and $J(Q',W')_{C'}$ which induces $\phi_k$. This is the origin of the third equation.
\end{rem}
\subsection{Mutation and a motivic identity}
For a strict source $k$, we define open subsets
$\mathrm{M}(Q;\mathbf{v})_k\subset \mathrm{M}(Q;\mathbf{v})$ and
$\mathrm{M}(Q_C;\mathbf{v})_k\subset \mathrm{M}(Q_C;\mathbf{v})$ by
\begin{align*}
\mathrm{M}(Q;\mathbf{v})_k&:=\{ V\in \mathrm{M}(Q;\mathbf{v})\mid \Hom(s_k,V)=0 \},\\
\mathrm{M}(Q_C;\mathbf{v})_k&:=\{ V\in \mathrm{M}(Q_C;\mathbf{v})\mid \Hom(s_k,V)=0 \}
\end{align*}
and put
\[
\mathrm{M}(Q,W,C;\mathbf{v})_k
:=
\mathrm{M}(Q_C;\mathbf{v})_k
\cap
\mathrm{M}(Q,W,C;\mathbf{v}).
\]
Note that we have $\pi^{-1}(\mathrm{M}(Q_C;\mathbf{v})_k)=\mathrm{M}(Q;\mathbf{v})_k$.
We put
\[
f_{W,\mathbf{v},k}
:=
f_{W,\mathbf{v}}|_{M(Q,W,C,\mathbf{v})_k}.
\]
We can prove the following in the same way as Theorem \ref{thm_cut_id} :
\begin{prop}\label{prop_43}
\[
[\varphi_{f_{W,\mathbf{v},k}}]
=
\mathbb{L}^{d}\cdot [\mathrm{M}(Q,W,C;\mathbf{v})_k].
\]
\end{prop}
We use the upper subscript $M(\cdots){}^k$ for the ones with the condition $\Hom(-,s_k')=0$.
\begin{prop}\label{prop_44}
\begin{equation}
\frac{[\mr{M}(Q,W,C;\mathbf{v})_k]}{[G(\mathbf{v})]}
=
\frac{[\mr{M}(Q',W',C';\mathbf{v}')^k]}{[G(\mathbf{v}')]}
\end{equation}
where $\mathbf{v}'=\phi_k(\mathbf{v})$.
\end{prop}
\begin{proof}
This is a consequence of Proposition \ref{prop_tilting} and Lemma \ref{lem_comporison}.
\footnote{In fact, the equivalence in Corollary \ref{prop_tilting} gives only the bijection of $\mathbb{C}$-valued points. But all the arguments in \S \ref{subsec_appendix} can be applied for families of representations and we can see the isomorphism of the moduli stacks.}
\end{proof}
Combining this with Proposition \ref{prop_43} and Lemma \ref{lem_43}, we get the following identity of the virtual motives of the same moduli stack with different Chern-Simons functionals :
\begin{thm}\label{thm_45}
\[
\mathfrak{M}_{\mr{vir}}(Q,W;\mathbf{v})_k
=
\mathfrak{M}_{\mr{vir}}(Q',W';\mathbf{v}')^k.
\]
\end{thm}
\subsection{Wall-crossing formula for motivic DT invariants}
In this subsection, we will work over $\mathbb{Q}$ and use the notations as $\hat{\mathcal{T}}^{\mathbb{Q}}_{Q}:=\hat{\mathcal{T}}_{Q}\otimes \mathbb{Q}$.
We put $N_{Q,Q'}:=\phi_k^{-1}(N)\cap N$ and
\[
\hat{\mathcal{T}}^{\mathbb{Q}}_{Q,Q'}:=
\prod_{\mathbf{v}\in N_{Q,Q'}} \widetilde{\mathcal{M}}_\mathbb{C}\cdot y_\mathbf{v}
\subset \hat{\mathcal{T}}^{\mathbb{Q}}_{Q}.
\]
Note that $\hat{\mathcal{T}}^{\mathbb{Q}}_{Q,Q'}$ is also a subalgebra of $\hat{\mathcal{T}}^{\mathbb{Q}}_{Q'}$. We put
\begin{align*}
\mca{A}_{Q,W,k}:=&\,
1+\sum_{\mathbf{v}\in N_{Q,Q'}}
\mathfrak{M}_{\mr{vir}}(Q,W;\mathbf{v})_k\cdot y_\mathbf{v}\\
=&\,
1+\sum_{\mathbf{v}\in N_{Q,Q'}}
\mathfrak{M}_{\mr{vir}}(Q',W';\mathbf{v}')^k\cdot y_\mathbf{v}\quad (\text{Theorem \ref{thm_45}})\\
\in&\
\hat{\mathcal{T}}^{\mathbb{Q}}_{Q,Q'}.
\end{align*}
and
\begin{align*}
\mathbb{E}(s_k)
&:=
\sum_{n\geq 0}\frac{[\mr{pt}]}{[\mr{GL}_n]_\mr{vir}}\cdot {y}_{\,[\,(s_k)^{\oplus n}]}\in \mca{T}_Q^{\mathbb{Q}},\\
\mathbb{E}(s'_k)
&:=
\sum_{n\geq 0}\frac{[\mr{pt}]}{[\mr{GL}_n]_\mr{vir}}\cdot {y}_{\,[\,(s_k')^{\oplus n}]}\in \mca{T}^{\mathbb{Q}}_{Q'}
\end{align*}
(see Example \ref{ex_dilog}).
In the same way as Theorem \ref{thm_FP}, we can see the following factorizations :
\begin{align*}
\mca{A}_{Q,W}&=\mca{A}_{Q,W,k}\times \mathbb{E}(s_k) \in \hat{\mathcal{T}}^{\mathbb{Q}}_{Q},\\
\mca{A}_{Q',W'}&=\mathbb{E}(s'_k) \times \mca{A}_{Q,W,k} \in \hat{\mathcal{T}}_{Q'}^{\mathbb{Q}}.
\end{align*}
Now we get the following wall-crossing formula for the motivic DT invariants :
\begin{thm}\label{thm_WC}
We have
\[
\mca{A}_{Q,W}\times \mathbb{E}(s_k)^{-1},\quad
\mathbb{E}(s'_k)^{-1} \times \mca{A}_{Q',W'}\in \hat{\mathcal{T}}_{Q,Q'}^{\mathbb{Q}}
\]
and they coincide.
\end{thm}
\subsection{Refined DT invariants}
\begin{NB}
We put
\[
\hat{{T}}_{Q}:=
\prod_{\mathbf{v}\in N} \mathbb{Q}(q)\cdot y_\mathbf{v}.
\]
and define the generating function of {\it refined Donaldson-Thomas invariants} by
\[
A_{Q,W}:=1+\sum_{\mathbf{v}\in N}
W\bigl(\mathfrak{M}_{\mr{vir}}(Q,W,\mathbf{v})\bigr)\cdot y_\mathbf{v}
\in \hat{{T}}_{Q}.
\]
\end{NB}
We define $\hat{T}_{Q,Q'}$ in the same way.
Taking the weight polynomial $\mathbb{W}$ of the equation in Theorem \ref{thm_WC}, we get the following formula which describes the relation between the refined DT invariants of $(Q,W)$ and $(Q',W')$ in terms of the quantum dilogarithm :
\begin{thm}
We have
\[
{A}_{Q,W}\times \mathbb{E}_q(y_k)^{-1},\quad
\mathbb{E}_q(y_k^{-1})^{-1} \times {A}_{Q',W'}\in \hat{{T}}_{Q,Q'}
\]
and they coincide.
\end{thm}
\subsection{Appendix : reminders on \cite{quiver-with-potentials}}\label{subsec_appendix}
For a QP $(Q,W)$ and a vertex $k$, we associate a new QP $\widetilde{\mu}_k(Q,W)=(\widetilde{Q},\widetilde{W})$ as follows.
We put $\widetilde{Q}_0=Q_0$ and $\widetilde{Q}_1$ is the union of
\begin{itemize}
\item all the arrows $c\in Q_1$ not incident to $k$,
\item a ``composite'' arrow $[ba]$ from $t(a)$ to $h(b)$ for each $a$ and $b$ with $h(a)=t(b)=k$, and
\item an opposite arrow $a^*$ (resp. $b^*$) for each incoming arrow $a$ (resp. outgoing arrow $b$) at $k$.
\end{itemize}
The new potential is given by
\[
\widetilde{W}:=[W]+\Delta
\]
where
\[
\Delta:=\sum_{a,b\in Q_1; h(a)=t(b)=k}[ba]a^*b^*
\]
and $[W]$ is obtained by substituting $[ba]$ for each factor $ba$ occurring in the expansion of $W$.
What we have been assuming is that there is an automorphism $\psi$ of $\mathbb{C}\widetilde{Q}$ such that we have a decomposition
\begin{equation}\label{eq_decomp}
(\psi(\widetilde{Q}),\psi(\widetilde{W}))
\simeq
(\psi(\widetilde{Q})_{\mr{red}},\psi(\widetilde{W})_{\mr{red}})
\oplus
(\psi(\widetilde{Q})_{\mr{triv}},\psi(\widetilde{W})_{\mr{triv}})
\end{equation}
with $\psi(\widetilde{Q})_{\mr{red}}=\mu_k(Q)$.
We put $\mu_k(W):=\psi(\widetilde{W})_{\mr{red}}$.
The reader may refer \cite[\S 5]{quiver-with-potentials} for the details\footnote{In \cite{quiver-with-potentials}, it is shown that for a generic $W$ we always have such an automorphism of the completion of $\mathbb{C}\widetilde{Q}$. Here we assume that have an automorphism of $\mathbb{C}\widetilde{Q}$, otherwise the mutation of the potential can be infinite. }.
We fix the automorphism $\psi$ and the decomposition \eqref{eq_decomp}.
Then a $J_{\widetilde{Q},\widetilde{W}}$-module is canonically identified with a $J_{\mu_k(Q,W)}$-module.
Take $V\in \mr{M}(Q;\mathbf{v})_k$.
We define $\widetilde{V}:=\oplus \widetilde{V}_i$ by $\widetilde{V}_i:=V_i$ for $i\neq k$ and
\[
\widetilde{V}_k
:=
\mr{coker}
\biggl(\
\sum_{t(b)=k}b
\colon
V_{k}\to \bigoplus_{t(b)=k}V_{h(b)}
\,\biggr).
\]
Note that the sum of the maps above is injective.
We define the action of $\mathbb{C}\widetilde{Q}$ on $\widetilde{V}$ as follows:
\begin{itemize}
\item for an arrow $c\in Q_1$ not incident to $k$, we associate
\[
\widetilde{V}_{t(c)}={V}_{t(c)} \overset{c\,\,}{\longrightarrow} {V}_{h(c)}=\widetilde{V}_{h(c)},
\]
\item for a ``composite'' arrow $[ba]$, we associate the composition
\[
\widetilde{V}_{t([ba])}={V}_{t(a)} \overset{b\circ a\,\,}{\longrightarrow} {V}_{h(b)}=\widetilde{V}_{h([ba])},
\]
\item for an opposite arrow $a^*$ of an incoming arrow $a$ at $k$, we associate the map induced by
\[
\sum_{t(b)=k} \partial_{[ba]}W\colon
\bigoplus_{t(b)=k}V_{h(b)}
\longrightarrow
V_{t(a)}=\widetilde{V}_{h(a^*)},
\]
\item for an opposite arrow $b^*$ of an outgoing arrow $b$ at $k$, we associate
\[
\widetilde{V}_{t(b^*)}={V}_{h(b)}
\hookrightarrow
\bigoplus_{t(b')=k}V_{h(b')}
\twoheadrightarrow
\widetilde{V}_k
=
\widetilde{V}_{h(b^*)}.
\]
\end{itemize}
Then $\widetilde{V}$ belongs to $\mr{M}(\widetilde{Q},\mathbf{v}')^k$ where $\mathbf{v}'=\phi_k(\mathbf{v})$.
We can verify that if $V\in \mr{M}({Q},W, \mathbf{v})_k$ then $\widetilde{V}\in \mr{M}(\widetilde{Q},\widetilde{W}, \mathbf{v}')^k$.
Take $U\in \mr{M}(\widetilde{Q};\mathbf{u})^k$.
We define $\widehat{U}:=\oplus \widehat{U}_i$ by $\widehat{U}_i:=U_i$ for $i\neq k$ and
\[
\widehat{U}_k
:=
\mr{ker}
\biggl(\
\sum_{h(b^*)=k}b^*
\colon
\bigoplus_{h(b^*)=k}U_{t(b^*)}
\to
U_k
\,\biggr).
\]
Note that the sum of the maps above is surjective.
We define the action of $\mathbb{C}\widetilde{\widetilde{Q}}$ on $\widehat{U}$ in a similar way.
If $U\in \mr{M}(\widetilde{Q},\widetilde{W}, {\mathbf{u}})^k$ then $\widetilde{U}\in \mr{M}(\widetilde{\widetilde{Q}},\widetilde{\widetilde{W}}, \phi_k^{-1}(\mathbf{u}))_k$.
In the proof of \cite[Theorem 5.7]{quiver-with-potentials}, an explicit equivalence between $(Q,W)$ and $(\widetilde{\widetilde{Q}},\widetilde{\widetilde{W}})$ is given.
This induces an identification of elements in $\mr{M}(\widetilde{\widetilde{Q}},\widetilde{\widetilde{W}}, \mathbf{u})_k$ with ones in $\mr{M}({Q},{W}, \mathbf{u})_k$
We can verify the composition of these there functors is identity.
Combined with the identification of $J_{\widetilde{Q},\widetilde{W}}$-modules and $J_{\mu_k(Q,W)}$-modules we get the following equivalence:
\begin{prop}
\begin{equation}\label{eq_tilting}
\Phi\colon \mr{mod}(J(Q,W))_k
\overset{\sim}{\longrightarrow}
\mr{mod}(J(Q',W'))^k.
\end{equation}
\end{prop}
\begin{rem}
The derived equivalence given by \textup{\cite{dong-keller}} induces the equivalence of the two categories above.
Here we use \cite{quiver-with-potentials}'s construction since we need the explicit description of $\Phi$ to prove the following proposition.
\end{rem}
\begin{prop}\label{prop_tilting}
The equivalence \eqref{eq_tilting} induces
\begin{equation*}
\mr{mod}(J(Q,W)_C)_k
\simeq
\mr{mod}(J(Q',W')_{C'})^k.
\end{equation*}
\end{prop}
\begin{proof}
First, take $V\in \mr{mod}(J(Q,W)_C)_k$.
For each composite arrow $[ba]$, the map $[ba]$ vanishes on $\Phi(V)$ since the map $a$ vanishes on $V$.
Hence we have $\Phi(V)\in \mr{mod}(J(Q',W')_{C'})^k$.
Next, assume that $\Phi(V)\in \mr{mod}(J(Q',W')_{C'})^k$ for $V\in \mr{mod}(J(Q,W))_k$.
For any $a$ with $h(a)=k$, $a$ vanishes on $V$ since
$\sum_{t(b)=k}[ba]$ vanished and $\sum_{t(b)=k}b$ is injective.
Hence we have $V\in \mr{mod}(J(Q,W)_{C})_k$.
\end{proof}
\begin{rem}
This gives a generalization of a part of the results of \cite{bridgeland-stern}.
\end{rem}
\begin{NB}
Let $\Gamma_{Q,W}$ be the Ginzburg dg algebra and
$\mathcal{D}_{Q,W}$ be the derived category of dg modules with finite dimensional cohomologies over $\Gamma_{Q,W}$.
Let
\[
\Phi=\Phi_k \colon \mathcal{D}_{Q,W} \overset{\sim}{\longrightarrow} \mathcal{D}_{Q',W'}
\]
be the derived equivalence (\cite{dong-keller,keller-completion}) which induces the isomorphism $\phi=\phi_k$ of the Grothendieck group given by
\[
\phi([s_i])=
\begin{cases}
[s'_i]+Q(i,k)[s'_k] & i\neq k.\\
-[s'_k] & i=k.
\end{cases}
\]
Assume that $(Q,W)$ is Calabi-Yau, that is, $\Gamma_{Q,W}$ is concentrated on degree $0$.
In such a case
\begin{itemize}
\item[(1)] $\mathcal{D}_{Q,W}\simeq D^b(\mr{mod}J(Q,W))$,
\item[(2)] $\mu_k(Q,W)$ is also Calabi-Yau (\cite{keller-completion}).
\end{itemize}
Let $\Phi$ be the derived equivalence between $D^b(\mr{mod}J(Q,W))$ and $D^b(\mr{mod}J(\mu_kQ,\mu_kW))$. For any $V\in D^b(\mr{mod}J(Q,W))$, we have
\[
H^j(\Phi(V))_i=0\ (j\neq 0),\quad H^0(\Phi(V))_i=V_i
\]
for $i\neq k$ and the map
\[
[ba] \colon H^0(\Phi(V))_{s(a)} \to H^0(\Phi(V))_{t(b)}
\]
is given by the composition $b\circ a$.
We define the following full subcategories
\begin{align*}
\mr{mod}(J(Q,W))_k
&:=
\{V\in \mr{mod}(J(Q,W))\mid \Hom(s_k,V)=0\},\\
\mr{mod}(J(Q,W)_C)_k
&:=
\mr{mod}(J(Q,W))_k
\cap
\mr{mod}(J(Q,W)_C),\\
\mr{mod}(J(Q',W'))^k
&:=
\{V\in \mr{mod}(J(Q',W'))\mid \Hom(V,s'_k)=0\},\\
\mr{mod}(J(Q',W')_{C'})^k
&:=
\mr{mod}(J(Q',W'))^k
\cap
\mr{mod}(J(Q',W')_{C'}).
\end{align*}
Note that for $V\in \mr{mod}(J(Q,W))$ the condition $\Hom(s_k,V)=0$ is equivalent to the injectiveness of
\[
\sum_{s(b)=k}V_k\to \biggl(\,\bigoplus_{s(b)=k} V_{t(b)}\biggr)
\]
and for $V\in \mr{mod}(J(Q',W'))$ the condition $\Hom(V,s'_k)=0$ is equivalent to the surjectiveness of
\[
\sum_{t(a)=k}\biggl(\,\bigoplus_{t(a)=k} V_{s(a)}\biggr)\to V_k.
\]
It is shown in \cite{keller-completion}, the derived equivalence $\Phi$ induces
\[
\mr{mod}(J(Q,W))_k
\simeq
\mr{mod}(J(Q',W'))^k.
\]
\begin{prop}\label{prop_tilting}
The derived equivalence induces
\[
\mr{mod}(J(Q,W)_C)_k
\simeq
\mr{mod}(J(Q',W')_{C'})^k.
\]
\end{prop}
\begin{proof}
First, take $E\in \mr{mod}(J(Q,W)_C)_k$.
For a composition $ba$ through $k$, $[bc]$ vanishes on $\Phi(E)$ since $a$ vanishes on $E$.
Hence we have $\Phi(E)\in \mr{mod}(J(Q',W')_{C'})^k$.
Next, assume that $\Phi(E)\in \mr{mod}(J(Q',W')_{C'})^k$ for $E\in \mr{mod}(J(Q,W))_k$.
For any $a$ with $t(a)=k$, $a$ vanishes on $E$ since
$\sum_{s(b)=k}[ba]$ vanished and $\sum_{s(b)=k}b$ is injective.
Hence we have $E\in \mr{mod}(J(Q,W)_{C})_k$.
\end{proof}
\begin{rem}
This gives a generalization of a part of the results of \cite{bridgeland-stern}.
\end{rem}
\end{NB}
\bibliographystyle{amsalpha}
|
\section{Introduction}
Let $(A,\mathfrak m)$ be a commutative Noetherian local ring and $I$ be an
ideal of $A$. The graded algebras, the Rees algebra, $R(I) :=
\oplus_{n\geq0}I^nt^n \subset R[t]$, the associated graded ring, $G(I)
:= \oplus_{n\geq0}I^n/I^{n+1} \cong R(I)/IR(I)$ and the fiber cone,
$F(I) := \oplus_{n\geq0}I^n/\mathfrak m I^n$ are together known as the blowup
algebras associated to $I$.
In this article, our aim is to study the Castelnuovo-Mumford
regularity and the Gorensteinness of the fiber cone. We do this by
relating them with the corresponding properties of certain other
graded modules. Cortadellas and Zarzuela, in a series papers, used
certain graded modules associated to filtrations of modules to study
the depth properties of the fiber cone \cite{cz1}, \cite{cz2},
\cite{co2}. We use these graded modules and other blowup algebras to
study the regularity and the Gorensteinness of the fiber cone.
For a standard graded algebra $S = \oplus_{n\geq0}S_n$ over a
commutative Noetherian ring $S_0$ and a finitely generated graded
$S$-module $M = \oplus_{n\geq0}M_n$, define
$$
a(M) := \left\{
\begin{array}{ll}
\operatorname{max}\{n \; | \; M_n \neq 0\} & \mbox{ if } M \neq 0 \\
-\infty & \mbox{ if } M = 0.
\end{array}
\right.
$$
For $i \geq 0$, set
$$
a_i(M) := a(H^i_{S_+}(M)),
$$
where $S_+$ denotes the ideal of $S$ generated by the homogeneous
elements of positive degree and $H^i_{S_+}(M)$ denotes the $i$-th
local cohomology module of $M$ with respect to the ideal $S_+$. The
{\em Castelnuovo-Mumford regularity} (or {\em regularity}) of
$M$ is defined as the number
$$
\operatorname{reg}(M) := \operatorname{max}\{a_i(M) + i \; | \; i \geq 0 \}.
$$
Let $(A,\mathfrak m)$ be a local ring and $I$ be any ideal.
The Castelnuovo-Mumford regularity of $R(I)$ and $G(I)$ have been well
studied in the past. Ooishi proved that $\operatorname{reg} R(I) = \operatorname{reg} G(I)$,
\cite{o} (see also \cite{t1}). In \cite{t1}, Trung studied the
vanishing behavior of the local cohomology modules of the
associated graded ring and the Rees algebra and derived that for any
ideal in a Noetherian local ring $\operatorname{reg} R(I) = \operatorname{reg} G(I)$. It
can easily be seen that such an equality is not true in the case of
the fiber cone and the associated graded ring (see Section 2). In
Section 2, we prove that for any ideal of analytic spread one in
a Noetherian local ring, the regularity of the fiber cone is bounded
above by the regularity of the associated graded ring. If the ideal
contains a regular element, we show that the equality holds in the
above case (Theorem \ref{thm7}). We also prove that,
under some assumptions, the regularity of the fiber cone is
bounded below by the regularity of the associated graded ring and
obtain certain sufficient conditions for the equality.
In Section 3, we study the Gorenstein property of the fiber cone.
The Gorenstein property of the Rees algebra and the
associated graded ring has been very well studied, see for example
\cite{gn}, \cite{hrs}, \cite{hku}, \cite{eh}, \cite{tvz}. It is known
that the Gorenstein fiber cones behave differently. For
example, it is known that, unlike in
the case of associated graded ring, fiber cone can be Gorenstein
without the ambient ring being Gorenstein. Also, it can easily be seen
that the symmetry of the Hilbert series of the fiber cone does not
assure the Gorensteinness for fiber cone (\cite[Example 6.2]{jpv}). In
Section 3, we obtain an expression for the canonical module of the
fiber cone (Proposition \ref{pro4}). Using the structure of the
canonical module, we obtain a necessary and sufficient condition for
the Gorenstein property of the fiber cone (Theorem \ref{gor-char}). We
also obtain an upper bound for the regularity of the canonical module
of $F(I)$. We end the article by proving that the multiplicity of the
canonical module of the fiber cone is strictly less than the multiplicity
of the canonical module of the associated graded ring, except for the
maximal ideal. Throughout this article, $(A,\mathfrak m)$ will always denote a
Noetherian local ring of dimension $d$ and with infinite residue
field. All the computations in this article have been performed using
the computer algebra package CoCoA, \cite{co}. \\
{\bf Acknowledgement :} We would like to sincerely thank the referee for
a thorough reading of the manuscript, pointing out some mistakes and
suggesting some improvements.
\section{Castelnuovo-Mumford regularity of the fiber cone}
In this section we study the regularity of the fiber cone. Unlike in
the case of the associated graded ring and the Rees algebra, there is
no equality between the regularities of the fiber cone and the Rees
algebra. For example if $A$ is a non-Buchsbaum ring and $I$ is
generated by a system of parameters and not a $d$-sequence (such a
sequence exists by \cite[Proposition 1.7]{h}), then from Corollary 5.2
in \cite{t1} we have $\operatorname{reg} R(I) > 0$. Since $I$ is generated by a
system of parameters $\operatorname{reg} F(I) = 0$. Therefore $\operatorname{reg} F(I) < \operatorname{reg}
R(I)$. We show that under some hypothesis, the regularity of the fiber
cone is at most that of the Rees algebra. We also provide some
sufficient conditions for the equality.
Let $A$ be a Noetherian local ring of dimension $d > 0$ and
$I \subset A$ be an ideal. Consider the filtration
$$
\mathcal F : A \supset \mathfrak m \supset \mathfrak m I \supset \mathfrak m I^2 \supset \cdots
$$
Let $R(\mathcal F) := A \oplus \mathfrak m t \oplus \mathfrak m It^2 \oplus \mathfrak m I^2t^3 \oplus
\cdots$. Then $R(\mathcal F)$ is a finitely generated graded $R(I)$-module.
Consider the exact sequences of $R(I)$-modules:
\begin{equation}\label{eqn3}
0 \rightarrow R(I) \rightarrow R(\mathcal{F}) \rightarrow \mathfrak m G(I)(-1) \rightarrow 0
\end{equation}
\begin{equation} \label{eqn6}
0 \rightarrow \mathfrak m G(I) \rightarrow G(I) \rightarrow F(I) \rightarrow 0
\end{equation}
We use the above two exact sequences and the corresponding long exact
sequence of local cohomology modules to the study the vanishing
properties of the local cohomology modules of the fiber cone.
Throughout this section we assume that $I$ is an ideal of analytic
spread := $\dim F(I) = \ell > 0$. We begin this section with a remark on
the top local cohomology modules of $R(I)$ and $R(\mathcal F)$.
\begin{remark} \label{rmk2}
Suppose $\underline{x}=x_1,\ldots,x_\ell$ generates a minimal
reduction of $I$. Note that $R(I)_+$ is generated radically by $\ell$
elements. Denote by $\underline{x^k}$ the sequence
$x_1^k,\ldots,x_\ell^k$. Then for all $n \in \mathbb Z$,
$$\left [H^\ell_{R(I)_+}(R(I))
\right ]_n\cong \underset{\underset{k}{\longrightarrow}}{\lim~} \frac{I^{\ell k+n}}{(\underline{x}^k)
I^{(\ell-1)k+n}} \mbox{ and }
\left [ H^\ell_{R(I)_+}(R(\mathcal{F}))\right ]_n \cong \underset{\underset{k}{\longrightarrow}}{\lim~} \frac{\mathfrak m
I^{\ell k+n-1}}{(\underline{x}^k) \mathfrak m I^{(\ell-1)k+n-1}}.$$
This implies that $a_\ell(R(\mathcal{F}))-1 \leq a_\ell(R(I))$.
\end{remark}
In the following theorem, we show that the regularity of the fiber
cone is bounded above by the regularity of the associated graded ring
when $\ell = 1$. For convenience, we will denote $H^i_{R(I)_+}(M)$ by
$H^i(M)$ for the rest of the section.
\begin{theorem} \label{thm7}
Let $(A, \mathfrak m)$ be a Noetherian local ring and $I$ be an ideal of $A$
with $\ell= 1$. Then $\operatorname{reg} F(I) \leq \operatorname{reg} G(I)$. Furthermore, if
$\operatorname{grade} I = 1$, then $\operatorname{reg} F(I)=\operatorname{reg} G(I) = r(I),$ where $r(I)$
denotes the reduction number of $I$.
\end{theorem}
\begin{proof}
Since by hypothesis $\ell=1$, $I$ is generated by a single
element up to radical. Therefore $H^i(M)= (0)$ for all $i \geq 2$ and
for any finitely generated graded $R(I)$-module $M$. From the exact
sequence \eqref{eqn6}, we have the long exact sequence:
\begin{eqnarray*}
0 &\rightarrow& H^0(\mathfrak m G(I)) \rightarrow H^0(G(I)) \rightarrow H^0(F(I)) \\
&\rightarrow& H^1(\mathfrak m G(I)) \rightarrow H^1(G(I)) \rightarrow H^1(F(I)) \rightarrow 0.
\end{eqnarray*}
It follows that $a_0(\mathfrak m G(I)) \leq a_0(G(I))$ and $a_1(F(I)) \leq
a_1(G(I))$. From the exact sequence \eqref{eqn3}, we have the long
exact sequence:
\begin{eqnarray*}
0 &\rightarrow& H^0(R(I)) \rightarrow H^0(R(\mathcal{F})) \rightarrow H^0(\mathfrak m G(I))(-1) \\
&\rightarrow& H^1(R(I)) \rightarrow H^1(R(\mathcal{F})) \rightarrow
H^1(\mathfrak m G(I))(-1) \rightarrow 0 .
\end{eqnarray*}
Therefore $a_1(\mathfrak m G(I)(-1)) = a_1(\mathfrak m G(I))+1 \leq
a_1(R(\mathcal{F}))$. From Remark \ref{rmk2}, it follows that
$a_1(\mathfrak m G(I)) \leq a_1(R(I))$. Since $G_+$ is radically generated
by one element, $H^1(G(I)) \neq 0$ and hence by Theorem 3.1 of
\cite{t1}, we get $a_1(R(I)) \leq a_1(G(I))$ so that $a_1(\mathfrak m G(I))
\leq a_1(G(I))$. Therefore $\operatorname{reg} \mathfrak m G(I) \leq \operatorname{reg} G(I)$. Now the
regularity behaviour under the exact sequence \eqref{eqn6} yields
that $$\operatorname{reg} F(I) \leq \operatorname{max}\{\operatorname{reg} G(I),\operatorname{reg} \mathfrak m G(I)-1\}= \operatorname{reg} G(I).$$
\vskip 0.2cm Now assume that $\operatorname{grade} I = 1$. Then by (\cite{cz3},
page 764) we have $\operatorname{reg} F(I)=r_J(I)$, for any minimal reduction $J$
of $I$. Also we have $\operatorname{reg} G(I)= r_J(I)$ (see for example
(\cite{hz}, Proposition 3.6) ). Therefore $\operatorname{reg} F(I)=\operatorname{reg} G(I) =
r(I)$.
\end{proof}
Now we give a lower bound for the regularity of fiber cone under some
assumptions. In \cite{cz3}, Cortadellas and Zarzuela proved that if
the depth of the fiber cone and the associated graded ring is at least
$\ell -1$, then the regularities of these two algebras are equal. We
generalize this result in the following theorem and retrieve their
result in the above mentioned case.
For $x \in I \; \backslash \mathfrak m I$, let $x^*$ denote
the image of $x$ in $I/I^2$ and $x^o$ denote the image of $x$ in $I/\mathfrak m
I$.
\begin{theorem} \label{pro8}
Let $(A,\mathfrak m)$ be a Noetherian local ring and $I$ be an ideal of $A$.
Suppose $\operatorname{grade} I = \ell$ and $\operatorname{grade} G(I)_+ \geq \ell-1$. Then
$\operatorname{reg} F(I) \geq \operatorname{reg} G(I)$. Furthermore, if $\operatorname{depth} F(I) \geq
\ell-1$, then $\operatorname{reg} F(I)= \operatorname{reg} G(I)$.
\end{theorem}
\begin{proof} If $\ell=1$, then the proposition follows from Theorem
\ref{thm7}. Suppose $\ell \geq 2$. Let $x_1,\ldots,x_\ell$ be a
minimal generating set for a minimal reduction $J$ of $I$ such that
$x_1^*,\ldots,x_\ell^* \in I/I^2$ is a filter regular sequence for
$G(I)$ and $x_1^o,\ldots,x_\ell^o \in I/\mathfrak m I$ is a filter regular
sequence for $F(I)$. Since $\operatorname{grade} G(I)_+ \geq \ell -1, \;
x_1^*,\ldots,x_{\ell-1}^*$ is $G(I)$-regular. Let ``-'' denote
modulo $(x_1,\ldots,x_{\ell-1})$. Then $G(\bar{I}) \cong
G(I)/(x_1^*,\ldots,x_{\ell-1}^*), \; F(\bar{I}) \cong
F(I)/(x_1^o,\ldots,x_{\ell-1}^o)$ and $\operatorname{reg} G(\bar{I})= \operatorname{reg} G(I)$.
Since $\dim F(\bar{I}) = 1$, by Theorem \ref{thm7} we get $\operatorname{reg}
F(\bar{I})= \operatorname{reg} G(\bar{I}) = \operatorname{reg} G(I)$. From \cite[Proposition
1.2]{ch}, it follows that $\operatorname{reg} F(I) \geq \operatorname{reg} F(\bar{I})$. This
implies that $\operatorname{reg} F(I) \geq \operatorname{reg} G(I)$. \vskip 2mm Now assume
$\operatorname{depth} F(I) \geq \ell-1$. Then $x_1^o,\ldots,x_{\ell-1}^o$ is
$F(I)$-regular. Then $\operatorname{reg} F(I)= \operatorname{reg} F(\bar{I})$. Therefore $\operatorname{reg}
F(I)= \operatorname{reg} G(I)$ as required.
\end{proof}
Now we give certain instances where the regularity of the fiber cone
is equal to the regularity of the Rees algebra or the associated
graded ring.
\begin{proposition}\label{pro6}
Let $(A,\mathfrak m)$ be a Noetherian local ring and $I$ be an
ideal of $A$. If $\operatorname{reg} R(\mathcal{F}) \leq \operatorname{reg} R(I)$, then $\operatorname{reg}
F(I)= \operatorname{reg} R(I)$.
\end{proposition}
\begin{proof}
From the exact sequence \eqref{eqn3} and the fact that $\operatorname{reg} \mathfrak m
G(I)(-1)=\operatorname{reg} \mathfrak m G(I)+1$, it follows that $\operatorname{reg} \mathfrak m G(I)+1 \leq \operatorname{max}\{
\operatorname{reg} R(I)-1, \operatorname{reg} R(\mathcal{F})\}$. Since $\operatorname{reg} R(\mathcal{F}) \leq
\operatorname{reg} R(I)$, the above inequality implies that $\operatorname{reg} \mathfrak m G(I)+1 \leq \operatorname{reg}
R(I)$. From the exact sequence \eqref{eqn6}, we get $\operatorname{reg} F(I) \leq
\operatorname{max} \{\operatorname{reg} \mathfrak m G(I)-1, \operatorname{reg} G(I)\}$. Since $\operatorname{reg} R(I) = \operatorname{reg} G(I)$,
the above inequality implies that $\operatorname{reg} F(I) \leq \operatorname{reg} R(I)$.
\vskip 2mm
Now from the exact sequence \eqref{eqn6}, $\operatorname{reg} G(I)\leq \operatorname{max} \{\operatorname{reg}
\mathfrak m G(I),\operatorname{reg} F(I)\}$. Since $\operatorname{reg} G(I)=\operatorname{reg} R(I)$ and $\operatorname{reg} \mathfrak m G(I)
\leq \operatorname{reg} R(I)-1,$
the above inequality yields that
$$\operatorname{max} \{\operatorname{reg} \mathfrak m G(I),\operatorname{reg} F(I)\}= \operatorname{reg} F(I) \mbox{ and }
\operatorname{reg} R(I) = \operatorname{reg} G(I) \leq \operatorname{reg} F(I).$$ Therefore we have
$\operatorname{reg} R(I)=\operatorname{reg} F(I)$. This completes the proof.
\end{proof}
Note that, if $\operatorname{grade}_{R(I)_+} R(\mathcal F) \geq \ell$, then
$H^i(R(\mathcal F)) = 0$ for $i < \ell$ and from Remark \ref{rmk2}, it follows
that $\operatorname{reg} R(\mathcal F) \leq \operatorname{reg} R(I)$. The next proposition gives yet
another instance of the equality of the regularity of these graded
algebras.
\begin{proposition}\label{pro9}
Let $(A,\mathfrak m)$ be a Noetherian local ring and $I$ be an $\mathfrak m$-primary
ideal of $A$ such that $\operatorname{grade} I > 0$. Suppose $I^{n_{0}}= \mathfrak m
I^{n_{0}-1}$ for some $n_0 \in \mathbb N$. Then $\operatorname{reg} F(I)= \operatorname{reg} G(I)$.
\end{proposition}
\begin{proof}
Since $I^{n_{0}}= \mathfrak m I^{n_{0}-1}$ for some $n_0 \in
\mathbb N$, it follows that $\mathfrak m G(I)$ is Artinian. If $I = \mathfrak m$, then the
assertion of the theorem follows trivially. If $I \neq \mathfrak m$, then $\mathfrak m
G(I) \neq 0$. Therefore $H^0(\mathfrak m G(I))= \mathfrak m G(I) \neq 0$
and $H^i(\mathfrak m G(I))= 0$ for all $i > 0$.
From the exact sequence \eqref{eqn6} we have
$$0 \rightarrow H^0(\mathfrak m G(I)) \rightarrow H^0(G(I)) \rightarrow
H^0(F(I)) \rightarrow 0$$
and $H^i(G(I)) \cong H^i(F(I))$ for all $i > 0$. Therefore $a_0(F(I))
\leq a_0(G(I))$ and $a_i(G(I))= a_i(F(I))$ for $i > 0$. If $\operatorname{depth}
G(I) >0$, then $H^0(G(I))=(0)$ and hence from the above exact sequence
we get $H^0(\mathfrak m G(I))= 0$, which is a contradiction. Therefore $\operatorname{depth}
G(I)=0$. Then by the Proposition 6.1 in \cite{t1}, we have
$$a_0(F(I)) \leq a_0(G(I)) < a_1(G(I)) = a_1(F(I)).$$
Therefore $\operatorname{reg}
F(I)= \operatorname{max}\{a_i(F(I))+i : i \geq 1 \}= \operatorname{max}\{a_i(G(I))+i : i \geq 1\}=
\operatorname{reg} G(I)$ as required.
\end{proof}
\begin{proposition}
Let $(A,\mathfrak m)$ be a Noetherian local ring and $I$ be an ideal of $A$
such that $\operatorname{grade} I > 0$. Assume that $\mathfrak m G(I)$ is a Cohen-Macaulay
$R(I)$-module of dimension $\ell$. Then
\begin{enumerate}
\item[(i)] $\operatorname{reg} F(I) \leq \operatorname{reg} R(I)$;
\item[(ii)] if $a_\ell(R(\mathcal{F}))-1 < a_\ell(R(I))$, then $\operatorname{reg}
F(I) = \operatorname{reg} R(I)$;
\item[(iii)] if $a_\ell(R(\mathcal{F}))-1 = a_\ell(R(I)),$ then $\operatorname{reg}
\mathfrak m G(I) \leq \operatorname{reg} R(I)$ and $\operatorname{reg} F(I) \leq \operatorname{reg} R(I)$. Furthermore,
if $\operatorname{reg} \mathfrak m G(I) < \operatorname{reg} G(I)$, then $\operatorname{reg} F(I) = \operatorname{reg} R(I)$.
\end{enumerate}
\end{proposition}
\begin{proof}
From the short exact sequence \eqref{eqn3}, there is a long exact
sequence of the local cohomology modules:
$$\cdots \rightarrow H^i(R(I)) \rightarrow H^i(R(\mathcal{F})) \rightarrow H^i(\mathfrak m G(I)(-1)) \rightarrow H^{i+1}(R(I)) \rightarrow \cdots.$$
Since $\mathfrak m G(I)$ is Cohen-Macaulay, $H^i(\mathfrak m G(I))=0$ for $i \neq \ell$.
Therefore, it follows from the above long exact sequence that
$H^i(R(I)) \cong H^i(R(\mathcal{F}))$ for $i < \ell$. This implies
that $a_i(R(\mathcal{F})) = a_i(R(I))$ for $i < \ell$. Using Remark \ref{rmk2},
we have $a_\ell(R(\mathcal{F}))-1 \leq a_\ell(R(I))$. Therefore (i)
follows from (ii) and (iii).
\vskip 0.2cm
If $a_\ell(R(\mathcal{F}))-1 < a_\ell(R(I))$, then
$a_\ell(R(\mathcal{F})) \leq a_\ell(R(I))$. Therefore $a_i(R(\mathcal{F}))
\leq a_i(R(I))$ for all $i$. This implies that $\operatorname{reg} R(\mathcal{F})
\leq \operatorname{reg} R(I)$. Therefore from the Proposition \ref{pro6} we have
$\operatorname{reg} F(I)= \operatorname{reg} R(I)$. This proves (ii).
\vskip 0.2cm
Suppose $a_\ell(R(\mathcal{F}))-1 = a_\ell(R(I))$. Since
$a_i(R(\mathcal{F})) = a_i(R(I))$ for $i < \ell,$ we have
$\operatorname{reg} R(\mathcal{F}) \leq \operatorname{reg} R(I)+1$. Therefore from the exact sequence
\eqref{eqn3}, it follows that
$$\operatorname{reg}(\mathfrak m G(I)(-1)) = \operatorname{reg} \mathfrak m G(I)+1 \leq \operatorname{max}\{ \operatorname{reg} R(I)-1, \operatorname{reg}
R(\mathcal{F})\} \leq \operatorname{reg} R(I)+1.$$
From the exact sequence \eqref{eqn6}, it follows that $\operatorname{reg} F(I) \leq \operatorname{max} \{\operatorname{reg} \mathfrak m G(I)-1, \operatorname{reg} G(I)\} \leq \operatorname{reg} R(I)$.
Thus $\operatorname{reg} F(I) \leq \operatorname{reg} R(I)$.
\vskip 0.2cm
Now assume that $\operatorname{reg} \mathfrak m G(I) < \operatorname{reg} G(I)$. From the exact sequence
\eqref{eqn6} it follows that $\operatorname{reg} G(I)\leq \operatorname{max} \{\operatorname{reg} \mathfrak m G(I),\operatorname{reg}
F(I)\}$. Since $\operatorname{reg} \mathfrak m G(I) < \operatorname{reg} G(I)$, the above inequality gives
that $\operatorname{max} \{\operatorname{reg} \mathfrak m G(I),\operatorname{reg} F(I)\}= \operatorname{reg} F(I)$. Therefore $\operatorname{reg}
G(I) \leq \operatorname{reg} F(I)$. Since $\operatorname{reg} G(I)= \operatorname{reg} R(I)$, we have $\operatorname{reg}
R(I) \leq \operatorname{reg} F(I)$. The other inequality is already proved.
Therefore $\operatorname{reg} F(I) = \operatorname{reg} R(I)$. This proves (iii).
\end{proof}
We conclude this section by giving some examples to illustrate the regularity behavior of the
fiber cone. The following example shows that the regularity of the
fiber cone can be strictly less than the regularity of the associated
graded ring even when $\operatorname{reg} F(I) > 0$. Let $R = \oplus_{n\geq 0}R_n$
be a finitely generated standard graded algebra. Then $R \cong
R_0[X_1, \ldots, X_m]/J$ for some $m$ and a homogeneous ideal $J$,
where $X_1, \ldots, X_m$ are indeterminates over $R_0$. Then the
relation type of $R$, denoted by $\operatorname{reltype}(R)$ is defined to be the
maximum degree of a minimal generating set of $J$. It is known that
$\operatorname{reltype}(R) \leq \operatorname{reg} R + 1$, \cite{t1}. Let $k$ denote a field.
\begin{example}\label{reg-less}
Let $A = k[\![X,Y,Z]\!]/(X^2,Y^2,XYZ^2)$ and $\mathfrak m = (x,y,z)$, where
$x=\bar{X}, y=\bar{Y}, z=\bar{Z}$ and $k$ is a field. Then $A$ is a one dimensional
Noetherian local ring. Let $I =(y,z)A$. Then $F(I) \cong k[Y,Z]/(Y^2)$. Therefore $\operatorname{reg} F(I)=1$. Let $\psi : A/I[U,V]\rightarrow G(I)$
be $A/I$-algebra homomorphism defined by $\psi(U)= yt$ and $\psi(V)= zt$. Then $ker(\psi)$ has a generator $(x+I)UV^2$. Therefore
$\operatorname{reltype} G(I) \geq 3$. Since $\operatorname{reltype} G(I) \leq \operatorname{reg} G(I)+1$ we have $\operatorname{reg} G(I) \geq 2$. Thus $\operatorname{reg} F(I) < \operatorname{reg} G(I)$. Here $\operatorname{grade} I = 0$.
\end{example}
\begin{example}
Let $A= k[\![x,y,z]\!]/J$, where $$J = (xy^3, xy^2z, xyz^2, xz^3, x^3z^2,
x^4, y^3z, x^3y, x^2y^2, y^4)$$ and $k$ is a field. Let $I=
(\bar{x}^2,\bar{y},\bar{z})$ and $\mathfrak m= (\bar{x},\bar{y},\bar{z})$. Then
$A$ is one dimensional non-Cohen-Macaulay Noetherian local ring and
$I$ is an $\mathfrak m$-primary ideal satisfying $I^4= \mathfrak m I^3$. Let $S=
k[U,V,W]$. Then $F(I) \cong \frac{S}{(U^2,UV^2,V^3W,UVW^2,UW^3,V^4)}$.
By using CoCoA, \cite{co}, one can see that the minimal free resolution
of $F(I)$ as an $S$-module is
$$0 \rightarrow S(-6)^3 \rightarrow S(-4) \oplus S(-5)^7 \rightarrow
S(-2)\oplus S(-3) \oplus S(-4)^4 \rightarrow S \rightarrow 0.$$
From this the regularity of $F(I)$ is $3$. Therefore by the
Proposition \ref{pro9} we have $\operatorname{reg} F(I)= \operatorname{reg} G(I)=3$. Note that
$\operatorname{grade} I=0$.
\end{example}
The following example shows that the reduction number of $I$ can be
strictly smaller than the regularity of $F(I)$.
\begin{example}
Let $A= k[\![X,Y,Z]\!]/J$, where $J=(X^4,XY^2Z,XYZ^2,YZ^4,Z^5)$. Let
$I=(x^3,y^2,z^2)$ and $\mathfrak m=(x,y,z)$, where $x=\bar{X},y=\bar{Y}$ and
$z=\bar{Z}$. Then $A$ is a non-Cohen-Macaulay Noetherian local ring.
Let $J =(y^2)$, then $I^4 = JI^3$. Therefore $r(I) \leq 3$. Let
$S=k[U,V,W]$. Then $F(I)\cong \frac{S}{(U^2,UVW,W^3,V^2W^2,V^4W)}$.
By using CoCoA, \cite{co}, one can see that the minimal free
resolution of $F(I)$ as $S$-module is
\begin{eqnarray*}
0 &\longrightarrow& R(-6)^2 \oplus R(-7) \longrightarrow R(-4) \oplus
R(-5)^4 \oplus R(-6)^2 \\
&\longrightarrow& R(-2) \oplus R(-3)^2 \oplus R(-4) \oplus R(-5)
\longrightarrow R \longrightarrow 0.
\end{eqnarray*}
Hence $\operatorname{reg}(F(I)) = 4$. Therefore $r(I) < \operatorname{reg} F(I)$.
\end{example}
Though we have proved that under certain conditions, $\operatorname{reg} F(I)$ is
bounded below by $\operatorname{reg} G(I)$, we have not been able to find an example
when this inequality is strict. Therefore we ask:
\begin{question}
Let $(A,\mathfrak m)$ be a Noetherian local ring with infinite residue field
and $I$ an ideal such that $\operatorname{grade} I > 0$. Is $\operatorname{reg} F(I) \leq \operatorname{reg}
G(I)$?
\end{question}
\section{Gorenstein fiber cones}
In this section, we study the Gorenstein property of the fiber cone.
We begin by obtaining an expression for the canonical module of the
fiber cone. We fix the notation for this section. Throughout this
section, we assume that $G(I)$ and $F(I)$ are Cohen-Macaulay.
Let $\omega_{G(I)}$ and $\omega_{F(I)}$ denote the canonical modules of the
associated graded ring $G(I)$ and the fiber cone $F(I)$ respectively. For the
definition and basic properties of canonical modules, see \cite{bh}.
In the original manuscript, the result given below was proved in a
weaker form. We would like to thank the referee for suggesting the
following improved form.
\begin{proposition} \label{pro4}
Let $(A,\mathfrak m)$ be a Noetherian local ring and $I$ be an $\mathfrak m$-primary ideal
such that the associated graded ring $G(I)$ is Cohen-Macaulay. Let
$\omega_{G(I)}= \oplus_{n \in \mathbb Z} \omega_n$ and $\omega_{F(I)}$
be the canonical modules of $G(I)$ and $F(I)$ respectively. Then
\begin{enumerate}
\item[(1)] $\omega_{F(I)}\cong \oplus_{n \in \mathbb Z}(0:_{\omega_n}\mathfrak m )$;
\item[(2)] $a(F(I))=a(G(I))= r-d$, where $r$ is the reduction number
of $I$ with respect to any minimal reduction $J$ of $I$;
\item[(3)] for any $k \in \mathbb N$, $a(F(I^k))= [\frac{a(F(I))}{k}]= [\frac{r-d}{k}]$;
\item[(4)] if $G(I)$ is Gorenstein, then
\end{enumerate}
$$\omega_{F(I)} \cong \bigoplus_{n \in \mathbb Z} \frac{(I^{n+r-d+1}:\mathfrak m)\cap
I^{n+r-d}}{I^{n+r-d+1}}.$$
\end{proposition}
\begin{proof}
(1) Since $G(I)$ is Cohen-Macaulay and $F(I)=G(I)/\mathfrak m G(I)$ is such
that $\dim G(I)=\dim F(I)= d$ we have by (\cite{hio}, Corollary (36.14))
that:
$$\omega_{F(I)} \cong Hom_{G(I)}(F(I), \omega_{G(I)}) \cong (0:_{\omega_{G(I)}} \mathfrak m G(I)) =
(0:_{\omega_{G(I)}} \mathfrak m ) = \oplus_{n \in \mathbb Z} (0:_{\omega_{n}}
\mathfrak m).$$ (2) By definition $a(F(I))= -\min\{ n| [\omega_{F(I)}]_n \neq
0 \}$. Since $\omega_n$ is a finitely generated $A/I$-module for any
$n$ and $I$ is $\mathfrak m$-primary, $A/I$ is of finite length and so
$\omega_n$. As a consequence $(0:_{\omega_n}\mathfrak m) \neq 0$ if and only
if $\omega_n \neq 0$. Therefore $a(F(I)) = a(G(I))$.
On the other hand, it is known (see for instance (\cite{hz}, Proposition 3.6)) that if $G(I)$
is Cohen-Macaulay then $a(G(I))=r-d$, where $r:=r_J(I)$ is the reduction number of $I$ with
respect to any minimal reduction $J$ of $I$.
(3) Since $G(I)$ is Cohen-Macaulay, $G(I^k)$ is also Cohen-Macaulay for any positive integer
$k$ (see for instance (\cite{hz}, Corollary 4.6) ). On the other hand, by
(\cite{hz}, Corollary 4.6) $a(G(I^k))=[\frac{a(G(I))}{k}]$. Thus by (2)
$$a(F(I^k))= a(G(I^k))= [\frac{a(G(I))}{k}]= [\frac{a(F(I))}{k}]= [\frac{r-d}{k}].$$
(4) If $G(I)$ is Gorenstein then $\omega_{G(I)} \cong G(I)(a(G(I)))= G(I)(r-d)$. So
$\omega_n= I^{n+r-d}/I^{n+r-d+1}$ for any $n$ and by (1)
$$\omega_{F(I)} \cong \oplus_{n \in \mathbb Z} (0:_{\omega_n} \mathfrak m) = \bigoplus_{n \in \mathbb Z} (I^{n+r-d+1}:\mathfrak m) \cap I^{n+r-d}/I^{n+r-d+1}.$$
\end{proof}
We know that if $G(I)$ and $F(I)$ are Cohen-Macaulay then $\operatorname{reg} G(I)=r=\operatorname{reg} F(I)$.
Now we prove that $\operatorname{reg} \omega_{G(I)}$ and $\operatorname{reg} \omega_{F(I)}$ are equal.
\begin{corollary}
Suppose $G(I)$ and $F(I)$ are Cohen-Macaulay. Then $\operatorname{reg}
\omega_{G(I)}= \operatorname{reg} \omega_{F(I)}$. In addition if $G(I)$ or $F(I)$
is Gorenstein then $\operatorname{reg} \omega_{G(I)} = d = \operatorname{reg} \omega_{F(I)}$.
\end{corollary}
\begin{proof}
Let $J$ be a minimal reduction of $I$ minimally generated
by $x_1,\ldots,x_d$ such that $x_1^{*},\ldots,x_d^{*} \in G(I)$ and
$x_1^{o},\ldots,x_d^{o} \in F(I)$ are regular sequences. Then $F(I)/J^o \cong F(I/J)$
and $G(I)/J^* \cong G(I/J)$.
Therefore $\operatorname{reg} \omega_{F(I)}= \operatorname{reg} (\omega_{F(I)}/J^0\omega_{F(I)})= \operatorname{reg} \omega_{F(I/J)}$
and $\operatorname{reg} \omega_{G(I)} = \operatorname{reg} \omega_{G(I/J)}$. But $\operatorname{reg} \omega_{F(I/J)}= a(\omega_{F(I/J)})$ and
$\operatorname{reg} \omega_{G(I/J)}= a(\omega_{G(I/J)})$. By Proposition \ref{pro4}(1) we have
$\omega_{F(I/J)}\cong \oplus_{n \in \mathbb Z}(0:_{[\omega_{G(I/J)}]_n}\mathfrak m )$. From this we have
$[\omega_{F(I/J)}]_n \neq 0$ if and only if $[\omega_{G(I/J)}]_n \neq 0$ for any $n$. Therefore
$a(\omega_{F(I/J)}) = a(\omega_{G(I/J)})$. Thus $\operatorname{reg} \omega_{G(I)}= \operatorname{reg} \omega_{F(I)}$.
Now assume $G(I)$ is Gorenstein. Then $\omega_{G(I)} \cong
G(I)(r-d)$. Thus $\operatorname{reg} \omega_{G(I)}= \operatorname{reg} G(I)(r-d) = \operatorname{reg}
G(I)-r+d$. Since $G(I)$ is Cohen-Macaulay $\operatorname{reg} G(I)= r$. Therefore
$\operatorname{reg} \omega_{G(I)}= d$. If $F(I)$ is Gorenstein, then one can prove
the statement in a similar manner.
\end{proof}
\begin{corollary}
Let $(A,\mathfrak m)$ be a Gorenstein local ring of dimension $d > 0$
and $I$ is an $\mathfrak m$-primary ideal with $G(I)$ is Cohen-Macaulay. Then
$\omega_{F(I)} \cong \bigoplus_{n \in \mathbb Z} \frac{(J^{n+r-d+1}:\mathfrak m
I^r)\cap (J^{n+r-d}:I^r)}{(J^{n+r-d+1}:I^r)}$. If in addition $F(I)$
is Gorenstein, then $$\lambda\left( \frac{(J^{n+1}:\mathfrak m I^r)\cap
(J^{n}:I^r)}{(J^{n+1}:I^r)}\right) = \lambda\left(\frac{I^n}{\mathfrak m
I^n}\right)$$ for all
$n \in \mathbb Z^{+}$.
\end{corollary}
\begin{proof}
By (\cite{hku}, Theorem 4.1), $\omega_{B}= \bigoplus_{n \in \mathbb Z} (J^{n+r}: I^r) t^{n+d-1}$, where
$B= A[It,t^{-1}]$ the extended Rees algebra of $I$. Since $\omega_{G(I)}= \omega_{B/t^{-1}B}= (\omega_B/t^{-1}\omega_B)(-1)$,
we have $\omega_{G(I)}= \bigoplus_{n \in \mathbb Z} \frac{(J^{n+r-d}:I^r)}{(J^{n+r-d+1}:I^r)} $.
From Proposition \ref{pro4}(1),
\begin{eqnarray*}
\omega_{F(I)} &=& \oplus_{n \in \mathbb Z} (0:_{[\omega_{G(I)}]_n} \mathfrak m) \\
&=& \bigoplus_{n \in \mathbb Z} \frac{(J^{n+r-d+1}:\mathfrak m I^r)\cap
(J^{n+r-d}:I^r)}{(J^{n+r-d+1}:I^r)}
\end{eqnarray*}
Assume $F(I)$ is Gorenstein. Then $\omega_{F(I)} \cong F(I)(r-d)= \bigoplus_{n \in \mathbb Z} I^{n+r-d}/\mathfrak m I^{n+r-d}$.
Therefore $$\bigoplus_{n \in \mathbb Z} \frac{(J^{n+r-d+1}:\mathfrak m I^r)\cap (J^{n+r-d}:I^r)}{(J^{n+r-d+1}:I^r)}=
\omega_{F(I)}= \bigoplus_{n \in \mathbb Z} I^{n+r-d}/\mathfrak m I^{n+r-d}.$$ This implies that
$$\lambda\left( \frac{(J^{n+1}:\mathfrak m I^r)\cap
(J^{n}:I^r)}{(J^{n+1}:I^r)}\right) =
\lambda\left(\frac{I^n}{\mathfrak m I^n}\right)$$ for all $n \in \mathbb Z^{+}$.
\end{proof}
\begin{remark} \label{rmk1}
Suppose $G(I)$ and $F(I)$ are Cohen-Macaulay rings. Let
$J=(x_1,\ldots,x_d)$ be a minimal reduction of $I$ such that
$x_1^*,\ldots,x_d^* \in G(I)$ and $x_1^o,\ldots,x_d^o \in F(I)$ are
superficial sequences and hence regular sequences. Denote
$J_i=(x_1,\ldots,x_i)$ and $J_0=(0)$. Then for any $i$ such that $1
\leq i \leq d$, $G(I)/(x_1^*,\ldots,x_i^*)\cong G(I/J_i)$ and
$F(I)/(x_1^o,\ldots,x_i^o) \cong F(I/J_i)$. Then $G(I)$ is Gorenstein if and
only if $G(I/J_i)$ is Gorenstein and
$F(I)$ is Gorenstein if and only if $F(I/J_i)$ is Gorenstein.
\end{remark}
Now we give a characterization for $F(I)$ to be Gorenstein in terms of
certain length conditions involving a minimal reduction of $I$ if $G(I)$
is Gorenstein.
\begin{theorem}\label{gor-char}
Let $(A,\mathfrak m)$ be a Noetherian local ring, $I$ be an $\mathfrak m$-primary ideal
and $J$ be a minimal reduction of $I$ with reduction number $r$.
Assume that $G(I)$ is a Gorenstein ring and $F(I)$ is
Cohen-Macaulay. Then $F(I)$ is Gorenstein if and only if
$$\lambda\left(\frac{((I^{n+1}+J):\mathfrak m)\cap
I^n}{I^{n+1}+JI^{n-1}}\right)=\lambda\left(\frac{I^n}{\mathfrak m
I^n+JI^{n-1}}\right)$$ for $0 \leq n \leq r$.
\end{theorem}
\begin{proof}
Since $G(I)$ and $F(I)$ are Cohen-Macaulay, we may choose a generating
set for $J$ such that the corresponding images form a regular sequence
in $G(I)$ as well as in $F(I)$. Therefore we have
\begin{eqnarray*}
\omega_{F(I/J)} & = & (\omega_{F(I)}/J^o\omega_{F(I)})(r) \\
&= &\bigoplus_{n \in \mathbb Z}\left[\frac{((I^{n+r+1}+J):\mathfrak m)\cap
I^{n+r}+J}{I^{n+r+1}+J}\right].
\end{eqnarray*}
Using the isomorphism theorems and the fact that $G(I)$ is
Cohen-Macaulay, one obtains the isomorphism:
$$
[\omega_{F(I/J)}]_{n-r} \cong \frac{((I^{n+1}+J):\mathfrak m)\cap
I^{n}}{I^{n+1}+JI^{n-1}}.
$$
Suppose $F(I)$ is Gorenstein. Then from the Remark \ref{rmk1}, it
follows that $F(I/J)$ Gorenstein with canonical module
$\omega_{F(I/J)}=F(I/J)(r)$. Therefore
\begin{eqnarray*}
\lambda\left(\frac{((I^{n+1}+J):\mathfrak m)\cap
I^n}{I^{n+1}+JI^{n-1}}\right)& =& \lambda([\omega_{F(I/J)}]_{n-r})\\
& = & \lambda([F(I/J)(r)]_{n-r})=\lambda\left(\frac{I^n}{\mathfrak m
I^n+JI^{n-1}}\right)
\end{eqnarray*}
for all $n$.
Hence, in particular,
we get the required equality of lengths for
$0 \leq n \leq r$.
\vskip 2mm Conversely assume that $ \lambda(((I^{n+1}+J):\mathfrak m)\cap
I^n/I^{n+1}+JI^{n-1})=\lambda(I^n/\mathfrak m I^n+JI^{n-1})$ for $0 \leq n
\leq r$. Then by the above isomorphisms we have
$\lambda([\omega_{F(I/J)}]_n)=\lambda([F(I/J)(r)]_n)$ for all $n$.
This implies that
$\lambda(\omega_{F(I/J)})=\lambda(F(I/J)(r))=\lambda(F(I/J))$. Let $
\eta: P \rightarrow \omega_{F(I/J)}$ be the natural surjective map
from a graded free $F(I/J)$-module $P$ of rank equal to
$\mu(\omega_{F(I/J)})$, the minimal number of generators of
$\omega_{F(I/J)}$. Since $\omega_{F(I/J)}$ has finite injective
dimension, its injective dimension is equal to $\operatorname{depth} F(I/J) = 0.$
Therefore $\omega_{F(I/J)}$ is an injective module and hence
$\operatorname{Hom}_{F(I/J)}(-,\omega_{F(I/J)})$ is an exact functor. Applying
this exact functor to the map $\eta,$ we get a surjective map
$\eta^*:\operatorname{Hom}_{F(I/J)}(\omega_{F(I/J)},\omega_{F(I/J)})\longrightarrow
\operatorname{Hom}_{F(I/J)}(P,\omega_{F(I/J)})$. But by the definition of
canonical module,
$$\operatorname{Hom}_{F(I/J)}(\omega_{F(I/J)},\omega_{F(I/J)})\cong F(I/J).$$ Note
that $$\operatorname{Hom}_{F(I/J)}(P,\omega_{F(I/J)})\cong \bigoplus_{rank(P)}
\omega_{F(I/J)}.$$ Hence there is a surjective map
$F(I/J)\rightarrow \bigoplus_{rank(P)} \omega_{F(I/J)}$. This
implies that
$$\lambda\left(\bigoplus_{rank(P)} \omega_{F(I/J)}\right) \leq
\lambda(F(I/J)).$$ This gives that $rank(P) \cdot
\lambda(\omega_{F(I/J)}) \leq \lambda(F(I/J))$. Since
$\lambda(\omega_{F(I/J)})=\lambda(F(I/J))$, the above inequality
gives that $rank(P)=1$. That is $\mu(\omega_{F(I/J)})=1$. Hence
$F(I/J)$ is Gorenstein with canonical module
$\omega_{F(I/J)}=F(I/J)(r)$. Since $J^o$ is generated by a regular sequence in
$F(I)$, $F(I)$ is Gorenstein with canonical module
$\omega_{F(I)} = F(I)(r-d)$.
\end{proof}
It is known that if $G(I)$ is Gorenstein, then $A$ is Gorenstein and
that such an analogue is not true in the case of fiber cone. We
show that if, in addition, $F(I)$ is Gorenstein, then $A/I$ is
Gorenstein.
\begin{corollary} \label{cor6}
Let $(A,\mathfrak m)$ be a Noetherian local ring and $I$ is an $\mathfrak m$-primary
ideal. Suppose $G(I)$ and $F(I)$ are Gorenstein. Then $A/I$
is Gorenstein.
\end{corollary}
\begin{proof}
Put $n=0$ in the Theorem \ref{gor-char}, we get $\lambda((I:\mathfrak m)/I)=\lambda(A/\mathfrak m)=1$.
This implies that $A/I$ is Gorenstein.
\end{proof}
Let $\operatorname{pd}_A(M)$ denote the projective dimension of $M$ as an $A$-module.
\begin{corollary} \label{cor7}
Suppose $I$ is an $\mathfrak m$-primary ideal such that $G(I)$ and $F(I)$ are Gorenstein.
Then $\operatorname{pd}(A/I) < \infty$ if and only if $I$ is generated by a regular
sequence.
\end{corollary}
\begin{proof}
Since $G(I)$ and $F(I)$ are Gorenstein, from the Corollary
\ref{cor6} it follows that $A/I$ is Gorenstein. Suppose $\operatorname{pd}(A/I)<
\infty$. Then from Theorem 2.6 of \cite{nv}, $I$ is generated by
a regular sequence. Conversely assume $I$ is generated by a regular
sequence. Then the Koszul complex $K_{\bullet}(I;A)$ is minimal free
resolution of $A/I$. Therefore $\operatorname{pd}(A/I)< \infty$.
\end{proof}
\begin{corollary}
Suppose $(A,\mathfrak m)$ is a regular local ring and $I$ is an $\mathfrak m$-primary ideal of $A$ such that $G(I)$ and
$F(I)$ are Gorenstein. Then $I$ is generated by a regular sequence.
\end{corollary}
\begin{proof}
By hypothesis $A$ is regular, so we have
$\operatorname{pd}(A/I) < \infty$. Therefore from Corollary \ref{cor7}, $I$ is
generated by a regular sequence.
\end{proof}
\begin{remark}
Suppose $G(I)$ and $F(I)$ are Gorenstein rings and $\mu(I)= d+2$. Then by Corollary
\ref{cor6}, $A/I$ is Gorenstein. By (\cite{hku}, Remark 2.9(1),
(2)) we have $\operatorname{pd}_A(G(I)) < \infty$ if and only if $G(I)$ is a
complete intersection. That is the defining ideal of $G(I)$ is
generated by a regular sequence. In this case $F(I)$ is also a
complete intersection.
\end{remark}
The expression we have obtained for the canonical module of the fiber
cone does not reveal when it can be realized as a submodule of
$F(I)$. Below we give a sufficient condition for $\omega_{F(I)}$
to be a submodule of $F(I)$.
\begin{corollary}
Let $(A, \mathfrak m)$ be a Noetherian local ring and $I$ be an $\mathfrak m$-primary
ideal. Assume that $G(I)$ is a
Gorenstein ring. Suppose $F(I)$ is Cohen-Macaulay and $A/I$ is Gorenstein.
If $(I^{n+r-d+1}:\mathfrak m) \cap \mathfrak m I^{n+r-d}= I^{n+r-d}$ for all $n \geq d-r+1$,
then $\omega_{F(I)}$ is an ideal of $F(I)$.
\end{corollary}
\begin{proof}
For any $n \geq d-r$, there is a natural map
$\psi_n:(I^{n+r-d+1}:\mathfrak m) \cap I^{n+r-d}/I^{n+r-d}\longrightarrow
I^{n+r-d}/\mathfrak m I^{n+r-d}$ which gives rise to a natural $F(I)$-linear
map $\psi : \omega_{F(I)} \longrightarrow F(I)$. The kernel of
$\psi_n$ is $(I^{n+r-d+1}:\mathfrak m) \cap \mathfrak m I^{n+r-d}/I^{n+r-d}$. Then the
$F(I)$-linear map $\psi:\omega_{F(I)}\longrightarrow F(I)(r-d)$ has
kernel $\oplus_{n \geq d-r}(I^{n+r-d+1}:\mathfrak m) \cap \mathfrak m
I^{n+r-d}/I^{n+r-d}$. Then $\omega_{F(I)}$ is an ideal of $F(I)$ if
$ker(\psi)=0$, i.e., if $(I^{n+r-d+1}:\mathfrak m) \cap \mathfrak m I^{n+r-d}=
I^{n+r-d}$ for all $n \geq d-r+1$. This completes the proof.
\end{proof}
The following Proposition shows that $e_0(\omega_{F(I)}) <
e_0(\omega_{G(I)})$ unless $I = \mathfrak m$.
\begin{proposition}
Let $(A, \mathfrak m)$ be a Noetherian local ring and $I$ be an $\mathfrak m$-primary
ideal. Suppose $G(I)$ and $F(I)$ are Cohen-Macaulay. Then
$e_0(\omega_{F(I)}) \leq e_0(\omega_{G(I)})$. Furthermore, if $G(I)$
is Gorenstein, then equality holds if and only if $I=\mathfrak m$.
\end{proposition}
\begin{proof}
From Proposition \ref{pro4}(1), it is clear that
$\lambda([\omega_{F(I)}]_n) \leq \lambda([\omega_{G(I)}]_n)$ for all
$n$. Hence $e_0(\omega_{F(I)}) \leq e_0(\omega_{G(I)})$. Let
$J=(x_1,\ldots,x_d)$ be a minimal reduction of $I$ with reduction
number $r$ such that $x_1^o,\ldots,x_d^o \in F(I)$ is a regular
sequence for $F(I)$, $\omega_{F(I)}$ and $x_1^*,\ldots,x_d^* \in
G(I)$ is $G(I)$ and $\omega_{G(I)}$ regular sequence. Then
$e_0(\omega_{G(I)})=e_0(\omega_{G(I/J)})=\lambda(\omega_{G(I/J)})$
and
$e_0(\omega_{F(I)})=e_0(\omega_{F(I/J)})=\lambda(\omega_{F(I/J)})$.
Now assume $e_0(\omega_{F(I)}) = e_0(\omega_{G(I)})$. Then
$\lambda(\omega_{F(I/J)}) = \lambda(\omega_{G(I/J)})$. By
Proposition \ref{pro4}(1) we have $\omega_{F(I/J)} \subseteq
\omega_{G(I/J)}$. Therefore $\omega_{F(I/J)} =\omega_{G(I/J)}$.
This implies that $(I:\mathfrak m)/I=A/I$. This gives
$\lambda((I:\mathfrak m)/I)= \lambda(A/I)$. This implies $(I:\mathfrak m)=A$. That is
$I=\mathfrak m$. The converse always holds. This completes the proof.
\end{proof}
We conclude this article by providing two examples. First we give an
example of an ideal whose associated graded ring is Gorenstein but the
fiber cone is not.
\begin{example}
Let $A=k[\![t^4,t^9,t^{10}]\!]$ and $I=(t^8,t^{18},t^{10}),\mathfrak m =
(t^4,t^9,t^{10})$. Then $A$ is a one dimensional Gorenstein local
domain and $I$ is an $\mathfrak m$-primary ideal. $J=(t^8)$ is a minimal
reduction of $I$ of reduction number $1$. Then it follows from the
Corollary 4.5 (5) of \cite{hku} that $G(I)$ is Gorenstein. Since $I$
has reduction number $1$, $F(I)$ is Cohen-Macaulay. Since $\mu(I) = 3
> \dim A + 1$, by Proposition 4.1 of \cite{jpv}, $F(I)$ is not
Gorenstein. Also, note that
$$
5 = \lambda\left (\frac{((I^2+J):\mathfrak m) \cap I}{I^2+J} \right ) \neq
\lambda \left ( \frac{I}{\mathfrak m I+J} \right )= 1.
$$
\end{example}
In the example below we apply our result to obtain an example of a
Gorenstein fiber cone.
\begin{example}[\cite{hku}, Example 2.5]
Let $A=k[\![t^4,t^9,t^{10}]\!]$ and $I := (t^8,t^9,t^{10})$.
Then $A$ is a one dimensional Gorenstein local
domain, $I$ is an $\mathfrak m$-primary ideal and $J=(t^8)$ is a minimal
reduction of $I$ with reduction number $2$. Let $R=A/I[X,Y,Z]$. Then
$G(I) \cong \frac{R}{(XZ-Y^2, wX^2-Z^2)}$, where $w$ is the image of
$t^4$ in $A/I$. Since $XZ-Y^2, wX^2-Z^2$ is an $R$-regular sequence
and $A/I$ is Gorenstein therefore $G(I)$ is Gorenstein. Since $\mathfrak m I^n
\cap J = \mathfrak m JI^{n-1}$ for all $n \geq 1$, $F(I)$ is Cohen-Macaulay.
By using CoCoA, \cite{co}, it can easily be seen that
$$\lambda\left (\frac{((I^2+J):\mathfrak m) \cap I}{I^2+J} \right )=
2 = \lambda \left ( \frac{I}{\mathfrak m I+J} \right ) $$ and $$\lambda\left
(\frac{((I^3+J):\mathfrak m) \cap I^2}{I^2+JI} \right )= 1= \lambda \left (
\frac{I^2}{\mathfrak m I^2+JI} \right ).$$ Therefore by the Theorem
\ref{gor-char}, $F(I)$ is Gorenstein.
\end{example}
|
\section{Introduction: The power of dualities}
\label{sec1}
The term {\it duality} is pervasive in physics, mathematics and
philosophy. In general, a duality connects and contrasts two aspects or
realizations of a given entity. In this article we will be concerned
only with dualities in physics. These are specific mathematical
transformations that we will uncover as we proceed connecting seemingly
unrelated {\it physical phenomena}.
Over time, dualities have appeared in various guises in nearly all
disciplines of physics \cite{witten}. The electromagnetic (EM) duality
of Maxwell's equations in the absence of sources, noticed by Heaviside
in 1884, is probably the oldest well-known duality in modern physics.
Later, the wave-particle duality of quantum mechanics \cite{dirac}
became a fundamental tenet of the modern physical description of
reality. This Fourier transform-based duality has since appeared in
numerous arenas, including in recent years various branches of quantum
statistical mechanics and field theory (see, for example,
\cite{shastry}), and it is likely to continue to play an ever
increasing role. Kramers and Wannier introduced dualities in statistical
mechanics in their foundational 1941 paper \cite{KW}. These authors
discovered an elegant relation between the two-dimensional classical
Ising model on a square lattice {\it at high temperature}, and the same
model {\it at low temperature} (hence the origin of the name
``self-duality" in this case), and used it to determine that model's
{\it exact critical temperature} some years before Onsager
\cite{onsager} published its exact solution. This first {\it
quantitative} success was followed by other similar ones, and so
dualities became a standard tool in statistical mechanics since they
could also provide {\it qualitative} insight. The spectacular
cross-fertilization between statistical mechanics and quantum field
theory (QFT) of the 1970's brought dualities to the attention of high
energy theorists, and soon it became apparent that dualities in QFT
combined features of the EM and statistical mechanics dualities while
retaining their distinct capability to produce weakly-coupled
representations of strongly-coupled problems.
Dualities can provide reliable qualitative or even exact quantitative
information about systems that need not be exactly solvable, partly
because they can put constraints on the phase boundaries and the exact
location of some critical or multi-critical points. Thus
(self-)dualities have been essential for investigating the phase
diagrams of numerous models in statistical mechanics and field theory
\cite{savit,malyshev}. This aspect encompasses some of the most
spectacular applications of dualities, and constitutes the legacy of
Kramers and Wannier. However, dualities can become even more potent when
fused with other tools, such as perturbation theory. A case in point is
the AdS-CFT correspondence, a topic that remains the focus of intense
research efforts. Since its original formulation in high energy physics
\cite{maldacena, GKP, AdS-Witten, AdS-review}, this {\it conjecture of
a duality} between a weakly-coupled five-dimensional gravity theory (on
an {\it Anti de Sitter} (AdS) background) and a strongly-coupled
four-dimensional conformal field theory (CFT) has been generalized and
exploited in other branches of physics. At present, the range of
applications of the AdS-CFT correspondence includes problems as diverse
as electronic transport properties \cite{electron}, quantum critical
dynamics \cite{understanding_qpt}, and the physics of strongly coupled
quark-gluon plasmas \cite{hydro}. The efforts to apply the AdS-CFT
correspondence to other strongly-coupled models continue, but the
problem of pushing it beyond a conjecture into a rigorous mathematical
statement remains. We think that turning the AdS-CFT correspondence
into a mathematically rigorous duality is essential to understanding
its potential generalizations.
These examples (old and new) attest to the power of dualities and
justifies the efforts of numerous researchers to exploit them to address
hard problems by simple, elegant means. This article is a self-contained
exposition of extensive original developments, including many new
dualities and self-dual models, in the general bond-algebraic theory
of dualities first introduced in Reference \cite{con} (bond algebras
have also been employed in the analysis and spectral resolution of
exactly solvable models \cite{bondDec08,OQC}, but this application is
not discussed in this paper). In the context of some specific models,
quantum dualities have been well understood for many years. Reference
\cite{con} introduced a coherent framework supporting a systematic study
of both quantum {\it and} classical dualities on an equal footing, and
providing a systematic way to compute dual (disorder) variables. The
new, unified, theory that emerges is rigorous, easy to use,
algorithmic, and of practical significance to theoretical studies and
numerical simulations. It has the potential to provide invaluable
insight into a myriad of pressing problems that are beyond perturbation
theory. Our theory of dualities rests on a single key observation: The
{\it bonds or interactions} of a Hamiltonian or transfer matrix are
more relevant to a duality transformation than the elementary degrees
of freedom. Those bonds, or interaction terms, are organized into a
{\it bond algebra}. In contrast, symmetries or properties of the
elementary degrees of freedom are largely irrelevant from a duality
mapping perspective \cite{bondDec08,con}. Our bond-algebraic approach
to dualities may shed light on problems like the characterization and
classification of collective (topological) excitations of lattice
models, and the AdS-CFT correspondence \cite{noc}, when complemented
with recent developments in the theory of dimensional reduction
\cite{batista_nussinov, tqo}.
Hamiltonians that meet basic physical requirements like causality are
usually realized by adding together (or integrating with a given
measure in QFT) simple local or quasi-local operators, like few-body
interactions and kinetic energy single-body operators. These operators
(or some simple function of them) constitute the {\it bond operators},
or bonds for short. Together they generate, in a sense to be specified
in Section \ref{sec3.2}, an operator von Neumann algebra that we call a
{bond algebra}. Bond algebras are model-specific algebras of
interactions that can be {\it radically different} from the algebras
that embody elementary degrees of freedom like bosons, fermions, or
spins. Our new theory states that dualities are structure-preserving
mappings (homomorphisms) of bond algebras, {\it that are typically local
in the bonds}. This {\it purely algebraic} characterization is
physically meaningful because homomorphism of bond algebras are {\it
always} equivalent to a unitary transformation (or, if gauge symmetries
are involved, to a partial isometry) connecting the Hilbert spaces on
which the two models connected by a duality are defined. Thus quantum
dualities are revealed as unitary transformations, and dual models must
share identical spectra and level degeneracies; that is, provided gauge
symmetries are not involved. The precise connection between gauge
symmetries and dualities will be one of the central themes of this
paper.
Remarkably, these ideas can be extended to include classical dualities
\cite{con}, thus unifying the theory of classical and quantum
dualities in a way that had been overlooked up to now. The key is to
notice that many problems in classical statistical physics can be
formulated in terms of a transfer matrix or operator. In this context,
physical requirements like locality become manifest in the {\it
multiplicative} structure of the transfer matrix or operator, that will
be in general a {\it product} of local or quasi-local bonds. Once the
bonds are identified, the theory of dualities proceeds as before. Thus,
a duality mapping for the bond algebra of the transfer matrix realizes a
dual transfer matrix that determines a dual representation \(\mathcal{Z}^D\) of
the partition function \(\mathcal{Z}\). But, since dualities are unitary
transformations, it follows that \(\mathcal{Z} \propto \mathcal{Z}^D\). This is, in
short, the bond-algebraic approach to classical dualities. In this way
one can show, for instance, that the self-duality of the two-dimensional
classical Ising model \cite{KW, bookEPTCP}, and the self-duality of the
quantum Ising chain \cite{fradkin_susskind} are two manifestations of a
symmetry possessed by one and the same bond algebra. But more
importantly, we will illustrate in this paper that the bond-algebraic
approach has the potential to extend classical dualities beyond its
present boundaries, as set for example in the reviews
\cite{savit,malyshev} (see Appendix \ref{appA}). We think that this
strongly indicates that our technique can be relevant to solving the
open problem of constructing dual representations of non-Abelian
lattice models, although so far we have only some examples to support
this claim.
Bond algebras are most efficient in revealing dualities because they are
model (Hamiltonian/transfer matrix)-specific, and thus capture, in the
form of a local mapping, the precise information that makes two
operators dual. This is different from techniques based on essentially
fixed mappings, like generalized Jordan-Wigner transformations
\cite{GJW}. These transformations are isomorphisms (that we call {\it
dictionaries}) connecting operator algebras (i.e., {\it languages}),
typically representing elementary degrees of freedom. Since these
dictionaries are insensitive to the specific structure of physical
models, they can easily spoil important physical characteristics, as
illustrated by standard fermionization in more than one space
dimension. However, it is important to underscore that a bond algebra
is {\it not} uniquely determined by the Hamiltonian/transfer matrix
that motivates its definition, but rather by the choice of bond
decomposition of those operators. Since that choice is not unique, (i)
any single Hamiltonian may produce many different bond algebras, and
(ii) any one bond algebra may be related to many different Hamiltonians.
This flexibility is essential for classification purposes and to the
success of our technique. It enables us to explore a variety of bond
algebras for one and the same Hamiltonian/transfer matrix to capture its
full range of dual representations. Similarly, it allow us to apply one
and the same duality mapping to different problems. The self-duality
for the spin \(1/2\) XY model in a transverse field of Section
\ref{sec3.10.2} provides an example of the first observation (i)
above.
In this case, the self-duality {\it emerges} on a special line in
coupling space, and it takes a special choice of bonds to generate the
bond algebra that makes this self-duality apparent. The second
observation (ii) is illustrated by the new dualities of Section
\ref{sec3.6} for the quantum Heisenberg model in any number of space
dimensions \(d\). These dualities for a {\it non-Abelian} model are
based on dualities for the (Abelian!) $d$-dimensional quantum Ising
model and highlight some difficulties with the standard notion of
non-Abelian dualities \cite{savit}. They point to the fact that the
concepts of symmetry and duality are quite often inaccurately related.
While symmetries of a theory represent isomorphisms leaving the
Hamiltonian invariant, (self-)dualities do not preserve the form of
the theory, but rather preserve its spectrum and level degeneracies.
The relevance of bonds over elementary degrees of freedom is further
emphasized by the fact that dualities are {\it local} transformations in
terms of bonds, while they are {\it non-local} when described in terms
of elementary degrees of freedom. This follows because one can invert
the relations between bonds and elementary degrees of freedom to obtain
the latter as non-local functions of the bonds. Once this is done, one
can compute the action of the duality mapping on the elementary degrees
of freedom, to obtain their dual image. Since dualities are
structure-preserving mappings, these {\it dual} elementary degrees of
freedom are guaranteed to be equivalent to the original ones. This {\it
systematic derivation} of (non-local in general) dual variables
(elementary degrees of freedom) from more basic, local objects is
extremely important, not only because it establishes a bridge between
the bond-algebraic and the traditional approach to quantum dualities,
but also because dual variables can have a fundamental physical meaning
as (generalized) {\it disorder variables}
\cite{Kadanoff_Ceva,fradkin_susskind}.
Starting with any specific bond algebra, one can proceed to look
systematically for its alternative realizations that feature local
representations of the bond generators. {\it Any of these realizations
defines a duality} (realizations that feature non-local representations
of the bond generators can appear, but would typically be discarded in
practice). This simple premise can lead to surprising results, as we
will see often in this paper. Exact dimensional reduction can appear as
a duality, and, as illustrated by our new derivation of the
Jordan-Wigner transformation, {\it statistical transmutation} can
appear as a duality too. {\it Symmetry transmutation} is also not
unusual as illustrated for example by our new derivation of the
duality between the XY and solid-on-solid (SoS) models, obtained here
for the first time {\it without invoking the Villain approximation}
({symmetry transmutation} here refers to the fact that the \(U(1)\)
symmetry group of the XY model is not isomorphic, but rather the
Pontryagin dual of the \(\mathbb{Z}\) symmetry group of the SoS model).
But perhaps what is most surprising about bond-algebraic dualities is
their capability to {\it dispose of gauge symmetries}, as we explain
next.
Gauge symmetries are constraints on the state Hilbert space, encoded in
a large set of local operators that commute with the Hamiltonian, and
with any other (measurable) observable. Since bonds are in general
measurable, local observables, they generate a bond algebra of
gauge-invariant operators. This has the remarkable consequence that bond
algebras of gauge models can have dual representations on state spaces
of lower dimensionality (a fact that should not be confused with an
{\it emergent} duality as described in Section \ref{sec3.11}). We call
these dualities gauge-reducing; they map the original gauge model to a
model with less, or simply without gauge symmetries, and are represented
by {\it homomorphisms} of bond algebras, rather than isomorphisms.
Unlike ordinary dualities that are unitarily implementable,
gauge-reducing dualities are implemented by partial isometries (one
can think of a partial isometry as a rectangular unitary matrix, that
either maps a state to zero or to an isometric (equal norm) state). But
in spite of these mathematical twists, dualities are just as easy to
detect in gauge theories as in any other model. That is because, while
gauge symmetries do affect the algebra of elementary degrees of
freedom, {\it they do not affect the algebraic relations between bonds}.
Thus bond-algebraic dualities may provide in some instances a practical
solution to the problem of eliminating gauge constraints. This paper
describes in detail the mathematics and multiple applications of these
ideas to models with {\it Abelian} gauge symmetries. {\it Non-Abelian}
gauge symmetries abide by the same principles but are technically more
involved, so we defer their complete treatment to future publications.
This article contains many new results and is organized as follows.
Section \ref{sec2.1} starts with a discussion of the different physical
contexts in which dualities have been introduced, followed in Section
\ref{sec2.2} by a more detailed discussion of what was known about
quantum dualities prior to the publication of Reference \cite{con}, the
so-called standard approach. These discussions should help the reader
put the subject of dualities, and some of the problems they address, in
perspective. Section \ref{sec3} contains all the important formal
developments related to the bond-algebraic approach to dualities.
Overall it is devoted to quantum dualities and their intimate
connection to classical dualities, mainly discussing every major new
idea and mathematical technique that is of relevance. We start by
studying bond algebras in Sections \ref{sec3.2} and \ref{sec3.3}. Our
new definition of quantum dualities as mappings of bond algebras is
presented in Section \ref{sec3.4}, and the next section explains how the
standard approach described in Section \ref{sec2.2} follows immediately
from the bond-algebraic formalism (determination of dual variables).
Section \ref{sec3.6} discusses critically the notion of non-Abelian
duality, in the light of new dualities for the quantum Heisenberg model,
and Section \ref{sec3.7} develops new boundary conditions that preserve
duality properties {\it in finite size systems}.
Section \ref{sec3.8} expands, in more concrete
terms, on the connection between bond algebra mappings and
unitary transformations. The remaining sections are devoted to deriving
and explaining the precise relations between self-dualities, standard
quantum symmetries, Section \ref{sec3.9}, and disorder variables,
Section \ref{sec3.10}, and the novel concept of emergent duality,
Section \ref{sec3.11}. Disorder variables can be systematically
determined from bond-algebraic dual mappings and they share an
interesting relation with topological excitations. Section
\ref{sec3.12} explains how bond-algebraic dualities afford a practical
way to eliminate gauge symmetries completely, and finally Section
\ref{classical&quantum} explains how many classical dualities follow from
bond-algebraic techniques.
The rest of the article is dedicated to unveil old and discover new
dualities in a broad spectrum of problems of physical relevance. There
are so many examples that in the following we only indicate a few of
them. Sections \ref{sec4} and \ref{sec5} exemplify the many ideas ideas
and techniques developed in Section \ref{sec3} with self-dualities and
dualities in quantum lattice models, of arbitrary spatial dimensions,
mainly of interest in condensed matter physics. For example, Section
\ref{sec5.5} describes a new duality for the extended Kitaev toric code
model in arbitrary space dimensions to a well known model of
Hamiltonian lattice field theory, the \(\mathbb{Z}_2\) Abelian Higgs
model \cite{fradkin_shenker}. The application of bond-algebraic
techniques to QFT is developed in Section \ref{sec6}.
Sometimes we show results directly in the continuum but we want to
stress the fact that the lattice Hamiltonian formalism is the most
convenient approach for interacting field theories. We study in full
detail compact and non-compact versions of quantum electrodynamics in
various spatial dimensions and make an attempt to introduce a version
of quantum electrodynamics without vector potential. In Section
\ref{sec6.4} we introduce a new family of self-dual models related to
the Abelian Higgs model, we prove the quantum St\"uckelberg model to be
self-dual in two-dimensions in Section \ref{sec_stuckel}, and Section
\ref{6.8} discusses self-dual field theories that display the
phenomenon of {\it dimensional reduction} of prime interest in the
theory of topological quantum order \cite{tqo}.
Section \ref{sec8} presents several problems of classical statistical
mechanics whose duality properties are uncovered by our bond-algebraic
approach. Section \ref{sec8.2} describes duality properties of the
Ising model in the Utiyama lattice, while the vector Potts model in two
dimensions is not only studied in great detail in Section \ref{sec8.3}
but also it is shown how to modify it to make it self-dual for
arbitrary couplings and states. Building upon these ideas we introduce a
new self-dual \(p\)-state approximation to four-dimensional lattice
electrodynamics in Section \ref{sec8.4}.
Section \ref{eightvertexM} establishes a link between the eight-vertex
model (in its Ashkin-Teller representation) and the quantum anisotropic
Heisenberg model, based on duality mappings rather than integrability.
The general connection between
quantum and classical disorder variables is discussed in Section
\ref{appB}. Finally, Section \ref{sec9} presents several important
applications and consequences of dualities. Those include general
properties of the spectrum of self-dual theories and the way to extract
exact relevant information in presence of phase transitions, and most
importantly a new way to look at {\it fermionization in arbitrary
spatial dimensions} as a general duality mapping. In particular, we
demonstrate how the Jordan-Wigner transformation is a consequence of bond
algebraic duality mappings when these are applied to nearest-neighbor
spinless fermion and spin $S=1/2$ systems on a chain. We further show,
how {\em no such duality can map} all (real-space) nearest-neighbor
spinless Fermi hopping terms to all spin $S=1/2$ exchange terms on
general lattices (and viceversa) in spatial dimensions $d>1$. That is,
in general, no extension of the $d=1$ Jordan-Wigner transformation that
connects all such individual {\it local terms} can appear in higher
dimensions. We show, however, that notwithstanding it is possible to
fermionize spin systems in $d>1$ dimensions (such as the $d=2$ quantum
Ising model) via a gauge reducing duality. Thus we further extend and
formalize, via the systematics of bond algebras, the scope of known
systems that can be fermionized. In Section \ref{sec_tqo_dim} we examine
the connection between the phenomenon of dimensional reduction and
duality in the context of topological quantum order systems \cite{tqo}.
We also indicate the most natural way to classify these systems in terms
of so-called $d$-dimensional gauge-like symmetries as opposed to using
a particular measure of entanglement since no single measure can
uniquely characterize a quantum state.
The appendices, a total of five, contain not only technical details
omitted in the text for pedagogical reasons but also some relevant
mathematical expressions not found in the literature, so we advice the
reader to consider consulting them if he or she wants to get a deeper
understanding of the subject.
\section{Traditional approaches to dualities}
\label{sec2}
This short section discusses dualities in physics from a historical
perspective. Mainstream, traditional ideas about dualities in
field theory, classical statistical mechanics and quantum mechanics
are summarized and illustrated with key examples.
\subsection{Dualities in perspective: What is a duality?}
\label{sec2.1}
The first appearance of the term {\it duality}
can be traced back to the early days of
electromagnetism as a field theory.
In 1884, Heaviside recast Maxwell's equations in vector form
\cite{nahin}, in the absence of sources,
\begin{eqnarray}
\nabla\cdot \vec{E}&=&0,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \,
\nabla\cdot\vec{B}=0,\label{maxwell's}\\
\nabla\times\vec{E}&=&-\partial_t\vec{B},\ \ \ \ \ \ \ \ \ \ \ \
\nabla\times\vec{B}=\partial_t\vec{E}\nonumber
\end{eqnarray}
(in rationalized Heaviside-Lorentz units, and the speed of
light $c=1$), and pointed out that the {\it duality}
\begin{equation}\label{em_duality}
\vec{E}\ \rightarrow\ \hat{\vec{E}}=\vec{B},\ \ \ \ \ \ \ \ \ \ \ \
\vec{B}\ \rightarrow\ \hat{\vec{B}}=-\vec{E}
\end{equation}
maps solutions of these equations to other solutions.
This notion of duality stresses {\it symmetries of
the equations of motion} showing that different
physical quantities are interchangeable.
Two of Maxwell's equations can be solved by introducing a four-vector potential
\(A_\mu\) (\(\mu,\nu=0,1,2,3\)), that is related to $\vec{E}=(E_1,E_2,E_3)$
and $\vec{B}=(B_1,B_2,B_3)$ through the antisymmetric field-strength tensor
\begin{equation}
F_{\mu \nu}=\partial_\mu A_\nu-\partial_\nu A_\mu,
\ \ \ \ \ \
F_{\mu \nu}=
\begin{pmatrix}
0& E_1& E_2& E_3\\
-E_1& 0& -B_3& B_2\\
-E_2& B_3& 0& -B_1\\
-E_3& -B_2& B_1& 0
\end{pmatrix} .
\end{equation}
Then the remaining Maxwell's equations read (\(\eta={\sf diag}(1,-1,-1,-1)\))
\begin{equation}\label{meqa}
\eta^{\mu\nu}\partial_\mu F_{\nu \gamma}=(
\eta^{\mu\nu}\partial_\mu\partial_\nu)A_\gamma-\partial_\mu(\eta^{\nu\gamma}
\partial_\nu A_\gamma)=0\ .
\end{equation}
This formalism is essential for well know reasons,
but the (self-)duality of Maxwell's equations
is not readily reflected by Equation \eqref{meqa}, partly
because the gauge transformation, with arbitrary scalar function $\chi({\bm{x}})$,
\begin{equation}
A_\mu({\bm{x}}) \mapsto {A'}_\mu({\bm{x}})= A_\mu({\bm{x}})+\partial_\mu \chi({\bm{x}}),
\end{equation}
leaves the EM field \(F_{\mu \nu}\) unchanged but renders \(A_\mu\)
unobservable. This motivates one of this paper's central theme: the
interplay between dualities and gauge symmetries.
A second, seemingly very different, concept of duality was developed
in the early 1940s in the context
of {\it classical} statistical mechanics. It originated in
the work of Kramers and Wannier \cite{KW} on the
two-dimensional ($D=2$) Ising model on a square lattice,
with partition function
\begin{equation}
\mathcal{Z}_{\sf I}(K)=\sum_{\{\sigma_{\bm{r}}\}}\ \exp \Bigl[K
\sum_{\bm{r}}\sum_{\nu=1,2}\ \sigma_{{\bm{r}}+\bm{e_\nu}}\sigma_{\bm{r}}\ \Bigr] .
\end{equation}
Here $\sigma_{\bm{r}}=\pm 1$ stands for a classical spin degree of freedom
located on the vertex (site) ${\bm{r}}=(r^1,r^2)=r^1 \bm{e_1} +r^2 \bm{e_2}$, $r^1,r^2\in
\mathbb{Z}$, of a
square lattice with unit vectors $\bm{e_1}, \bm{e_2}$, see Figure
\ref{notation_links}. The exchange constant is $J=K k_BT$, with $T$
temperature and $k_B$ Boltzmann's constant.
Kramers and Wannier noticed that \(\mathcal{Z}_{\sf I}\) satisfies a relation
\begin{equation}\label{first_time_statistics}
\mathcal{Z}_{\sf I}(K)=A(K,K^*)\mathcal{Z}_{\sf I}(K^*),
\end{equation}
that, since
\begin{equation}\label{critical_K_ising}
K^*=-\frac{1}{2}\ln \tanh K,
\end{equation}
connects the high-temperature (weak \(K\)) behavior of the Ising
model to its low temperature (strong $K$) behavior
(\(A(K,K^*)\) is a known, analytic (non-singular) proportionality factor).
It follows that every singularity
of \(\mathcal{Z}_{\sf I}\) at coupling \(K\) must be matched
by another singularity at a corresponding dual coupling \(K^*\), so that if
\(\mathcal{Z}_{\sf I}\) has only one phase transition (one singularity) at
a critical coupling \(K_c\), it must be located at the self-dual point
\(K_c=K_c^*\). Then from Equation \eqref{critical_K_ising}
\begin{equation}
K_c=\frac{1}{2}\ln(1+\sqrt{2}),
\end{equation}
which determines the {\it exact} critical temperature of the Ising model
$T_c=2/\ln(1+\sqrt{2})$ (in units of $J/k_B$).
This spectacular result drew considerable attention to duality
transformations,
and soon after the publication of Reference \cite{KW}, Wannier
\cite{wannier} showed how to obtain duality relations of the form
\begin{equation}\label{genidclass}
\mathcal{Z}_{\Lambda}(K)=A(K,K^*)\mathcal{Z}_{\Lambda^*}(K^*),
\end{equation}
for Ising models defined on nearly arbitrary planar lattices \(\Lambda\).
These transformations where dubbed dualities \cite{wannier}
because \(\Lambda^*\) stands for the lattice dual to \(\Lambda\)
({\it dual} here is used in the sense of old algebraic topology, see for
instance, Reference \cite{bredon}, Chapter IV, Section 6.)
Dual lattices are defined in Appendix \ref{appA}.
We see from Equation \eqref{genidclass} that the Ising model on a square
lattice is self-dual partly because so is the square lattice,
while the Ising model on a triangular lattice is dual to that same model
on a hexagonal lattice (and viceversa). This last duality
is part of a simple approach to determining these models' critical
temperature \cite{baxter}.
Potts extended the duality transformation of Wannier to some of the models now known
under his name in his 1952 paper \cite{potts}.
It was Wegner, however, who uncovered the general structure underlying
dualities, by showing in his 1973 paper
\cite{wegner} that duality transformations
could be obtained for a wide class of models
(including models in more than two dimensions) by considering the (group-theoretical)
Fourier transform of individual Boltzmann weights.
The 1975 paper \cite{wu_wang} popularized Wegner's approach to
the point that it became the standard, traditional approach to classical dualities.
This traditional approach is described thoroughly in the
review articles by Savit \cite{savit} and Malyshev and Petrova
\cite{malyshev}, but it is also summarized in Appendix \ref{appA}
(see Reference \cite{bookEPTCP} for a pedagogical introduction).
The key fact that the notion of a Fourier transform exists for {\it any}
group allows for application of the traditional approach to produce dual representations
of any {\it Abelian} model displaying appropriate interactions, but it is impossible
to take the same traditional approach to produce dual representations
for lattice {\it non-Abelian} models (that is, models like
Wilson's non-Abelian lattice gauge theories). To construct dual representations
of such models remains one of the most difficult open problems in the
theory of dualities \cite{savit} (let us point out that {\it Abelian} and
{\it non-Abelian} here does not refer to the symmetry group of the model of
interest, but rather to the nature of that model's degrees of freedom,
see Section \ref{sec3.6}).
Yet a third concept of duality was developed for
{\it quantum} many-body problems in the late 1970s, mainly in a series
of papers \cite{fradkin_susskind,horn_yankielowicz,horn,kogut,shankar}
concerned with Hamiltonian lattice
quantum field theory (LQFT). Starting with Reference \cite{fradkin_susskind},
these papers settled a notion of
quantum duality that we interpret and formalize as follows:
two Hamiltonians
\begin{equation}
\label{h1h2ok}
H_1[\lambda_1,\lambda_2,\cdots](o_\Gamma),\ \ \ \ \ \ \ \ \
H_2[\lambda_1,\lambda_2,\cdots](o_\Gamma),
\end{equation}
that feature coupling constants $\lambda_\nu$, $\nu=1,2,\cdots$,
and elementary degrees of freedom $\{{o}_\Gamma\}$ labelled by some labels \(\Gamma\)
are {\it dual} if there
exists an {\it alternative representation} $\{\hat{o}_\Gamma\}$ of the
algebra of the $\{{o}_\Gamma\}$ such that
\begin{equation}\label{d_condition}
H_1[\lambda_1,\lambda_2,\cdots](o_\Gamma)=
H_2[\lambda_1^*,\lambda_2^*,\cdots](\hat{o}_\Gamma).
\end{equation}
That is, \(H_1\) and \(H_2\) are dual if they are equal up to an {\it
operator change of variables} \(o_\Gamma \rightarrow \hat{o}_\Gamma\), together
with a readjustment of couplings \(\lambda_\nu\rightarrow\lambda_\nu^*\).
In summary, we have described notions of duality that arose
in three different fields
of physics. It is not clear a {\it a priori} that they are related
beyond general, conceptual features, but we will show in this paper
that there is a common mathematical background underlying all three of them.
\subsection{The traditional approach to quantum dualities}
\label{sec2.2}
As described near the end of the last section, quantum dualities
have the following far reaching consequence: If Equation
\eqref{d_condition} holds, it must be that the energy spectra of the
two dual Hamiltonians satisfy
\begin{equation}\label{same_levels}
E_1(\lambda_1,\lambda_2,\cdots)=E_2(\lambda_1^*,\lambda_2^*,\cdots).
\end{equation}
This suggests that a quantum duality may be a unitary equivalence of the form
\begin{equation}\label{d_as_u}
\mathcal{U}_{\sf d}\, H_1[\lambda_1,\lambda_2,\cdots](o_\Gamma)\,
\mathcal{U}^{\dagger}_{\sf d}=H_2[\lambda_1^*,\lambda_2^*,\cdots](\hat{o}_\Gamma),
\end{equation}
that would imply Equation \eqref{same_levels} right away. Surprisingly,
the definition of quantum dualities just introduced
seems to be incompatible with Equation \eqref{d_as_u}, as we
will explain below.
Let us consider a specific, non-trivial example in detail, the quantum,
one-dimensional ($d=1$) Ising model in a transverse field
(or ``quantum Ising chain'' for short), specified by the Hamiltonian
\begin{equation}
H_{\sf I}[h,J](\sigma)=\ \sum_i\ (h\sigma^x_i+\ J\sigma^z_i\sigma^z_{i+1})
\label{infinite_ising_transverse}.
\end{equation}
\(H_{\sf I}\) features \(S=1/2\) spins located at each
site $i\in\mathbb{Z}$ of a chain, represented
by Pauli matrices \(\sigma^x_i, \sigma^z_i\). They constitute
the elementary degrees of freedom that were denoted by \(o_\Gamma\)
near the end of last section. The goal is to show, along the lines of
Equation \eqref{d_condition} that \(H_{\sf I}\) is
self-dual (that is, dual to itself). So
we must find a new representation $\mu^x_i,\ \mu^z_i$ of the Pauli matrices
and new values \(J^*,\ h^*\) of the couplings so that \(H_{\sf I}[h,J](\sigma)=
H_{\sf I}[h^*,J^*](\mu)\). Since we know from the exact solution of \(H_{\sf I}\)
\cite{pfeuty} that its energy levels are symmetric in \(J,\ h\) (so that
\(E_{\sf I}(J,h)=E_{\sf I}(h,J)\)), we set \(J^*=h\) and \(h^*=J\).
We are left then with the problem of finding an appropriate dual representation
of the Pauli matrices.
The equality
\begin{equation}\label{sdisingold}
\sum_{i}\ (h\sigma^x_i+J\sigma^z_i\sigma^z_{i+1})\ =\ \sum_{i}\ (J\mu^x_i
+h\mu^z_i\mu^z_{i+1}),
\end{equation}
suggests setting up the relations,
\begin{eqnarray}
\label{ldis}
\mu^z_i\mu^z_{i+1}&=&\sigma^x_{i+m},\ \ \ \ \ \ \ \ \ \ \ \ \ \ \
m=?\nonumber\\ \mu^x_i&=&\sigma^z_{i+m'}\sigma^z_{i+1+m'}, \ \ \ \ m'=?.
\end{eqnarray}
As underscored by the question marks in Equation \eqref{ldis}, we have
to decide what $m$ and $m'$ should be. The obvious choice $m=m'=0$
leads to
\begin{equation}\label{baddualising}
\mu^z_i\mu^z_{i+1}=\sigma^x_{i},\ \
\ \ \ \ \ \
\mu^x_i=\sigma^z_{i}\sigma^z_{i+1}.
\end{equation}
On the other hand, if the new spin variables $\mu$ exist at all,
they must satisfy ${(\mu^z_i)}^2=1$. Thus,
\begin{equation}\label{mustring}
\mu^z_i=\mu^z_i\mu^z_{i+1}\times
\mu^z_{i+1}\mu^z_{i+2}\times\cdots=\prod_{m=i}^\infty\sigma^x_m.
\end{equation}
But then we see from Equations \eqref{baddualising} and \eqref{mustring} that
\(\mu^x_i\) {\it commutes} with \(\mu^z_i\).
Let us set then \(m=0\) and \(m'=-1\) in Equation \eqref{ldis}, so that
\begin{equation}
\mu^x_i=\sigma^z_{i-1}\sigma^z_i\ \ \ \ \ \ \mu^z_i\mu^z_{i+1}=\sigma^x_i.
\end{equation}
The solution to {\it this} set of equations is
\begin{equation}
\label{muxmuz}
\mu^x_i=\sigma^z_{i-1}\sigma^z_i,\ \ \ \ \ \ \mu^z_i=
\prod_{m=i}^\infty\sigma^x_m,
\end{equation}
and now \(\mu^x_i\), \(\mu^z_i\) do satisfy the correct spin-1/2 algebra.
This completes the proof in the traditional approach to quantum dualities
that $H_{\sf I}$ is self-dual \cite{fradkin_susskind}.
Admitting that this example is a fair representation of an
``average" quantum duality, we can infer that
\begin{enumerate}
\item{quantum dualities need not be strong-coupling/weak coupling
relations;}
\item{quantum dualities are {\it ``fundamentally" non-local},}
\item{quantum dualities are {\it not} unitarily implementable.}
\end{enumerate}
The last statement follows from this simple observation. Suppose
we could recast the self-duality of the Ising chain as a unitary
equivalence \(\mathcal{U}_{\sf d}H_{\sf I}[h,J]\mathcal{U}_{\sf d}^\dagger=
H_{\sf I}[J,h]\). Then we would have that
\begin{equation}
\label{paradox}
\mathcal{U}_{\sf d} H_{\sf I}[0,J]\mathcal{U}_{\sf d}^\dagger=
\mathcal{U}_{\sf d}\left(\sum_i\ J\sigma^z_i\sigma^z_{i+1}\right)
\mathcal{U}_{\sf d}^\dagger=H_{\sf I}[J,0]=\sum_i\ J\sigma^x_i,
\end{equation}
but this cannot possibly be right, because \(H_{\sf I}[0,J]\)
and \(H_{\sf I}[J,0]\) {\it have different level degeneracies}.
This is not to say, however, that the traditional approach, based
on operator changes of variables, goes successfully
beyond unitary equivalence. Equation \eqref{sdisingold}
implies that
\begin{equation}
\sum_i\ J\sigma^z_i\sigma^z_{i+1}=\sum_i\ J\mu^x_i.
\end{equation}
If the $\mu^x_i,\mu^z_i$ are truly an alternative representation of the
Pauli matrices, this equation cannot possibly be right either, for the
same reasons as before.
\section{Bond-algebraic approach to quantum dualities}
\label{sec3}
This section is devoted to explaining our theory of quantum and classical
dualities based on bond algebras, and it
discusses every major new idea and mathematical technique that
we introduce to the subject (some of them advanced in Reference \cite{con}).
We will argue that
\begin{enumerate}
\item{quantum and (a very large class of) classical dualities are unitary
equivalences (or projective unitary
equivalences if the duality eliminates gauge symmetries), and that}
\item{the easiest way to search for dualities
is to look for structure-preserving mappings between Hamiltonian-dependent
{\it bond algebras}, Section \ref{sec3.8}.}
\end{enumerate}
Conceptually, (1) is more important than (2), yet mathematically it
follows from (2), as basically does everything else in this paper.
\subsection{Bond algebras and the concept of locality}
\label{sec3.2}
It is a basic fact of physics that the Hamiltonian of a system determines
its dynamics and thermodynamics (some important consequences of this
statement are reviewed in Appendix \ref{sec3.1}).
Bond algebras \cite{bondDec08} were devised to exploit
a simplifying feature common to most Hamiltonians, and rooted in
fundamental physical principles: {\it Hamiltonians are sums
of (possibly a huge number of) simple, local terms}
({\it local} is used in a broad sense in this paper, either to indicate
that interactions are local in space/space-time in the sense of field
theory, or that they involve only a few degrees of freedom).
Take for example the Hamiltonian for $N$ electrons of
charge $e$ and mass $m$ in an external potential,
\begin{equation}
\label{general_cond-mat}
H_e=\sum_{i=1}^{N}\left(\frac{1}{2m}{\bm p}_i^2+V({\bm{x}}_i)\right)
+\sum_{i\neq j}\frac{e^2}{\vert {\bm{x}}_i-{\bm{x}}_j\vert}.
\end{equation}
The elementary degrees of freedom \(x_i^\mu, p_i^\nu\), satisfy
the Heisenberg relations
\begin{equation}\label{heisenberg}
[x_i^\mu,p_i^\nu]=i\hbar\delta_{\mu,\nu},\ \ \ \ \ \ \mu,\nu=1,2,3,
\end{equation}
or commute otherwise (from now on, we set \(\hbar=1\)).
This is in no way specific to the problem at hand (understanding
\(H_e\)), but is just a general fact. On the other
hand, \(H_e\) is the sum of \(2N+ N(N-1)/2\) individual operators
\begin{equation}
{\bm p}_i^2,\ \ \ \ \ \ V({\bm{x}}_i),\ \ \ \ \ \ \frac{1}{\vert {\bm{x}}_i-{\bm{x}}_j\vert},
\end{equation}
that we would like to consider altogether on an equal footing, and so we
call them generically {\it bonds}. Taken individually bonds
are still elementary to understand, but, in contrast to elementary degrees of
freedom, they satisfy algebraic relations that are {\it specific to the
problem at hand}. Also, there is
{\it an algebraic notion of connectivity} for bonds that
reflects locality, in the sense that if two bonds commute, they can
only influence each other {\it indirectly}.
Thus the algebraic relations between bonds strike a balance
between general physical principles and model specificity
suggesting that algebras of bonds (as opposed to algebras of elementary
degrees of freedom) could potentially become important mathematical
tools.
This idea was pioneered in Reference
\cite{bondDec08}, where algebras of bonds where exploited to solve exactly
several quantum lattice models of interest in the context of topological quantum order.
Let us introduce next the formal definition of a {\it bond algebra}.
Consider a Hamiltonian operator $H$, written as
a sum of bond operators $h_\Gamma$
\begin{equation}
H=\sum_\Gamma\ \lambda_\Gamma\ h_\Gamma.
\label{bae}
\end{equation}
with c-number coupling constants \(\lambda_\Gamma\).
The index ``\(\Gamma\)'' is completely general. It could stand for
a particle index, or for
a site, a link, or some other subregions of a lattice $\Lambda$, or may
denote a point \({\bm{x}}\) in space or a Fourier mode,
or may stand for any other suitable label one can think of.
\begin{definition}\label{defba}
A bond algebra for the Hamiltonian \(H=\sum_\Gamma\ \lambda_\Gamma\ h_\Gamma\)
with bond decomposition \(\{h_\Gamma\}_\Gamma\),
is the von Neumann algebra ${\cal A}\{h_\Gamma\}$ generated
by the bonds.
\end{definition}
The basic mathematical aspects of this definition (including the
definition of a von Neumann algebra) are discussed in the next section.
The rest of this section is devoted to an informal discussion of bond algebras.
Intuitively speaking, \({\cal A}\{h_\Gamma\}\) is an algebra of
operators generated by taking all possible {\it finite}, complex, linear
combinations of
powers and products of bonds, their Hermitian conjugates, and the
identity operator \(\mathbb{1}\),
\begin{eqnarray}\label{oo}
\{\mathbb{1}, h_\Gamma, h_\Gamma^\dagger,
h_\Gamma h_{\Gamma'}, h_\Gamma^\dagger h_{\Gamma'},
h_{\Gamma'}^\dagger h_\Gamma, h_{\Gamma'}^\dagger h_\Gamma^\dagger,
h_\Gamma h_{\Gamma'} h_{\Gamma''}, \cdots \ \}.
\end{eqnarray}
So by construction, if an operator $\mathcal{O}
\in {\cal A}\{h_\Gamma\}$ then $\mathcal{O}^\dagger
\in {\cal A}\{h_\Gamma\}$ as well. This intuitive picture of bond
algebras suffices to understand most of the rest of the paper.
It is important to understand that a bond algebra
is {\it not} determined by a Hamiltonian \(H\), but rather
by its bond decomposition \(\{h_\Gamma\}_\Gamma\). Any single
Hamiltonian may produce many
different bond algebras, since different decompositions
\begin{equation}
H=\sum_\Gamma\ \lambda_\Gamma\ h_\Gamma=\sum_\Sigma\ {\lambda'}_\Sigma
\ {h'}_\Sigma
\end{equation}
define different, equally valid sets of bonds that can potentially
generate very different bond algebras. Conversely,
any one bond algebra may be related to many different Hamiltonians.
Consider for illustration the single site spin Hamiltonian
\begin{equation}
H_{\sf I}=h_x\sigma^x+h_y\sigma^y.
\end{equation}
One can take \(\sigma^{x}\) and \(\sigma^{y}\)
as generating bonds, or the single bond
$(h_x\sigma^x+h_y\sigma^y)$. Then we get
two bond algebras that are clearly different,
\begin{equation}\label{hxhyex}
{\cal A}\{\sigma^x,\sigma^y\}\ \neq\ {\cal A}\{h_x\sigma^x+h_y\sigma^y\},
\end{equation}
since \({\cal A}\{h_x\sigma^x+h_y\sigma^y\}\) is commutative while
\({\cal A}\{\sigma^x,\sigma^y\}\) is not.
This flexibility of the concept of bond algebra turns out to be an
essential advantage.
{\it Applications dictate what bond decomposition is best for
any given problem}.
The complexity of a bond algebra can vary, and a practical measure
of that complexity is simply afforded by considering
the Hilbert space \(\mathcal{H}\) on which the bond algebra acts on.
Then one can recognize three increasingly difficult (in the number of
resources) scenarios:
\begin{enumerate}
\item{\(\mathcal{H}=\mathcal{H}_1\otimes\cdots\otimes\mathcal{H}_N\), where each factor
in the tensor product is finite dimensional;}
\item{\(\mathcal{H}=\mathcal{H}_1\otimes\cdots\otimes\mathcal{H}_N\), where some or all
factors are infinite dimensional;}
\item{ \(\mathcal{H}=\bigotimes_{\alpha\in I} \mathcal{H}_\alpha\),
where the index set \(I\) is infinite, so that \(\mathcal{H}\) is an {\it
infinite tensor product} \cite{vonneumannII}. For example,
\(\mathcal{H}=\bigotimes_{i\in\mathbb{Z}} \mathbb{C}^2_i\)}.
\end{enumerate}
The first scenario (1) is elementary (bond algebras are then just matrix
algebras), and (2) is moderately simple,
but (3) is directly connected to the thermodynamic limit
and/or the continuum limit of QFTs \cite{sewell,barton},
and it is the source of endless fascination and complications.
In practice, problems in (3) must be regularized (turned into problems
in (1) or (2)) before any progress can be made (see
Appendix \ref{appE} on LQFT). But, as long
as bonds are chosen to be {\it local} (specifically
in the sense that they act non trivially only
on a {\it finite} number of factors), bond algebras are perfectly well defined
\cite{vonneumannII} even if the state space were as complicated an object
as (3) above.
Let us notice next that the generators listed in
Equation \eqref{oo} need not be in general linearly independent.
Then one can find a (potentially much) smaller basis
\(\{\mathcal{O}_{\alpha}\}\) for the bond algebra \({\cal A} \{h_\Gamma\}\),
and decompose products of bonds as
\begin{equation}
h_{\Gamma_1}\cdots h_{\Gamma_N}\cdots=
\sum_\alpha\ c_\alpha{\cal{O}}_{\alpha}.
\end{equation}
It is interesting to recognize that bond algebras must have a basis, because
this shows that a kind of ``reducibility hypothesis" \cite{kadanoff,Kadanoff_Ceva}, or
``operator product expansion" formula
\begin{eqnarray}\label{Aijk}
{\cal{O}}_{\alpha} {\cal{O}}_{\beta} = \sum_{\gamma}
A_{\alpha \beta}^\gamma\ {\cal{O}}_{\gamma}
\end{eqnarray}
holds. The c-numbers \(A_{\alpha \beta}^\gamma\) are {\it structure constants}
for the bond algebra.
We can get a feeling for the physical meaning
of the structure constants by taking the expectation value of Equation \eqref{Aijk}
(vacuum expectation value, or thermal
average, etc.),
\begin{equation}\label{algebra_averages}
\langle{\cal{O}}_{\alpha} {\cal{O}}_{\beta}\rangle = \sum_{\gamma}
A_{\alpha \beta}^\gamma\ \langle{\cal{O}}_{\gamma}\rangle.
\end{equation}
This shows that bond algebras afford a partial realization of the idea
of algebras of fluctuating variables (see \cite{kadanoff, Kadanoff_Ceva},
and references therein).
{}From a different perspective, the structure constants
\(A_{\alpha \beta}^\gamma\) can be seen
as generalized constants of motion.
In the Heisenberg picture, the basis of the bond algebra evolves as
\begin{equation}
\mathcal{O}_{\alpha}(t)=\mathcal{U}(t)^\dagger\, \mathcal{O}_{\alpha}\,
\mathcal{U}(t),
\end{equation}
where \(\ \mathcal{U}(t)=\widehat{T}\ e^{-i\int_0^t H dt'}\), and $\widehat{T}$
is the time-ordering symbol. Then,
\begin{equation}
{\cal{O}}_{\alpha}(t) {\cal{O}}_{\beta}(t)\ =\ \sum_{\gamma}
A_{\alpha \beta}^\gamma\ {\cal{O}}_{\gamma}(t)
\end{equation}
{\it with the same structure constants} as in Equation \eqref{Aijk} at
time $t=0$.
\subsection{Some mathematical aspects of bond algebras}
\label{sec3.3}
By definition, bond algebras are von Neumann algebras of operators.
In this section we spell out the meaning and far reaching
consequences of this requirement. We start by recalling the definition of a
von Neumann algebra \cite{vonneumannI,jones}.
Let \(\mathcal{H}\) be a Hilbert space (the space of quantum states),
and let \(B(\mathcal{H})\) denote
the algebra of bounded operators on \(\mathcal{H}\) (an operator \(\mathcal{O}\)
is bounded if there is some number \(0\leq C<\infty\) such that
\(||\mathcal{O} v||\leq C||v||\), for every vector \(v\in \mathcal{H}\)).
If \(\mathcal{S}\subset B(\mathcal{H})\) is an arbitrary subset, its commutant
\(\mathcal{S}'\in B(\mathcal{H})\) is the subalgebra defined by
\begin{equation}
\mathcal{S}'=\{\mathcal{O}\in B(\mathcal{H})\ |\
\forall\mathcal{R}\in\mathcal{S},\ \ \
\mathcal{O}\mathcal{R}=\mathcal{R}\mathcal{O}\}.
\end{equation}
\begin{definition}
A subalgebra \({\cal A}\subset B(\mathcal{H})\) is a {\it von Neumann algebra}
if it satisfies three algebraic conditions \cite{jones}:
\begin{itemize}
\item{It contains the identity operator, \(\mathbb{1}\in{\cal A}\).}
\item{ It is closed under Hermitian conjugation, if \(\mathcal{O}\in{\cal A}\),
then \(\mathcal{O}^\dagger \in {\cal A}\) as well.}
\item{It is equal to its bycommutant, \({\cal A}={\cal A}''\).}
\end{itemize}
\end{definition}
Since von Neumann algebras are algebras of {\it bounded} operators,
the sense in which a bond decomposition \(\{h_\Gamma\}_\Gamma\)
generates a (von Neumann) bond algebra \(\mathcal{A}\{h_\Gamma\}\)
varies according to whether the bonds are bounded operators or not.
If the bonds are all bounded operators, the bond algebra they generate
is simply the smallest von Neumann algebra \({\cal A}\{h_\Gamma\}\subset
B(\mathcal{H})\) that contains every bond.
If the bonds are {\it not} all bounded operators, this notion needs to be refined.
An operator \(\mathcal{O}\) (not necessarily bounded) is {\it affiliated}
to a von Neumann algebra \(\mathcal{A}\) if it commutes \(\mathcal{O}U=
U\mathcal{O}\) with every unitary operator \(U\in\mathcal{A}'\).
Every operator that is affiliated and bounded belongs to $\mathcal{A}$.
This notion is useful for the following reason.
Suppose \(\mathcal{O}\) is unbounded and affiliated to \(\mathcal{A}\), and
suppose also that \(\mathcal{O}\) admits an spectral decomposition, so that
we can construct operators \(f(\mathcal{O})\) that are functions of \(\mathcal{O}\)
in the usual way. Then one can show \cite{vonneumannI} that every {\it bounded}
\(f(\mathcal{O})\) {\it is} an operator in \(\mathcal{A}\),
\(f(\mathcal{O})\in\mathcal{A}\),
even though \(\mathcal{O}\) itself is not. So we define:
If the set of bonds generators \(\{h_\Gamma\}_\Gamma\)
includes unbounded operators, then the bond algebra they generate
is the {\it smallest} von Neumann algebra \({\cal A}\{h_\Gamma\}\subset
B(\mathcal{H})\) such that every bond \(h_\Gamma\) is affiliated to
\({\cal A}\{h_\Gamma\}\) (such an algebra always exists \cite{sunder}).
In summary, whatever the nature of the bonds may
be, their bond algebra is a convenient (since it contains only bounded operators)
yet faithful representative of the structure of the interactions that are
embodied in the bonds.
A mapping of von Neumann algebras \(\Phi:\mathcal{A}_1\rightarrow\mathcal{A}_2\)
is an {\it homomorphism} if
\begin{eqnarray}
\Phi(\mathbb{1})&=& \mathbb{1},\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
\ \ \ \ \ \ \ \ \ \ \ \ \, \Phi(\mathcal{O}^\dagger)=
\Phi(\mathcal{O})^\dagger,\label{homomorphism}\\
\Phi({\cal O}_1{\cal O}_2)&=&\Phi({\cal O}_1)\Phi({\cal O}_2),
\ \ \ \ \ \ \Phi({\cal O}_1+\lambda {\cal O}_2)=
\Phi({\cal O}_1)+\lambda\Phi({\cal O}_2).\nonumber
\end{eqnarray}
If \(\Phi\) is also one-to-one and onto, then it is called
an {\it isomorphism}. As mentioned before, one of the main goals of this paper is
to establish a connection between dualities and unitary transformations. The following
theorem \cite{jones} then explains to a great extent our insistence in embedding bonds
in a von Neumann algebra.
\begin{theorem}\label{isoarespatial}
Let \(\mathcal{A}_i\) be von Neumann algebras of operators on the Hilbert spaces
\(\mathcal{H}_i\), for \(i=1,2\). If \(\Phi:\mathcal{A}_1\rightarrow\mathcal{A}_2\)
is an isomorphism, then there exists
\begin{itemize}
\item{a Hilbert space \(\mathcal{M}\), and}
\item{a unitary transformation \(\mathcal{U}:\mathcal{H}_1\otimes\mathcal{M}
\rightarrow\mathcal{H}_2\otimes\mathcal{M}\) such that}
\end{itemize}
\begin{equation}\label{spatial}
\Phi({\cal O})\otimes\mathbb{1}=\mathcal{U}({\cal O}\otimes\mathbb{1})\mathcal{U}^\dagger ,
\end{equation}
\end{theorem}
\noindent
where \(\mathbb{1}\)
stands for the identity operator on \(\mathcal{M}\). In this paper we will often
be able to take \(\mathcal{M}=
\mathbb{C}\), so that Equation \eqref{spatial} simplifies to
\(\Phi({\cal O})=\mathcal{U}{\cal O}\mathcal{U}^\dagger\). Then we say that \(\Phi\)
is unitarily implementable.
\subsection{Dualities as isomorphisms of bond algebras}
\label{sec3.4}
In this section we introduce and illustrate our definition of
quantum duality based on bond algebras. It will be refined in Section
\ref{sec3.12} to include models with gauge symmetries. Classical
dualities will be defined similarly in
Section \ref{classical&quantum}, after we discuss
how to associate bond algebras to classical models
of statistical mechanics.
Our new approach to dualities is based on the recognition that, if we exclude
models with gauge symmetries for the moment,
{\it quantum dualities are isomorphisms of bonds algebras}. More precisely \cite{con},
\begin{definition}\label{algebraically_dual}
Two Hamiltonians $H_{1}$ and $H_2$ are dual if there is a bond algebra
\(\mathcal{A}_{H_1}\) for \(H_1\) isomorphic to some bond algebra
\(\mathcal{A}_{H_2}\) for \(H_2\), and if the isomorphism
$\Phi_{{\sf d}}:\mathcal{A}_{H_1}\rightarrow\mathcal{A}_{H_2}$ maps
\(H_1\) to \(H_2\).
\end{definition}
Since \(H_1\) and \(H_2\) are self-adjoint, Equation \eqref{spatial}
implies that these Hamiltonians share identical spectra and level degeneracies,
and so
\begin{equation}\label{isoduu}
H_2=\mathcal{U}_{{\sf d}} H_1\mathcal{U}^\dagger_{{\sf d}}.
\end{equation}
A Hamiltonian \(H[\lambda]\) which depends on some set
of coupling parameters $\lambda=(\lambda_1,\lambda_2,\cdots)$ is
self-dual if it is dual to itself, up to a change in the coupling
$\lambda\rightarrow\lambda^*$, with \(\lambda^*\) the {\it dual couplings}.
Notice that by Equation \eqref{homomorphism}, a bond algebra
homomorphism $\Phi_{{\sf d}}$ preserves the equations of motion of an arbitrary
observable ${\cal O}$
\begin{equation}\label{heisenberg_covariant}
\frac{d {\cal O}}{dt}-i[H_1,{\cal O}]=0\ \ \ \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \ \ \
\frac{d\Phi_{{\sf d}}({\cal O})}{dt}-i[H_2,\Phi_{{\sf d}}({\cal O})]=0.
\end{equation}
Now that we have a precise definition of duality, we need to:
\begin{itemize}
\item{show that it includes the known dualities, and}
\item{show that it is useful.}
\end{itemize}
To start with, let us show that the
quantum Ising chain is self-dual in the sense of definition \ref{algebraically_dual}.
Take the basic bonds in \(H_I\) of Equation \eqref{infinite_ising_transverse}
to be $ \{\sigma^z_i\sigma^z_{i+1}\},\{ \sigma^x_i \}$.
They generate a bond algebra $\mathcal{A}_{\sf I}$
that we can characterize in terms of relations:
\begin{enumerate} \label{relations}
\item{\((\sigma^z_i\sigma^z_{i+1})^2=\mathbb{1}=(\sigma^{x}_i)^2\);}
\item{Any bond $\sigma^x_i$ {\it anti-commutes} with two other
bonds, $\sigma^z_{i-1}\sigma^z_i$ and $\sigma^z_{i}\sigma^z_{i+1}$, and
{\it commutes} with all other bonds;}
\item{Any bond $\sigma^z_{i}\sigma^z_{i+1}$ {\it anti-commutes} with two
other bonds, $\sigma^x_i$ and $\sigma^x_{i+1}$, and {\it commutes} with
all other bonds.}
\end{enumerate}
We will assume that these relations characterize the bond algebra.
While this may seem plausible it is far from obvious, since it is not
hard to argue that $\mathcal{A}_{\sf I}$ is reducible. There
are, however, consistency checks that we can run on the results
that we will obtain from this assumption. Also let us point out that the bond algebra
$\mathcal{A}_{\sf I}$ is a well defined algebra of bounded operators, in spite
of the fact that it is generated by an infinite number of bonds. This follows
because the bonds act locally on the infinite tensor product
\(\bigotimes_{i\in \mathbb{Z}}\mathbb{C}^2_i\) \cite{vonneumannII}.
Coming back to the set of relations above, we see that
\(\sigma^x_i\) and \(\sigma^z_{i}\sigma^z_{i+1}\) play
perfectly symmetrical roles, and so we can set up the relation-preserving
mapping
\begin{equation}\label{aut_ising1}
\Phi_{\sf d}(\sigma^z_{i}\sigma^z_{i+1} )=\sigma^x_i,\ \ \ \ \ \ \ \
\Phi_{\sf d}(\sigma^x_{i})=\sigma^z_{i-1}\sigma^z_i.
\end{equation}
These equations define the action of \(\Phi_{\sf d}\) on bonds alone
but, since it preserves all the algebraic relations among them, it extends
to a unique isomorphism of the {\it full} bond algebra $\mathcal{A}_{\sf I}$.
\(\Phi_{\sf d}\) is illustrated in Figure \ref{sdinfiniteising}.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.75\columnwidth]{./sdinfiniteising.eps}
\end{center}
\caption{A graphic representation of two quantum Ising chains,
connected by the self-duality isomorphism \(\Phi_{{\sf d}}\) of
Equation \eqref{aut_ising1}. The crosses
$\times$ represent the bonds $\sigma^x_i$, and the thick lines
between crosses represent the bonds $\sigma^z_i\sigma^z_{i+1}$.
\(\Phi_{{\sf d}}\) exchanges the two while preserving all algebraic relations.}
\label{sdinfiniteising}
\end{figure}
Furthermore, \(\Phi_{{\sf d}}\) maps \(H_{\sf I}[h,J]\) to \(H_{\sf I}[J,h]\). It follows
that the quantum Ising chain is {\it self-dual} in the sense of
definition \ref{algebraically_dual}.
It is not hard to see that \(\Phi_{{\sf d}}\) is unitarily implementable (recall the definition
after Theorem \ref{isoarespatial}), so that there is a \(\mathcal{U}_{{\sf d}}\) such that
\begin{equation}
\mathcal{U}_{{\sf d}}\ \sigma^z_{i}\sigma^z_{i+1}\ \mathcal{U}_{{\sf d}}^\dagger =\sigma^x_i,
\ \ \ \ \ \
\mathcal{U}_{{\sf d}}\ \sigma^x_{i}\ \mathcal{U}_{{\sf d}}=\sigma^z_{i-1}\sigma^z_i, \ \ \ \ \ \
\forall i\in \mathbb{Z} .
\end{equation}
The homomorphism \(\Phi_{{{\sf d}}}\) reveals that the Ising chain is
self-dual due to a {\it local mapping} that reflects a symmetry of its {\it local}
interactions. In contrast, the traditional approach
seems to imply that dualities are of necessity non-local, because
it focuses on non-local transformations of
elementary degrees of freedom. Notice also that the self-duality mapping
\(\Phi_{{{\sf d}}}\) determines a large family of perturbations of \(H_{\sf I}\)
that preserve its self-dual character. For example,
\begin{equation}
H=H_{\sf I}+\lambda \sum_i\ ( \sigma^y_i\sigma^z_{i+1}
+\sigma^z_i\sigma^y_{i+1})
\end{equation}
is self-dual because the perturbation (the term proportional to
\(\lambda\)) is invariant under \(\Phi_{{{\sf d}}}\) (the action of
\(\Phi_{{{\sf d}}}\) on \(\sigma^y_i\sigma^z_{i+1}\) for instance can
be determined by factoring \(\sigma^y_i\sigma^z_{i+1}=
-i \sigma^z_i\sigma^z_{i+1} \sigma^x_i\)). Also
we can apply \(\Phi_{{{\sf d}}}\) to Hamiltonians other than \(H_{\sf I}\),
as long as they are affiliated to $\mathcal{A}_{\sf I}$.
Consider the \(d=1\) dimensional, spin \(S=1/2\) XY-model,
\begin{equation}
H_{\sf XY}=\sum_i\ (J_x\sigma^x_i\sigma^x_{i+1}+
J_z\sigma^z_i\sigma^z_{i+1}).
\end{equation}
The bonds \(\sigma^z_i\sigma^z_{i+1}\) of
$H_{\sf XY}$ are already bonds of \(H_{\sf I}\). The
\(\sigma^x_i\sigma^x_{i+1}\) in $H_{\sf XY}$ are the products of two
bonds of \(H_{\sf I}\). Thus it is possible to use the isomorphism of
the quantum Ising model to compute a dual form of the XY-model. As
\(\sigma^x_i\sigma^x_{i+1}\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}\
\sigma^z_{i-1}\sigma^z_{i} \sigma^z_{i} \sigma^z_{i+1}\), we find that
\begin{equation}
H_{\sf XY}\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}\
H_{\sf Inn}=\sum_i\ (J_x\sigma^z_{i-1}\sigma^z_{i+1}+J_z\sigma^x_i) .
\end{equation}
The fact that, in \(d=1\), \(H_{\sf XY}\) and \(H_{\sf Inn}\) share the
same energy spectra was first noticed in Reference \cite{pfeuty},
but explained only later in Reference \cite{grady}
(\(H_{\sf Inn}\) is trivially dual to two decoupled Ising chains).
Next we would like to establish the
precise connection between the bond-algebraic and traditional
approach to quantum dualities of Section \ref{sec2.2}.
\subsection{Connection to the traditional approach: Determination of
dual variables}
\label{sec3.5}
The traditional approach to dualities (Section \ref{sec2.2})
focuses on dual variables, that is, on
operator change of variables that are non-local in general. In
contrast, the bond-algebraic approach to dualities
of the previous section focuses on local mappings of bonds.
How can the two be related? As it turns out, the isomorphism
of bond algebras determines uniquely the dual variables of the problem.
This is the bridge between the bond-algebraic and the traditional approach.
Let us illustrate this point with the quantum
Ising chain. To start with, consider the relation (see Equation
\eqref{aut_ising1})
\begin{equation}
H_{\sf I}[h,J]=\sum_i
(h\Phi_{\sf d}(\sigma^z_i\sigma^z_{i+1})+J\Phi_{\sf d}(\sigma^x_i)).
\end{equation}
If the individuals spins \(\sigma^z_i\) happen to belong to the
bond algebra \(\mathcal{A}_{\sf I}\), then we can further write
\begin{equation}\label{first_step_dual_variables}
H_{\sf I}[h,J]=\sum_i (h\Phi_{\sf d}(\sigma^z_i)\Phi_{\sf
d}(\sigma^z_{i+1})+J \Phi_{\sf d}(\sigma^x_i)).
\end{equation}
If we now compare this last relation to Equation \eqref{sdisingold},
we see that the dual variables could be connected to the self-duality isomorphism
as
\begin{eqnarray}
\Phi_{\sf d}(\sigma^z_i)=\mu^z_i,\ \ \ \ \ \
\Phi_{\sf d}(\sigma^x_i)=\mu^x_i.
\end{eqnarray}
But \(\Phi_{\sf d}\) is defined by \eqref{aut_ising1}. This makes
sense of $\mu^x_i$ as \(
\mu^x_i\equiv \Phi_{\sf d}(\sigma^x_i)= \sigma^z_{i-1}\sigma^z_i
\), but it is not clear what the action of \(\Phi_{\sf d}\) on \(\sigma^z_i\)
should be.
Now, at least formally,
\begin{eqnarray}\label{infprodsz}
\sigma^z_i=\prod_{m=i}^\infty\sigma^z_m\sigma^z_{m+1}.
\end{eqnarray}
Unfortunately, this does not quite show that \(\sigma^z_i\in\mathcal{A}_{\sf I}\),
because the left-hand side of Equation \eqref{infprodsz} features an {\it infinite}
product of bonds. An infinite combination (sum and/or product)
of bonds will only be an element in the
bond algebra if it converges to some bounded operator in the strong or weak
operator topology \cite{vonneumannI}. {\it Suppose though for now} that
\(\sigma^z_i\in\mathcal{A}_{\sf I}\). Then we can compute
\begin{eqnarray}
\Phi_{\sf d}(\sigma^z_i)=\Phi_{\sf
d}\left(\prod_{m=i}^\infty\sigma^z_m\sigma^z_{m+1}
\right)=\prod_{m=i}^\infty\Phi_{\sf d}(\sigma^z_m\sigma^z_{m+1})=
\prod_{m=i}^\infty\sigma^x_m=\mu^z_i.
\end{eqnarray}
{\it Thus the expressions we obtain for the dual variables
\(\mu^x_i, \mu^z_i\) are identical to
the ones derived by traditional arguments in {\rm Section \ref{sec2.2}}}
[see Equation \eqref{mustring} in particular]. Because $\Phi_{\sf d}$
is an algebra isomorphism, the dual variables are guaranteed to
satisfy the same algebra as the original variables \(\sigma^x_i,\sigma^z_i\).
But we can view this from a different perspective. The fact that the
dual variables satisfy the correct algebra affords an independent check
supporting that \(\Phi_{{\sf d}}\) is indeed an isomorphism, and thus the relations
that were assumed to characterize the bond algebra are complete.
In summary, the structure of the bond
algebra determines the self-duality homomorphism, and the
self-duality homomorphism enables us to compute the dual variables.
Thus, we have both a test for self-duality and an algorithm to construct
dual variables.
Now that we have the intuitive picture, let us point out for the sake
of mathematical rigor that \(\sigma^z_i\notin\mathcal{A}_{\sf I}\) and
\(\Phi_{{\sf d}}(\sigma^z_i)\) is not defined. The reason is that formally we can
also write \(\sigma^z_i=\prod_{m=-\infty}^{i-1}\sigma^z_m\sigma^z_{m+1}\).
Then it would follow from computing the action of \(\Phi_{\sf d}\)
of both representations of \(\sigma^z_i\) that
\begin{equation}
\prod_{m=i}^\infty\sigma^x_m\ \stackrel{?}{=}\
\prod^{i-1}_{m=-\infty}\sigma^x_m.
\end{equation}
But this cannot possibly hold true. It is important to notice that
this is {\it not} a limitation of the bond-algebraic approach to dualities
(that managed to establish in the previous section the self-duality of the Ising
model purely by well-defined manipulations involving bonds), but rather of the concept of
dual variables in infinite systems. In practice, however, infinite systems are
studied as limits of finite ones, for which dual variables are
well defined and can be computed as above.
We will come back to this issue in Section \ref{sec3.7}.
\subsection{Abelian versus non-Abelian dualities: the Heisenberg model}
\label{sec3.6}
Lattice Non-Abelian dualities constitute one of the greatest challenges in the
theory of dualities, the classical aspects of which are discussed in
Appendix \ref{appA}. In this section, we present preliminary
contributions of bond algebras to understanding this difficult problem.
There is a broad, well established consensus among physicists that
a duality is {\it non-Abelian} if the dual models have non-Abelian
symmetries, and is Abelian otherwise \cite{savit}. Bond
algebras can realize duality mappings of an Abelian origin in models with non-Abelian
symmetries,
suggesting that this classification is not appropriate.
The discussion of this
section, based on a new duality for the Heisenberg model
in any space dimension $d$, may help sharpen the notion of non-Abelian
duality beyond the somewhat inaccurate standard lore.
The Heisenberg model,
\begin{equation}\label{anyd_heisenberg_model}
H_{\sf H}=J\sum_{\bm{r}}\sum_{\nu=1}^d\ (\sigma^x_{\bm{r}}\sigma^x_{{\bm{r}}+\bm{e_\nu}}+
\sigma^y_{\bm{r}}\sigma^y_{{\bm{r}}+\bm{e_\nu}}+\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}),
\end{equation}
is one the fundamental models of magnetism (its application
to cuprates is reviewed in Reference \cite{manousakis}).
To our knowledge, exact dualities for the Heisenberg model have not been reported
before, and this may seem reasonable, since it has a non-Abelian
group of global symmetries (it is invariant under global $SU(2)$ rotations
in spin space). Thus it is surprising to find out that,
with the help of bond algebras, we can write a duality for \(H_{\sf H}\)
right away.
The starting point is the observation that the bond algebra
\begin{equation}
\mathcal{A}_{\sf H}\equiv \mathcal{A}\{\sigma^x_{\bm{r}}\sigma^x_{{\bm{r}}+\bm{e_\nu}},
\sigma^y_{\bm{r}}\sigma^y_{{\bm{r}}+\bm{e_\nu}},\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\}
\end{equation}
is a sub-algebra of the bond algebra of the quantum Ising model in \(d\) dimensions,
\begin{equation}\label{anyd_ising}
H_{\sf I}=\sum_{\bm{r}}\ \left(h\sigma^x_{\bm{r}} +\sum_{\nu=1}^d\
J\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\right),
\end{equation}
simply because the bonds of the Heisenberg model can be written
as products of bonds of the Ising model. Then {\it any (self-)duality
for the Ising model can be translated into a duality for the Heisenberg
model}. For example, we can use the self-duality mapping of the quantum
Ising chain, Equation \eqref{aut_ising1}
to find a dual form for the $d=1$ Heisenberg model. Since
\begin{equation}
\sigma^y_i\sigma^y_{i+1}=\sigma^x_i\sigma^z_i\sigma^z_{i+1}
\sigma^x_{i+1}\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}\
-\sigma^z_{i-1}\sigma^x_i\sigma^z_{i+1},
\end{equation}
we find that
\begin{equation}\label{dual_heisenberg}
H_{\sf H}\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}\
H_{\sf H}^D=\frac{J}{4}\sum_i\ (\sigma^z_{i-1}\sigma^z_{i+1}
-\sigma^z_{i-1}\sigma^x_i\sigma^z_{i+1}+
\sigma^x_i).
\end{equation}
Appendix \ref{appG} describes
a version of this duality for finite systems that can be checked numerically.
The Hamiltonian \(H_{\sf H}^D\) has an interesting connection to the eight-vertex
model that seems to have gone unnoticed to the best of our knowledge.
Section \ref{eightvertexM} discusses the relation between the eight-vertex
model and the anisotropic quantum Heisenberg model. In
particular, its is shown that \(H_{\sf H}^D\) is directly related to the
quantum Ashkin-Teller model.
In contrast to what standard practice would suggest,
we do not think that it is appropriate
to call Equation \eqref{dual_heisenberg} a non-Abelian duality.
The duality of
Equation \eqref{dual_heisenberg}
is strictly based on the self-duality
of the quantum Ising chain, that is Abelian on at least two accounts.
First, the Ising model has only Abelian symmetries. Second,
we will show (Section \ref{sec8})
that the self-duality of the quantum Ising chain is strictly equivalent to the
self-duality of the classical $D=2$ Ising model.
Classical dualities are strictly based on very special properties of Abelian groups (see
Appendix \ref{appA}). It seems fair to say that the duality of
Equation \eqref{dual_heisenberg}
avoids the non-Abelian structure of the model, and thus
it is inappropriate to call it a non-Abelian duality.
In other words, this duality for the Heisenberg model
suggests that {\it the group of symmetries of a model} is not the most
important factor in determining the character of a duality.
The results and ideas just discussed are not peculiar to one dimension,
but before discussing the higher-
dimensional analogues of Equation \eqref{dual_heisenberg},
we need to introduce a bit of notation to describe
degrees of freedom on the {\it links} of a lattice.
In general, to specify a
link \(\bm{l}\) of a {\it hyper-cubic lattice} of dimension \(d\), we
determine first the lattice site \({\bm{r}}\) and direction \(\nu\) such that
\(\bm{l}\) connects the two sites \({\bm{r}}\) and \({\bm{r}}+\bm{e_\nu}\). Then, we
denote \(\bm{l}\) by the pair \(({\bm{r}},\nu)\), as shown in Figure
\ref{notation_links}.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.9\columnwidth]{./links.eps}
\end{center}
\caption{(Left panel) Convention to denote vertices ${\bm{r}}=(r^1,r^2)=r^1
\bm{e_1}+ r^2\bm{e_2}$ in a two-dimensional square lattice with unit vectors
$\bm{e_1}, \bm{e_2}$, and (right panel) links, attached to a vertex
${\bm{r}}$, $({\bm{r}},\nu)$ with $\nu=1,2$.}
\label{notation_links}
\end{figure}
We can now place a spin \(S=1/2\) degree of freedom at each link
\(({\bm{r}},\nu)\), represented by sets of Pauli matrices
\(\sigma^\mu_{({\bm{r}},\nu)}\), \(\ \mu=x,y,z\). Let us introduce one more
piece of notation. Both the plaquette operator
\begin{equation}\label{general_plaquette}
B_{({\bm{r}},\mu\nu)}=\sigma^z_{({\bm{r}},\mu)}\sigma^z_{({\bm{r}}+\bm{e_\mu},\nu)}
\sigma^z_{({\bm{r}}+\bm{e_\nu},\mu)}\sigma^z_{({\bm{r}},\nu)},\ \ \ \ \ \
\mu\neq\nu=1,\cdots,d,
\end{equation}
that resides on the plaquette with vertex ${\bm{r}}$ and spanned by
$(\bm{e_\mu},\bm{e_\nu})$, and the vertex operator
\begin{equation}
A_{\bm{r}}=\prod_{\nu=1}^{d}\
\sigma^x_{({\bm{r}},\nu)}\sigma^x_{({\bm{r}}-\bm{e_\nu},\nu)}\ ,
\end{equation}
that resides on the lattice site ${\bm{r}}$, will show up repeatedly in this article.
Also, in dimensions \(d=2\) or \(3\),
we prefer a more compact notation for the plaquette operator,
\begin{equation}\label{gauge_plaquette}
B_{({\bm{r}},1)}\equiv B_{({\bm{r}},23)},\ \ \ \ B_{({\bm{r}},3)}\equiv B_{({\bm{r}},12)},
\ \ \ \ B_{({\bm{r}},2)}\equiv B_{({\bm{r}},31)}.
\end{equation}
With these conventions in place, we can introduce a model dual
to the Ising model in any dimension:
\begin{equation}\label{anyd_dual_ising}
H_{\sf I}^D=\sum_{\bm{r}}\ \left(h\ A_{\bm{r}} +\sum_{\nu=1}^d\ J\sigma^z_{({\bm{r}},\nu)}\right).
\end{equation}
The duality follows from these observations:
The vertex operators \(A_{\bm{r}}\) anti-commute with exactly \(2d\) spins \(\sigma^z_{({\bm{r}},\nu)}\),
just as the spins \(\sigma^x_{\bm{r}}\) of the Ising model of Equation \eqref{anyd_ising}
anti-commute with exactly \(2d\) bonds \(\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\).
Similarly, the spins \(\sigma^z_{({\bm{r}},\nu)}\) anti-commute with just two
vertex operators \(A_{\bm{r}}\) and \(A_{{\bm{r}}+\bm{e_\nu}}\), just as
\(\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\) anti-commutes with \(\sigma^x_{\bm{r}}\)
and \(\sigma^x_{{\bm{r}}+\bm{e_\nu}}\) only (a classical analogue of this duality
was introduced by Wegner in Reference \cite{wegner_ising}).
On the other hand, the relation
\begin{eqnarray}
\sigma^y_{\bm{r}}\sigma^y_{{\bm{r}}+\bm{e_\nu}}=-\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\
\sigma^x_{\bm{r}}\sigma^x_{{\bm{r}}+\bm{e_\nu}},
\end{eqnarray}
shows that the bond algebra of the Heisenberg model of Equation
\eqref{anyd_heisenberg_model} is a sub-algebra of the bond algebra of the
Ising model. Hence we can transfer the duality of Equation \eqref{anyd_dual_ising}
to the Heisenberg model:
\begin{equation}\label{dual_to_heisenberg}
H_{\sf H}^D=J\sum_{\bm{r}}\sum_{\nu=1}^d\ (A_{\bm{r}} A_{{\bm{r}}+\bm{e_\nu}}-
A_{\bm{r}}\sigma^z_{({\bm{r}},\nu)}A_{{\bm{r}}+\bm{e_\nu}}+\sigma^z_{({\bm{r}},\nu)}).
\end{equation}
The duality mapping reads,
\begin{equation}
A_{\bm{r}} A_{{\bm{r}}+\bm{e_\nu}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{\bm{r}}\sigma^x_{{\bm{r}}+\bm{e_\nu}}
,\ \ \ \ \ \ \ \ \ \ \ \ \
\sigma^z_{({\bm{r}},\nu)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}.
\end{equation}
Notice that \(H_{\sf H}\) has one
spin degree of freedom per lattice site, while \(H_{\sf H}^D\) has
\(d\) (one per link). Thus \(H_{\sf H}\) and \(H_{\sf H}^D\) do not
act on state spaces of the same dimensionality when $d>1$, and cannot be dual in the
strict sense of definition \eqref{algebraically_dual}. In order to
resolve this dilemma, one must appreciate that \(H_{\sf H}^D\) has a large group of
{\it gauge (local) symmetries} that is not shared by the Heisenberg
model (that has only global symmetries). All of the plaquette operators
\(B_{({\bm{r}}, \mu \nu)}\) defined in Equation \eqref{general_plaquette}
commute with \(H_{\sf H}^D\),
\begin{equation}
[B_{({\bm{r}},\mu\nu)},\ H_{\sf H}^D]=0,
\end{equation}
and rigorously speaking, \(H_{\sf H}\) and \(H_{\sf H}^D\) are
dual {\it up to the complete elimination of these gauge symmetries}
(Appendix \ref{appG} presents a version of this statement that can
be checked numerically. Note also that this discussion also applies to the
duality of Equation \eqref{anyd_dual_ising} between $H_{\sf I}$ and $H_{\sf I}^D$
for $d>1$).
This crucial refinement of the concept of duality will be discussed
at length in Section \ref{sec3.12} and will justify the need for homomorphisms,
as opposed to isomorphisms,
in the more general case.
Let us notice in closing this line of arguments, that there are also
several examples of {\it self-dual} models with a {\it non-Abelian symmetry group},
most notably the \(d=1\) vector Potts ($p$-clock) model of Section \ref{sec4.1},
and \(d=3\) \(\mathbb{Z}_p\) gauge theories of Section \ref{sec6.3}.
The fact that these models
have non-Abelian symmetries seems to have gone unnoticed in the literature.
We consider next a connection between self-dualities and non-Abelian
groups of a completely different character. Self-duality isomorphisms
connect one and the same Hamiltonian at different regions in coupling space, and
taken together they close a self-duality group, because the composition of
two self-dualities is another self-duality. This group acts linearly on the
bond algebra of the Hamiltonian, and can be either Abelian or non-Abelian.
To illustrate the premise, consider Kitaev's ``honeycomb
model'' \cite{Kitaev2006, nussinov_chen, bondDec08, pachosannals},
defined by the $S=1/2$ Hamiltonian,
\begin{eqnarray}
H_{\sf Kh}&=&-J_x\sum_{x{\sf-bonds}}\ \sigma^x_{i}\sigma^x_{j} -J_y
\sum_{y{\sf -bonds}}\sigma^y_{i}\sigma^y_{j} -J_z\sum_{z{\sf -bonds}}
\sigma^z_{i}\sigma^z_{j} \nonumber \\
&=& - \sum_{\langle ij \rangle} J_{\mu} \sigma_{i}^{\mu} \sigma_{j}^{\mu},
\ \ \ \ \ \ {\bm e}_{\mu} || (\vec{j} - \vec{i}),
\label{HK}
\end{eqnarray}
on a honeycomb lattice.
Here the spins are located at the sites $i,j$ (we use
the notation that is standard in the literature
to avoid any confusion). The three
nearest neighbor directions on the honeycomb lattice (at 120 degrees
relative to one another) are denoted by the
indices $\mu = x,y,z$ in Equation (\ref{HK}), see Figure
\ref{Kitaev_honeycomb}.
\begin{figure}[h]
\begin{center}
\includegraphics[width=5.2in]{./fig1+2.eps}
\end{center}
\caption{Kitaev's honeycomb model features $S=1/2$ spins
represented by a Pauli matrices $\vec{\sigma}_{k}$.
The model has three types of bonds, indicated by
the letter $\mu=x,y,z$, that represent the bond operators
$\sigma^\mu_i \sigma^\mu_j$. $\Phi_{\sf d}$ stands
for the duality mapping that realizes the exchange \(
\sigma^x_i \sigma^x_j\leftrightarrow\sigma^y_i \sigma^y_j\), and
that will be denoted in what follows as \(P_{yxz}\).}
\label{Kitaev_honeycomb}
\end{figure}
The Hamiltonian \(H_{\sf Kh}\) admits several simple self-dualities
that exchange any two of its couplings \(J_x,\ J_y\), and \(J_z\), and more
general permutations as well. Let \(\tau\in{\cal S}_3\), the group
of permutations of three elements, be the
permutation \(x,y,z\mapsto \tau(x), \tau(y),\tau(z)\). Then we denote
the corresponding self-duality mapping by $P_{\tau(x)\tau(y)\tau(z)}$,
that realizes the exchange \(J_x,J_y,J_z\mapsto J_{\tau(x)},
J_{\tau(y)},J_{\tau(z)}\)
(for example, the self-duality shown in Figure \ref{Kitaev_honeycomb}
will be denoted by $P_{yxz}$). We see that this family of
self-dualities affords a representation of the non-Abelian group of permutations
in the space of bonds of the model, but this group does not commute
with \(H_{\sf Kh}\) unless the Hamiltonian is fine-tuned to be at the self-dual
line \(J_x=J_y=J_z\).
We can write down representations for the pairwise
permutations. For instance, a global rotation about the $\sigma^z$ axis
by 90 degrees will exchange $\sigma^x \sigma^x$ with $\sigma^y \sigma^y$
and viceversa. It follows that
\begin{equation}
P_{yxz} = \exp\left [i \frac{\pi}{4} \sum_{j=1}^{N} \sigma_{j}^{z}\right ].
\label{pair_permute}
\end{equation}
The three pairwise permutations $\{P_{yxz}, P_{xzy}, P_{zyx}\}$ can be
represented as (non-commuting) rotations by 90 degrees. Any permutation
of the three bond types (or any bonds more generally in other systems)
can, of course, be written as a product of pairwise permutation
operators of the form of Equation (\ref{pair_permute}). For instance,
\begin{equation}
P_{zxy} = P_{zyx} P_{xzy}.
\end{equation}
Similar to
the group ${\cal S}_{3}$, we might embed other finite groups as acting on
a finite number of bond types.
\subsection{Exact dualities for finite systems}
\label{sec3.7}
Up to now we have only considered bond algebras of infinite systems.
This has advantages and disadvantages. On one hand, the
bond algebras are mathematically well defined, and the bond algebra
mappings of interest are typically simple. On the other hand,
it is of great interest to study the action of these mappings on operators
that are infinite combinations of bonds (e.g., the Hamiltonian).
This may be a concern because those operators need to be defined in the
infinite tensor product space where the bond algebra acts. We saw in Section
\ref{sec3.5} some of the problems that can arise from trying to extend the
action of bond algebra mappings to infinite combinations of bonds. Let us
take a look at these problems from a slightly different perspective that
will be useful later in this section.
It is standard practice to argue that the quantum Ising chain
of Equation \eqref{infinite_ising_transverse} has a \({\mathbb{Z}}_2\) symmetry generated by
\begin{equation}
Q=\prod_{i=-\infty}^{\infty}\ \sigma^x_i,\ \ \ \ \ \ [H_{\sf I}, Q]=0.
\end{equation}
Now, since formally we can write \(\mathbb{1}=\prod_{i=-\infty}^{\infty}
\ \sigma^z_i\sigma^z_{i+1}\), it would seem that the mapping of
Equation \eqref{aut_ising1} satisfies
\begin{equation}\label{phi_paradox}
\Phi_{{\sf d}}(\mathbb{1})=\prod_{i=-\infty}^{\infty}\
\Phi_{{\sf d}}(\sigma^z_i\sigma^z_{i+1})=Q.
\end{equation}
Since \(\Phi_{{\sf d}}(\mathbb{1})=\mathbb{1}\) must
hold true as well, it would seem that \(\Phi_{{\sf d}}\) is a multivalued mapping.
This problem was already
pointed out in Section \ref{sec2.2} from a different but equivalent perspective.
In general, duality mappings established in the
limit of infinite size, or in the continuum as in QFT (see
Section \ref{sec6}), are well defined on finite combinations of bonds,
but have ill-defined actions on for example global symmetries that involve infinite
combinations of bonds.
The practical solution to these problems is to work with bond algebras
of finite-size systems, and eventually take the thermodynamic limit if
one is interested in the infinite-size system.
But typically, standard boundary
conditions (BCs) (open, periodic, anti-periodic, etc.)
may spoil duality properties that are apparent in the infinite- size limit.
For example, both open and periodic BCs,
\begin{eqnarray}
H_{\sf I}^o&=&\sum_{i=1}^{N}\ h\sigma^{x}_i + \sum_{i=1}^{N-1}\
J\sigma^z_{i}\sigma^z_{i+1}, \ \ \ \ \ \ (\mbox{open BCs}),\
\ \ \ \ \mbox{}\label{io}\\
H_{\sf I}^c&=&H_{\sf I}^o+J\sigma^z_N\sigma^z_1, \ \ \ \ \ \
(\mbox{periodic (toroidal) BCs}), \label{ic}
\end{eqnarray}
spoil the self-duality \(J\leftrightarrow h\) of the quantum Ising chain,
and the same happens with many other (self-)dualities.
On the other hand, BCs can help to restore
in finite-size systems properties of the thermodynamic limit like
translation invariance. Similarly,
bond algebras can be exploited to find in a systematic way
{\it model specific} BCs that restore duality properties \cite{con}.
This is an impressive advantage of the bond-algebraic
over the traditional approach, because it puts models that are {\it exactly}
dual or self-dual at reach of computer simulations (the role
of BCs in connection to dualities was noticed from
time to time in the literature in the context of specific models,
see for example \cite{horn,hinrichsen}). Let us see how this works
with the models of Equations \eqref{io} and \eqref{ic}.
More complicated examples will be discussed in later sections.
Intuitively speaking, the Hamiltonian \(H_{\sf I}^o\) of Equation
\eqref{io} should not be self dual, because
it has $N$ bonds $\sigma^x_i$, but only $N-1$ bonds $\sigma^z_i\sigma^z_{i+1}$.
This suggests adding a bond
\begin{equation}\label{finiteisingsd}
H_{\sf I}^o\ \rightarrow\
\tilde{H}_{\sf I}^o=H_{\sf I}^o+\ J\sigma^z_N
\end{equation}
(see Figure \ref{automorphismfising}) that becomes {\it irrelevant in the
thermodynamic limit}, from the standpoint of bulk properties.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.65\columnwidth]{./automorphismfising.eps}
\end{center}
\caption{Two finite-size ($N=4$ sites) quantum Ising chains with
self-dual BCs that break \(Z_2\) invariance, connected
by the self-duality isomorphism \(\Phi_{{\sf d}}\) of Equation \eqref{automorphismfi}.
The big circle at the rightmost end of the chains represents the boundary correction
$\sigma^z_4$.}
\label{automorphismfising}
\end{figure}
The next step is to check that \(\tilde{H}_{\sf I}^o\)
is self-dual. To see this, we notice that if the model admits a
self-duality mapping \(\Phi_{\sf d}\), it must be that
\begin{eqnarray}
\Phi_{\sf d}(\sigma^x_1)=\sigma^z_N,
\end{eqnarray}
due to the structure of relations among bonds.
Next, to compute $\Phi_{\sf d}(\sigma^z_1\sigma^z_2)$, notice that
it must be one of the $\sigma^x$s, and that it must
anti-commute with $\Phi_{\sf d}(\sigma^x_1)=\sigma^z_N$. Thus it must be
that $\Phi_{\sf d}(\sigma^z_1\sigma^z_2)=\sigma^x_N$. Reasoning in this way,
we can reconstruct the full self-duality isomorphism
\begin{eqnarray}\label{automorphismfi}
\sigma^x_1&\ \stackrel{\Phi_{\sf d}}{\longrightarrow}\ & \sigma^z_N
\nonumber\\
\sigma^x_i&\ \stackrel{\Phi_{\sf d}}{\longrightarrow}\ &
\sigma^z_{r(i)}\sigma^z_{r(i)+1},\ \ \ \ \ \ i=2,3,\cdots,N\nonumber\\
\sigma^z_N&\ \stackrel{\Phi_{\sf d}}{\longrightarrow}\ &
\sigma^x_1\nonumber\\
\sigma^z_i \sigma^z_{i+1} & \ \stackrel{\Phi_{\sf d}}{\longrightarrow}\ &
\sigma^x_{r(i)},\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ i=1,2,\cdots,N-1,
\end{eqnarray}
$r(i)$ represents the {\it inversion} map
\begin{eqnarray}
r(i)=N+1-i.
\label{inversion_map_eq}
\end{eqnarray}
Notice that $\Phi_{{{\sf d}}}^2=\mathbb{1}$, the identity map.
In general, as discussed in Section \ref{sec3.9},
$\Phi_{{{\sf d}}}^2$ is related to a symmetry of the model under consideration.
The boundary term $J \sigma^z_N$ also makes it possible to compute
finite dual variables. The extra bond $\sigma^z_N$ guarantees that
the individual spins $\sigma^z_i$, \(i=1,\cdots,N\) are elements
in the bond algebra, since we can write
\begin{eqnarray}
\sigma^z_i=\sigma^z_N\times \sigma^z_{N}\sigma^z_{N-1}\times
\cdots\times \sigma^z_{i+1}\sigma^z_i.
\end{eqnarray}
Then the dual variables \(\mu^{x,y,z}_i=\Phi_{{{\sf d}}}(\sigma^{x,y,z}_i)\)
are
\begin{eqnarray}
\mu_1^x&=&\sigma^z_N\nonumber\\
\mu^x_i&=&\sigma^z_{r(i)} \sigma^z_{r(i)+1},\ \ \ \ \ \
i=2,3,\cdots,N,\nonumber \\
\mu^z_i&=&\prod_{m=i}^N\sigma^x_{r(m)}=\prod_{m=1}^{r(i)}\sigma^x_m.
\label{ssstring}
\end{eqnarray}
The mapping of Equation \eqref{automorphismfi} proves
that \(\tilde{H}_{\sf I}^o\) is indeed self-dual, and
is free of the mathematical inconsistencies embodied in
Equation \eqref{phi_paradox}. In particular, the self-dual boundary
term breaks the \({\mathbb{Z}}_2\) symmetry of the model, so the problem
inherent to Equation \eqref{phi_paradox} is no longer an issue. On the
other hand, one can find self-dual BCs that preserve the
\({\mathbb{Z}}_2\) symmetry \(Q=\prod_{i=1}^N\ \sigma^x_i\), namely
\begin{equation}
H_{\sf I}^o\ \rightarrow\ {\tilde{H'}}^{o}_{\sf I}=H_{\sf I}^o-h\sigma^x_N.
\end{equation}
The self-duality mapping for \({\tilde{H'}}_{\sf I}^{o}\) can be constructed just
as before, starting with \(\Phi_{{\sf d}}(\sigma^x_1)=\sigma^z_{N-1}\sigma^z_N\), but
since \(\sigma^x_N\) is no longer in \({\tilde{H'}}_{\sf I}^{o}\) , this will not determine the action
of \(\Phi_{{\sf d}}\) on \(\sigma^x_N\). This is important because we would like
to compute \(\Phi_{{\sf d}}(Q)\) and check that no inconsistency arises, and it is easy
to solve. The trick is to add \(\sigma^x_N\) to the list of bond generators, i.e.,
bond algebra,
{\it but not to} \({\tilde{H'}}_{\sf I}^{o}\), and extend the action of \(\Phi_{{\sf d}}\)
consistently. In this case, the result is that \(\Phi_{{\sf d}}(\sigma^x_N)=\sigma^z_1\),
and so
\begin{equation}
Q^D\equiv \Phi_{{\sf d}}(Q)= \sigma^z_N,\ \ \ \ \mbox{so that}\ \
[Q^D,\ {\tilde{H'}}_{\sf I}^{o}]=0.
\end{equation}
The self-duality exchanges the two non-trivial symmetries of the model.
The discussion of previous paragraphs illustrates very general
features of the problem of constructing (self-)dual boundary terms, features
that we will find also in more complex models in higher dimensions.
In general, {\it (self-)dual BCs are not unique}, and different
choices break and/or preserve different symmetries. This is intimately
connected to the topic of Section \ref{sec3.9}, and it is important in
practice to remember that the action of
a duality on a non-local symmetry cannot be understood with any precision in the
formal limit of infinite size or in the continuum (where, on the other hand,
dualities are most easily spotted!). For that, one has to choose the
(self-)dual BC that is best suited to the problem at hand,
to consider afterwards the action of the duality on the symmetries. Let us
illustrate next self-dual BCs that preserve translation invariance.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.45\columnwidth]{./isingperiodic.eps}
\end{center}
\caption{Quantum Ising chain featuring \(N\) sites,
with periodic BCs.}
\label{isingperiodic}
\end{figure}
In contrast to the open chain, the closed Ising chain
\(H_{\sf I}^c\) of Equation \eqref{ic} does not suffer from any bond
counting mismatch, so one could think that
\begin{equation}\label{circle_half_tr}
\sigma^x_i\ \mapsto\ \sigma^z_i\sigma^z_{i+1},\ \ \ \ \ \
\sigma^z_i\sigma^z_{i+1}\ \mapsto\ \sigma^x_{i+1}, \ \ \ \ \ \
(\mbox{with } N+1\equiv1),
\end{equation}
defines a self-duality. Unfortunately,
this is incorrect. If this mapping were an isomorphism, it would map
\begin{equation}
\mathbb{1}=\sigma^z_1\sigma^z_2\times\cdots\times\sigma^z_N
\sigma^z_1\ \mapsto\ \sigma^x_1\cdots\sigma^x_N,
\end{equation}
but isomorphisms can only map the identity \(\mathbb{1}\) to itself.
Fortunately, this analysis suggests the solution to the problem.
Let us change the boundary term as
\begin{equation}
H_{\sf I}^c\ \rightarrow\ \tilde{H}_{\sf I}^c=H_{\sf I}^o
+J\sigma^z_N Q\sigma^z_1,
\end{equation}
with
\begin{equation}
Q=\prod_{m=1}^N\sigma^x_m.
\end{equation}
Then, {\it with the understanding that \(\sigma^z_N\sigma^z_1\)
should be replaced by \(\sigma^z_N Q \sigma^z_1\)},
the mapping of Equation \eqref{circle_half_tr} does define a self-duality
isomorphism for \(\tilde{H}_{\sf I}^c\).
The boundary correction \(\sigma^z_N\sigma^z_1\rightarrow\sigma^z_NQ\sigma^z_1\)
seems to break the translational invariance of \(H^c_{\sf I}\). To see that
this is not the case, let us fix a uniform notation \(z_i\equiv\sigma^z_i\sigma^z_{i+1},\
i=1,\cdots,N-1\), \(z_N\equiv\sigma^z_NQ\sigma^z_1\), and compute the action
of \(\Phi_{{\sf d}}^2\) from Equation \eqref{circle_half_tr},
\begin{equation}
\Phi_{{\sf d}}^2(\sigma^x_i)=\sigma^x_{i+1}, \ \ \ \ \ \
\Phi_{{\sf d}}^2(z_i)=z_{i+1}, \ \ \ \ \ \ \mbox{with}\ \ i+N \equiv i.
\end{equation}
We see that \(\Phi_{{\sf d}}^2\) is in fact the generator of translations. When we
explain in the next section that \(\Phi_{{\sf d}}\) is unitarily implementable, this will
afford the proof that \(\tilde{H}_{\sf I}^c\) has translation invariance.
One easy way to check the correctness of the self-duality mapping
of Equation \eqref{circle_half_tr} (with corrected boundary term)
is to compute the dual variables, and
check that they satisfy the correct Pauli algebra. The problem with this
plan is that all the bonds in \(\tilde{H}_{\sf I}^c\) commute with the
symmetry \(Q\), and so the individual spins
\(\sigma^z_i\), \(i=1,\cdots, N\) cannot possibly be in the bond algebra.
The solution is to add one spin, say \(\sigma^z_N\) to
the set of bond generators, {\it but not to \(\tilde{H}_{\sf I}^c\)}.
In this way, we extend the bond algebra to include all the \(\sigma^z_i\)
without spoiling the fact that \(\tilde{H}_{\sf I}^c\) is self-dual,
provided we can extend the action of the self-duality isomorphism
to \(\sigma^z_N\) in a consistent fashion.
For the sake of concreteness, let us see how this works when there
are only \(N=3\) spins in the chain. Then the self-duality isomorphism
\begin{eqnarray}
&&\sigma^x_1\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_1\sigma^z_2,\ \ \ \ \ \ \ \ \ \ \
\sigma^z_1\sigma^z_2\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_2,\nonumber\\
&&\sigma^x_2\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_2\sigma^z_3,\ \ \ \ \ \ \ \ \ \ \
\sigma^z_2\sigma^z_3\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_3,\\
&&\sigma^x_3\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_3Q\sigma^z_1,\ \ \ \ \ \
\sigma^z_3Q\sigma^z_1\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_1,\nonumber
\end{eqnarray}
can be extended to \(\sigma^z_3\) as
\begin{equation}\label{pising_extension}
\sigma^z_3\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_1,
\end{equation}
preserving all the algebraic relations.
The dual variables that follow read
\begin{eqnarray}
&&\mu^z_1\equiv\Phi_{{\sf d}}(\sigma^z_1)=\sigma^z_1\sigma^x_2\sigma^x_3,
\ \ \ \ \ \
\mu^x_1\equiv\Phi_{{\sf d}}(\sigma^x_1)=\sigma^z_1\sigma^z_2,\nonumber\\
&&\mu^z_2\equiv\Phi_{{\sf d}}(\sigma^z_2)=\sigma^z_1\sigma^x_3,
\ \ \ \ \ \ \ \ \
\mu^x_2\equiv\Phi_{{\sf d}}(\sigma^x_2)=\sigma^z_2\sigma^z_3,\label{duali_circle}\\
&&\mu^z_3\equiv\Phi_{{\sf d}}(\sigma^z_3)=\sigma^z_1,
\ \ \ \ \ \ \ \ \ \ \ \
\mu^x_3\equiv\Phi_{{\sf d}}(\sigma^x_3)=\sigma^z_3Q\sigma^z_1.\nonumber
\end{eqnarray}
It is straightforward to extend this construction to \(N\) spins.
\subsection{Dualities as unitary transformations}
\label{sec3.8}
Last section's results strengthen our argument that the
bond-algebraic approach to dualities is truly practical.
One can always take a duality between infinite models, recast it as
a duality between finite renditions of those models, even check the
homomorphism
numerically, and be
free of all potential inconsistencies. Also,
the general definition of bond algebra and duality are meant
to settle the connection between dualities and unitary transformations.
It is not clear how to use them to construct explicitly
the unitaries that implement dualities (from now on, we use
the word {\it unitary} as short for unitary transformation).
Let us show for concreteness
how to build the self-duality unitary of the simplest self-dual
quantum Ising chain with only two sites.
Consider first the
Hamiltonian of Equation \eqref{finiteisingsd} with just
two sites, \(N=2\). Then (from Equation \eqref{automorphismfi}),
the self-duality isomorphism reads
\begin{eqnarray}\label{2sites}
\sigma^x_1\ \longleftrightarrow \sigma^z_2,\ \ \ \ \ \
\sigma^x_2\ \longleftrightarrow\ \sigma^z_1\sigma^z_2.
\end{eqnarray}
Let
\begin{equation}
|--\rangle,\ \ \ |-+\rangle,\ \ \ |+-\rangle,\ \ \
|++\rangle,
\end{equation}
be the simultaneous eigenstates of \(\sigma^x_1\) and \(\sigma^x_2\),
and let
\begin{eqnarray}
|-1\ -1\rangle&=&|\uparrow\downarrow\rangle,\ \ \ \ \
|-1\ 1\rangle=|\downarrow\uparrow\rangle,\\
|1,-1\rangle&=&|\downarrow\downarrow\rangle,
\ \ \ \ \ \ \ \ \ |1\ 1\rangle=|\uparrow\uparrow\rangle,\nonumber
\end{eqnarray}
be the simultaneous eigenstates of \(\sigma^z_1\sigma^z_2\) and
\(\sigma^z_2\). These latter bonds are dual to
\(\sigma^x_2\) and \(\sigma^x_1\) respectively,
thus we must have
\begin{eqnarray}
\mathcal{U}_{{\sf d}}&&= |-1\ -1\rangle\langle --|+|-1\ 1\rangle
\langle+-|+\\
&&\ \ \ \ \ \ \ |1\ -1\rangle\langle-+|+|1\ 1\rangle
\langle++|,\nonumber
\end{eqnarray}
or, in matrix form,
\begin{equation}
\mathcal{U}_{{\sf d}}=\frac{1}{2}
\begin{bmatrix}
1 &1 &1 &1\\
1 &-1 &-1 &1\\
1 &-1 &1 &-1\\
1 &1 &-1 &-1
\end{bmatrix}.
\end{equation}
The isomorphism \(\Phi_{{\sf d}}\) of Equation \eqref{2sites}
is its own inverse, and correspondingly,
\(\mathcal{U}_{{\sf d}}^2=\mathbb{1}\).
\subsection{Dualities and quantum symmetries}
\label{sec3.9}
A symmetry transformation is a modification of the observer's point of
view that does not change the outcome of an experiment performed on the
same system. Mathematically, in quantum mechanics, it is a mapping that
takes the Hilbert space of states $\cal H$ into an equivalent Hilbert
space. Wigner's theorem asserts that any transformation $\hat{T}$ which
preserves the transition probability between rays in the Hilbert space
$\cal H$, $| \Psi_1 \rangle, | \Psi_2 \rangle$,
\begin{equation}
|\langle \hat{T}^\dagger \Psi_1 | \hat{T} \Psi_2 \rangle|^2= |\langle
\Psi_1 | \Psi_2 \rangle|^2
\end{equation}
can be represented by a linear unitary or anti-linear anti-unitary map $U$
($U^\dagger=U^{-1}$) on $\cal H$. The discrete operation of
time-reversal is one of the few relevant examples in physics which
involves an anti-unitary operator. A symmetry transformation leaves the
Hamiltonian invariant. That is, $U \, H[\lambda] \, U^\dagger=
H[\lambda]$ or, equivalently, \([H,U]=0\).
Self-dualities have sometimes been alluded to as being symmetries
\cite{witten}. We think this is a misnomer for the following reasons:
As argued, self-dualities are usually unitarily implementable
transformations. However, while quantum symmetries are trivial isomorphisms
that leave the Hamiltonian $H$ invariant, self-duality transformations
do not preserve the form of $H$, but rather preserve its spectrum and
level degeneracies. In a sense, self-dualities capture non-trivial
isomorphisms aside from the
more trivial case of symmetries that leave $H$ itself invariant.
One basic connection between symmetries and dualities is that
symmetries control the variety of ways in which dualities can
manifest themselves. A closer look at any duality reveals that it could
be embodied in a wide variety of isomorphisms, but there is
a common denominator: these are all related by symmetry. This is
easy to understand on general grounds,
now that we know that dualities are unitary
equivalences. For suppose that you have two different dualities
\(\mathcal{U}_{{\sf d}}\) and \({\mathcal{U}'}_{{\sf d}}\) that connect Hamiltonians
\(H_1\) and \(H_2\),
\begin{equation}\label{H12H2}
\mathcal{U}_{{\sf d}} H_1\mathcal{U}_{{\sf d}}^\dagger=H_2,\ \ \ \ \ \mbox{and}
\ \ \ \ \ {\mathcal{U}'}_{{\sf d}} H_1{\mathcal{U}'}_{{\sf d}}^\dagger=H_2.
\end{equation}
Then
\begin{equation}
{\mathcal{U}'}_{{\sf d}}^\dagger\mathcal{U}_{{\sf d}} H_1\mathcal{U}_{{\sf d}}^\dagger{\mathcal{U}'}_{{\sf d}}
=H_1,\ \ \ \ \ \mbox{and}\ \ \ \ \
{\mathcal{U}'}_{{\sf d}}\mathcal{U}_{{\sf d}}^\dagger H_2\mathcal{U}_{{\sf d}}{\mathcal{U}'}_{{\sf d}}^\dagger=H_2,
\end{equation}
so that \({\mathcal{U}'}_{{\sf d}}^\dagger\mathcal{U}_{{\sf d}}\) is a symmetry of \(H_1\),
and \({\mathcal{U}'}_{{\sf d}}\mathcal{U}_{{\sf d}}^\dagger\) is a symmetry of \(H_2\).
Conversely if \(U_1\) (\(U_2\)) is a symmetry of \(H_1\) (\(H_2\)), and
\(\mathcal{U}_{{\sf d}}\) is a duality as above, then
\begin{equation}
U_2\mathcal{U}_{{\sf d}} U_1
\end{equation}
is a duality as well. These connections could be used to unveil
hidden symmetries through dualities.
The connection between dualities and symmetries is even stronger for
self-dual models \cite{con}. Suppose, for simplicity, that we have a Hamiltonian
$H[\lambda_1,\lambda_2,\cdots]$, dependent upon a set of couplings
$\lambda_\nu$, that is self-dual under the exchange
$\lambda_1\leftrightarrow\lambda_2$, that is,
\begin{eqnarray}\label{usym}
\mathcal{U}_{\sf
d}\, H[\lambda_1,\lambda_2,\cdots]\, \mathcal{U}^\dagger_{{\sf d}}=
H[\lambda_2,\lambda_1,\cdots],
\end{eqnarray}
with $\mathcal{U}_{{\sf d}}$ a unitary independent of the couplings.
$\mathcal{U}_{{\sf d}}$ relates to symmetries of $H$ in two ways.
First, it is clear that
\begin{eqnarray}
[H[\lambda_1,\lambda_2,\cdots],\ \mathcal{U}^{2n}_{{\sf d}}]=0,
\end{eqnarray}
i.e., $\mathcal{U}^{2n}_{{\sf d}}$ are symmetries of $H$ for any
$n=1,2\cdots$ up to the power that gives unity back.
We see that loosely speaking, a self-duality can
be seen as the square root of a symmetry.
Second, Equation \eqref{usym} shows that at the {\it self-dual point}
\begin{equation}
\lambda_1=\lambda_2\ \ \ \ \ \ \ \ (\mbox{self-dual point}) ,
\end{equation}
$\mathcal{U}_{{\sf d}}$ itself commutes with $H$. In other words,
$\mathcal{U}_{{\sf d}}$ {\it emerges as a new symmetry at the
self-dual point}. In fact, the full sequence of powers $\mathcal{U}_{{\sf d}},
\mathcal{U}^2_{{\sf d}}, \mathcal{U}^3_{{\sf d}}, \cdots$
is a sequence of quantum symmetries at the self-dual point.
The results just described suggest that self-dualities may
increase the symmetry of a model drastically at the self-dual point,
maybe even by becoming a continuous group of symmetries.
This is an especially attractive
possibility for models that exhibit a phase transition at the self-dual
point, but in fact, it can be excluded on general principles.
For suppose that one could find a self-duality transformation
$\mathcal{U}_{{\sf d}}(\theta)$ that depends on some set of continuous
coordinates $\theta$, so that
\begin{eqnarray}
\mathcal{U}_{{\sf d}}(\theta)\, H[\lambda_1,\lambda_2,\cdots]\,
\mathcal{U}^\dagger_{{\sf d}}(\theta)=H[\lambda_2,\lambda_1,\cdots]
\end{eqnarray}
for any value of $\theta\neq 0$, {\it independently of the values of
the couplings in $H$}. Such a group of self-duality unitaries would
become an extra continuous symmetry of the model at the self-dual
point. But this is impossible, because
$\mathcal{U}^2_{{\sf d}}(\theta)$ must be a symmetry always. Then, taking
$\theta=\epsilon$ infinitesimal so that $\mathcal{U}^2_{\sf
d}(\epsilon) \approx \mathbb{1}+2i\epsilon\cdot \breve{T}$, we see that the
generators $\breve{T}$ must always commute with $H$. But then
$\mathcal{U}_{\sf d}(\epsilon)\approx \mathbb{1}+i\epsilon\cdot \breve{T}$ must
commute with $H$ as well, rather than represent a self-duality.
The discussion above does not exclude the possibility that
$\mathcal{U}_{{\sf d}}$ may depend on the couplings in the Hamiltonian,
\begin{eqnarray}
\mathcal{U}_{{\sf d}}(\lambda_1,\lambda_2,\cdots) \,
H[\lambda_1,\lambda_2,\cdots] \, \mathcal{U}^\dagger_{\sf
d}(\lambda_1,\lambda_2,\cdots)= H[\lambda_2,\lambda_1,\cdots],
\end{eqnarray}
and this is in fact the case for the spin \(S=1/2\) XY model
discussed in Section \ref{sec3.11}.
But this would not turn the self-duality at the self-dual point
into continuous symmetry either. Rather, one would have a
discrete set of symmetries, one discrete set for each value of the
self-dual coupling $\lambda\equiv \lambda_1=\lambda_2$.
In closing, let us mention briefly two examples.
Consider first the finite, open, self-dual quantum Ising chain
$\tilde{H}^o_{\sf I}$ introduced
in Section \ref{sec3.7}, Equation \eqref{finiteisingsd}.
It is easy to verify that $\mathcal{U}^2_{{\sf d}}=\mathbb{1}$ (there is no
need to compute $\mathcal{U}_{{\sf d}}$ explicitly, just to note that
$\mathcal{U}_{{\sf d}}$ implements the mapping defined in Equation
\eqref{automorphismfi}). At the self-dual point $J=h$, $\mathcal{U}_{{\sf d}}$
becomes a non-trivial discrete symmetry of the model, the generator of
a $\mathbb{Z}_2$ symmetry group for $\tilde{H}^o_{\sf I}$.
This is especially interesting, since the standard \(\mathbb{Z}_2\) symmetry
of the Ising model is broken by the self-dual boundary term \(J\sigma^z_N\).
For the {\it infinite} quantum Ising chain, we have from Equation
\eqref{aut_ising1} that
\begin{equation}
\Phi_{\sf d}^2(\sigma^x_i)=\sigma^x_{i-1},\ \ \ \ \ \ \ \ \ \ \ \
\Phi_{\sf d}^2(\sigma^z_i\sigma^z_{i+1})
=\sigma^z_{i-1}\sigma^z_{i},
\end{equation}
Thus $\Phi_{\sf d}^{2}$ generates lattice translations to the left.
\subsection{Order and disorder variables for self-dual models}
\label{sec3.10}
Recognizing that self-dualities are unitary equivalences has consequences
that go beyond symmetry, and are intimately tied to the
behavior of the quantum
fluctuations that compete at a quantum phase transition.
For self-dual models, there is a natural way to associate a disorder
parameter to any order parameter (and viceversa), through the self-duality
unitary, and moreover, the eigenstates of the self-duality unitary are states at which
the expectation value of a pair of ``duality-conjugate" observables {\it are
equal}. While these states are not specially meaningful at general couplings,
at the self-dual point they can be chosen to be simultaneous eigenstates
of the Hamiltonian
(because the self-duality becomes a symmetry at the self-dual point).
The general setting we are going to consider in this section is that of
a self-dual Hamiltonian \(H[\lambda]\)
depending on any number of parameters
\(\lambda=(\lambda_1,\lambda_2,\cdots)\), with dual parameters \(\lambda^*\)
defined by
\begin{eqnarray}\label{dual_here}
H[\lambda^*]=\mathcal{U}_{{\sf d}}\,H[\lambda]\,\mathcal{U}_{{\sf d}}^\dagger.
\end{eqnarray}
The observable \({\cal O}_{{\sf d}}\) dual-conjugate to \({\cal O}\)
is defined by the equation
\begin{equation}
{\cal O}_{{\sf d}}=\mathcal{U}_{{\sf d}} {\cal O} \mathcal{U}_{{{\sf d}}}^\dagger
\end{equation}
For example, \(H[\lambda^*]\) is the dual-conjugate of \(H[\lambda]\),
and for the Ising models studied in Section
\ref{sec3.7}, the dual-conjugates of the spin operators
\(\sigma^x_i,\ \sigma^z_i\) are the dual variables
\(\mu^x_i,\ \mu^z_i\) of
Equations \eqref{ssstring} and \eqref{duali_circle}.
The first interesting consequence of this definition is that,
{\it relative to self-duality eigenstates \(|\phi_j\rangle\)},
\begin{equation}
{\mathcal{U}}_{{\sf d}}|\phi_j\rangle=e^{i\phi_j} |{\phi_j}\rangle,\ \ \ \ \ \
j=1,\cdots,{\sf dim}\mathcal{H},
\end{equation}
pairs of observables that are dual-conjugate have identical
expectation values,
\begin{equation}
\langle\phi_j|{\cal O}_{{\sf d}} |\phi_j\rangle=
\langle\phi_j|\mathcal{U}_{{\sf d}} {\cal O} \mathcal{U}_{{{\sf d}}}^\dagger
|\phi_j\rangle=\langle\phi_j|{\cal O} |\phi_j\rangle.
\end{equation}
This is especially interesting at the self-dual point \(\lambda_{\sf sd}=
\lambda_{\sf sd}^*\), where the states \(|\phi_j\rangle\) can be chosen
to be simultaneous eigenstates of \(\mathcal{U}_{{\sf d}}\) and
\(H[\lambda_{\sf sd}]\) (since
\(\mathcal{U}_{{\sf d}}\) is a symmetry of \(H\) at the self-dual point).
Next we would like to compare expectation values of dual-conjugate
pairs relative to arbitrary states \(|\psi\rangle\). For this it is convenient
to specialize the discussion to self-dualities that are their own inverses,
so that
\begin{equation}\label{very_sd}
\mathcal{U}_{{{\sf d}}}^\dagger=\mathcal{U}_{{\sf d}} .
\end{equation}
It is often possible to arrange for this to be the case, thanks to the
freedom in choosing \(\mathcal{U}_{{\sf d}}\) discussed in the previous section.
Then, we have on one hand that
\begin{equation}
\langle\psi|\mathcal{O}_{{\sf d}}|\psi\rangle=
\langle\psi_{{\sf d}}|\mathcal{O}|\psi_{{\sf d}}\rangle,
\end{equation}
where
\begin{equation}
|\psi_{{\sf d}}\rangle\equiv \mathcal{U}_{{\sf d}}|\psi\rangle.
\end{equation}
But thanks to Equation \eqref{very_sd}, a completely analogous
relation holds for \(\mathcal{O}\):
\begin{equation}
\langle\psi|\mathcal{O}|\psi\rangle=
\langle\psi|{\mathcal{U}}_{{\sf d}}^2\mathcal{O}{\mathcal{U}}_{{\sf d}}^2|\psi\rangle=
\langle\psi_{{\sf d}}|\mathcal{O}_{{\sf d}}|\psi_{{\sf d}}\rangle,
\end{equation}
It is in this specific sense that \(\mathcal{O}\) and \(\mathcal{O}_{{\sf d}}\)
show perfectly complementary behavior.
Let us apply these general results to quantum phase transitions.
Let \(\vert \Omega; \lambda\rangle\) denote a ground state for \(H[\lambda]\).
Then \(\mathcal{U}_{{\sf d}}\vert \Omega; \lambda\rangle\) is a ground state
for \(H[\lambda^*]\), that we denote \(\vert \Omega;\lambda^*\rangle\).
It follows that
\begin{eqnarray}
\langle \Omega;\lambda\vert {\cal O} \vert \Omega;\lambda\rangle &=&\langle \Omega
;\lambda^*\vert {\cal O}_{{\sf d}}\vert \Omega;\lambda^*\rangle,\ \ \ \ \ \ \mbox{and}\\
\langle \Omega;\lambda\vert {\cal O}_{{\sf d}} \vert \Omega;\lambda\rangle &=&\langle \Omega
;\lambda^*\vert {\cal O}\vert \Omega;\lambda^*\rangle.\nonumber
\end{eqnarray}
Hence, if the mean value of ${\cal O}$ happens to be related to the {\it
order parameter} associated with a phase transition that takes place as
the couplings \(\lambda\) are changed, it follows
immediately from the two relations above that \({\cal O}_{{\sf d}}\) represents
an operator related to the {\it disorder parameter}.
Consider for illustration the quantum Ising chain, and set
$\lambda \equiv J/h$, so that the dual $\lambda^*$,
resulting from the self-duality transformation $h \leftrightarrow J$,
is $\lambda^* = \lambda^{-1}$. Then we have that
\begin{equation}\label{sdsstring}
\langle 0;\lambda\vert \sigma^z_i\sigma^z_j\vert 0;\lambda\rangle=
\langle 0;\lambda^{-1}\vert \mu^z_i\mu^z_j\vert 0;\lambda^{-1}\rangle,\ \
\ \langle 0;\lambda\vert \mu^z_i\mu^z_j\vert 0;\lambda\rangle=
\langle 0;\lambda^{-1}\vert \sigma^z_i\sigma^z_j\vert 0;\lambda^{-1}\rangle.
\end{equation}
Equation \eqref{sdsstring} demonstrates that the string operator of
Equation \eqref{ssstring} is the disorder variable conjugate to the
order variable \(\sigma^z_i\) \cite{fradkin_susskind}. The relation
between our bond-algebraic approach to {\it quantum disorder variables}
and the work of Kadanoff and Ceva \cite{Kadanoff_Ceva} on (commutative)
algebras in the {\it classical} $D=2$ Ising model is elaborated on in
Section \ref{appB}. Our bond-algebraic approach generalizes the
work of \cite{Kadanoff_Ceva}. Here we would like to point out that in Reference
\cite{Kadanoff_Ceva}, it was argued (only in the context of the $D=2$
Ising model) that the product of an order and a neighboring disorder
variable should behave as a fermion. This is largely satisfied by our
operator order and disorder quantum variables for the quantum Ising
chain. If we define
\begin{eqnarray}
\gamma_i \equiv \sigma^z_i\mu^z_{i+1},
\end{eqnarray}
then it will be easy to check that the operator $\gamma_{i}$ represents
a Majorana fermion, that is, a Dirac fermion
that is self-conjugate (and consequently, its own anti-particle). For the
current purposes, it suffices to mention that Majorana fermions can be
expressed in terms of (spinless) fermion creation/annihilation
operators and satisfy the following anti-commutation relations
\begin{eqnarray}
\label{majorana_f}
\{\gamma_i, \gamma_j\}=2\delta_{i,j}.
\end{eqnarray}
This enables the standard Jordan-Wigner transformation \cite{GJW} that
maps $S=1/2$ spin degrees of freedom into spinless fermions (and
viceversa).
It seems to be a general feature of $d=1$ quantum models
that the product of the order and disorder variables satisfies simple and
interesting algebraic relations.
Unfortunately, this pattern seems to break down in higher dimensions.
\subsection{Emergent dualities}
\label{sec3.11}
Two standard ways to simplify models in condensed matter physics
are to restrict couplings to take very special values, and/or project
out some states of the full state space. A typical example is the
$t$-$J$ model \cite{auerbach}, that is obtained as a projection from the
Hubbard model in the strong-coupling limit. In this section,
we explain how emergent (self-)dual properties
can appear in the effective models that come out of such manipulations \cite{con},
even when the starting models are not (self-)dual to
start with.
We discuss the two scenarios (special couplings versus reduced state space)
separately for simplicity, but examples more complex than the ones
we are going to consider can well present a blend of both.
\subsubsection{Projective emergent dualities}
The projection of a Hamiltonian, and a corresponding bond algebra,
into a sector (subspace) \(\mathcal{W}\) of the full state space
\(\mathcal{H}\), produces a {\it new} bond algebra that may,
or may not, have new algebraic and duality properties.
For instance, bonds that do not commute in
general may commute when projected onto certain sectors, or vanish
(thus reducing the number of relations that characterize the algebra).
Hence it may well be that the projected bond algebra enjoys (self-)dualities
that are not available for the full model.
More precisely, a
projection will always change the structure of the
bond algebra, unless the projector
\(P_\mathcal{W}=P_\mathcal{W}^2\) commutes with all the bonds. To see this,
notice that the following relation always holds
\begin{equation}
P_\mathcal{W}(h_\Gamma+\lambda h_{\Gamma'})P_\mathcal{W}\ =\
P_\mathcal{W}h_\Gamma P_\mathcal{W}+\lambda P_\mathcal{W} h_{\Gamma'}
P_\mathcal{W}.
\end{equation}
So, projection always preserve the linear structure. Problems
can develop with respect to the multiplicative structure, since
\begin{equation}
P_\mathcal{W}(h_\Gamma h_{\Gamma'})P_\mathcal{W} = P_\mathcal{W}h_\Gamma
P_\mathcal{W}\ P_\mathcal{W} h_{\Gamma'}P_\mathcal{W}
\end{equation}
will generally hold only if for all $\Gamma$,
\begin{equation}
\label{hpw}
[h_\Gamma,P_\mathcal{W}]=0.\ \ \ \ \ \ \ \
\end{equation}
If this is the case, then
\begin{eqnarray}
P_\mathcal{W}(h_\Gamma h_{\Gamma'})P_\mathcal{W}\ =\
P_\mathcal{W}^2(h_\Gamma h_{\Gamma'})P_\mathcal{W}^2\ =\
P_\mathcal{W}h_\Gamma P_\mathcal{W}\ P_\mathcal{W}
h_{\Gamma'}P_\mathcal{W},
\label{ppp}
\end{eqnarray}
and the projection process preserves (to some extent)
the structure of the bond algebra (in other words,
the mapping \(\Phi(h_\Gamma)=P_\mathcal{W}h_\Gamma P_\mathcal{W}=
h_\Gamma P_\mathcal{W}\) is an algebra homomorphism).
However, it is unlikely that the projections of physical
interest will preserve the bond algebra
as in Equation \eqref{ppp}, and so we can expect that the
effective, projected model will have a different bond algebra
(Gauge models represent
the most important exception to this rule, see the next section).
We call {\it emergent dualities} those dualities that are brought
about by this change in the bond algebra due to a projection
(or a special fixing of the couplings, like in the next subsection),
to stress that these dualities {\it emerge} in some sector of the
theory but need not be exact relations for the full system.
In many
instances, the sector of interest is that of low energies. At low
temperatures, the system becomes more and more confined to this Hilbert
space sector (especially so when spectral gaps are present between the
low energy sector of $H$ and all other excited states). It is important
to appreciate that emergent dualities must be unitarily implementable,
just as ordinary dualities, with the extra freedom that the unitary
transformations need only be defined on certain subspaces such as that
spanned by the low energy states.
To make this lucid, we now consider two examples. The first
example is afforded by the elementary Hamiltonian
\begin{equation}
\label{hsl}
H_{\sf L}=L_{z}+\frac{1}{2},
\end{equation}
with the angular momentum operator \(L_{z}=-i \partial/\partial\theta\).
If one takes its domain to be the full Hilbert space of wave-functions on the
circle \(\langle \theta|\psi\rangle=\psi(\theta)\in{\cal L}^2(U(1))\),
then \(H_{\sf L}\) is not bounded below.
However, $H_{\sf L}$ has an {\it emergent} duality to the standard harmonic
oscillator
\begin{equation}
H_{\sf HO}=a^\dagger a+\frac{1}{2}, \ \ \ \ \ \ \ \ \ \
[a,a^\dagger]=1,
\end{equation}
on the sector of states of non-negative angular momentum.
To see how this works, consider the algebra,
\begin{equation}
[L_{z},\ A]=-A,\ \ \ \ \ \ \ \ [L_{z},\ A^\dagger]=A^\dagger,
\ \ \ \ \ \ \ \ [A,A^\dagger]=0
\end{equation}
where \(A\) and \(A^\dagger\) are the ladder operators associated with
\(L_{z}\), and act on wave-functions by multiplication,
\(\langle\theta\vert A\vert\psi\rangle=e^{-i\theta}\psi(\theta)\).
Then, if we let \( \{\vert m\rangle \}\) denote angular momentum eigenstates,
\begin{eqnarray}
L_{z}\vert m\rangle=\ m\vert m\rangle,\ \ \ \ \ \ \ \ \ \
\langle\theta\vert m\rangle=\frac{1}{\sqrt{2\pi}}e^{im\theta},
\end{eqnarray}
we have that \(\langle\theta\vert A|m\rangle=\langle\theta\vert m-1\rangle\) and
\(\langle\theta\vert A^\dagger|m\rangle=\langle\theta\vert m+1\rangle\).
We would like to study this algebra on the subspace spanned by the
states \(\vert m\rangle\) of non-negative angular momentum only, \(m=0,1,2,
\cdots\ \). If \(P\) denotes the orthogonal projector onto this
subspace, then it is not difficult to check that
\begin{equation}
\label{LPAP}
[L_P,\ A_P]=-A_P,\ \ \ \ \ \ \ \ [L_P,\ A_P^\dagger]=A_P^\dagger,
\ \ \ \ \ \ \ \ [A_P,\ A_P^\dagger]=\ P_0,\ \ \
\end{equation}
where \(P_0=\vert 0 \rangle\langle 0\vert\) denotes the projector onto
the eigenspace of \(0\) angular momentum (the ground state sector
for $PH_{\sf L}P$). In Equation \eqref{LPAP}, we employ the
shorthand \(M_P=PMP\) for general projected operators (and we further
abbreviate, $L_{P} = P L_{z} P$). The algebra of Equation \eqref{LPAP}
is isomorphic to that of the harmonic
oscillator \cite{remark_HS}, as the mapping
\begin{eqnarray}
&&L_P \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ a^\dagger a,\nonumber\\
&& A_P\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ (a^\dagger a+1)^{-1/2}a,
\ \ \ \ \ \ \ \ \ \
A_P^\dagger\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ a^\dagger (a^\dagger
a+1)^{-1/2}.
\label{Losc}
\end{eqnarray}
shows. Thus \(\Phi_{{\sf d}}(PH_{\sf L}P)=H_{\sf HO}\) embodies an
elementary {\it emergent duality}.
\begin{figure}[h]
\begin{center}
\includegraphics[width=.43\columnwidth]{./column_state} \hspace*{1cm}
\includegraphics[width=.43\columnwidth]{./top_eq_column_st.eps}
\end{center}
\caption{Two dimer coverings of the square lattice. Dimer coverings label
an orthonormal basis of states for the state space of the quantum dimer model
defined in the text.}
\label{qdm1}
\end{figure}
A more interesting example of an {\it emergent self-duality} is
afforded by the Quantum Dimer Model \cite{con}. This model's
Hamiltonian \cite{RK}
\begin{eqnarray}
H_{\sf QDM}&=&\sum_{\Box} \left[
-t\left(\left|\vv\right\rangle\left\langle\hh\right | +
\left|\hh\right\rangle\left\langle\vv \right | \right)\right . _{\Box}
\nonumber\\
&+&\left.v\left(\left|\vv\right\rangle\left\langle\vv\right|+
\left|\hh\right\rangle\left\langle\hh \right|\right)_{\Box} \right],
\label{QDM}
\end{eqnarray}
acts on a state space spanned by orthonormal states labelled by dense dimer
coverings of a lattice, see Figure \ref{qdm1} (the sum
\(\sum_{\Box}\) must include all the elementary plaquettes $\Box$). It
contains both a kinetic $t$ term that flips one dimer tiling of any
plaquette to another (a horizontal covering to a vertical one and
viceversa), and a potential $v$ term. On every plaquette, the potential
term operator is equal to the square of the kinetic term \cite{nn}.
At the so-called Rokhsar-Kivelson (RK) point $t=v$ \cite{RK}, the ground
states are equal amplitude superpositions of dimer coverings. If $P_{g}$
is the projection operator onto the ground state sector, then on any
plaquette,
\begin{eqnarray} \!\!\!\!\!\!\! \!\!\!\!\!
P_{g} [\left(\left|\vv\right\rangle\left\langle\hh\right| +
\left|\hh\right\rangle\left\langle\vv \right|\right)]_{\Box} P_{g}
= P_{g} [\left(\left|\vv\right\rangle\left\langle\vv\right|+
\left|\hh\right\rangle\left\langle\hh \right|\right)]_{\Box} P_{g} =
x_{\Box} P_{g} ,
\end{eqnarray}
with $x_{\Box} = 0$ or $1$ on the particular plaquette $\Box$. At the
RK point, the projected Hamiltonian becomes $P_{g} H_{\sf QDM} P_{g}
=0$. Since both the kinetic ($t$) and potential ($v$) terms are given by
$x_{\Box} P_{g}$ within the ground state sector, they can be
interchanged without affecting the bond algebra. This self-duality
emerges exclusively in the ground state sector of the model at the RK
point.
\subsubsection{Coupling-dependent emergent dualities}
\label{sec3.10.2}
Next we discuss dualities that emerge in some specific region
of the space of parameters of some models. To some extent, this
notion is already included in the general concept of bond-algebraic
duality,
but the examples we have studied do not emphasize it sufficiently.
The point to notice
is that we can choose the bond generators to
include the coupling and external fields in a non-trivial fashion.
Then the structure of the corresponding {\it bond algebra depends on
those couplings} as well, and varies with them. Thus it is possible that
some specific values of the couplings will afford (self-)dual
properties that may be absent in general.
Let us present an example to clarify this idea. In what follows, we
show that the spin \(S=1/2\) XY model
\begin{equation}\label{hxyn}
H_{\sf XY}= \sum_i\ (J_x\sigma^x_i\sigma^x_{i+1}+\
J_y\sigma^y_i\sigma^y_{i+1} +\ \bar{h} \sigma^z_i)
\end{equation}
presents an {\it emergent self-duality on the surface \(J_x=J=1/J_y\)}
in coupling space \((J_x,J_y,\bar{h}\)) \cite{hinrichsen}.
The reason is that on this surface we can re-write
\begin{equation}
H_{\sf XY}[J,h]=\sum_i\ (A_i(J,h)+B_i(J^{-1},h^{-1}))
\end{equation}
as a sum of bonds
\begin{equation}
A_i(J,h)\equiv J\sigma^x_{i}\sigma^x_{i+1}+\ h\sigma^z_i,\ \ \ \ \ \
B_i(J^{-1},h^{-1})=J^{-1}\sigma^y_i\sigma^y_{i+1}+\ h^{-1}\sigma^z_{i+1},
\end{equation}
provided we split \(\bar{h}=h+h^{-1}\),
thus generating a bond algebra symmetrical under the exchange $J\leftrightarrow h$
(so that after the duality transformation the dual magnetic field reads \(\bar{h}^*=J+J^{-1}\)).
In what follows, we write \(A_i, B_i\) for \(A_i(J,h), B_i(J^{-1},h^{-1})\),
and \(\hat{A}_i, \hat{B}_i\) for \(A_i(h,J), B_i(h^{-1},J^{-1})\).
The self-duality \(J\leftrightarrow h\) is already intuitively clear from the set
of relations that characterize the bond algebra generated by the \(A_i, B_i\):
\begin{eqnarray}
&&\ \ \ \ \ \ A_i^2=J^2+h^2,\ \ \ \ B_i^2=J^{-2}+h^{-2},\nonumber\\
&&\ \ \ \ \ \ \{B_i,A_i\}=0, \ \ \ \ \{B_i,A_{i+1}\}=2,\nonumber\\
&& A_i\ A_{i+1}\ A_i\ A_{i+1}\ +\ A_{i+1}\ A_i\ A_{i+1}\ A_i =\
2(J^4+h^4)\label{xy_relation}\\
&&B_i\ B_{i+1}\ B_i\ B_{i+1}\ +\ B_{i+1}\ B_i\ B_{i+1}\ B_i=\
2(J^{-4}+h^{-4})\nonumber\\
&&A_i\ B_{i+1}\ A_i\ B_{i+1}\ +\ B_{i+1}\ A_i\ B_{i+1}\ A_i=\
2(J^2h^{-2}+h^2J^{-2})\nonumber
\end{eqnarray}
(all other pairs of generators commute). Rigorously, the self-duality
follows from the observation that the mapping
\begin{equation}\label{iso_xy}
A_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \hat{A}_i,\ \ \ \ \ \ B_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \hat{B}_i,
\end{equation}
is an isomorphism, simply because the dual bonds
\(\hat{A}_i, \hat{B}_i\)
satisfy the same relations of Equation \eqref{xy_relation},
so that \(H_{\sf XY}[h,J]=\sum_i\ (\hat{A}_i+\hat{B}_i)\) is unitarily
equivalent to \(H_{\sf XY}[J,h]\).
The self-duality isomorphism of Equation \eqref{iso_xy} reduces to the
identity map at the self-dual point \(J=h\), because the bonds
\(\hat{A}_i, \hat{B}_i\) become identical to the \(A_i, B_i\)
there. This has interesting consequences in the dual variables
to be computed below. Also, the product
\begin{eqnarray}
iA_jB_j=(Jh^{-1}+hJ^{-1})\sigma^x_j\sigma^y_{j+1}=
i\hat{A}_j\hat{B}_j,
\end{eqnarray}
is invariant under duality and can be added to the Hamiltonian
with arbitrary couplings without spoiling its self-dual structure.
The dual variables for this problem are particularly interesting
because they depend on the coupling parameters, and have an additive
and multiplicative structure (while every other set of dual variables
considered in this paper are purely multiplicative). We will indicate
how to construct them in general, and write them explicitly for the
simplest case of just two sites (\(N=2\)).
The starting point is to notice
that the isomorphism of Equation \eqref{iso_xy}
works just as well when restricted to a finite bond
algebra generated by the \(2N\) bonds \(A_i, B_i\) with \(i=1,\cdots,N\).
This shows that the {\it finite} rendition \(H_{\sf XY}=\sum_{i=1}^{N-1}
\ (A_i+B_i)\) is self-dual (because of self-dual BCs, this
Hamiltonian has \(\sigma^z_1\) coupled to \(h\) only, and \(\sigma^z_{N}\)
coupled only to \(h^{-1}\)). The next step is to figure out whether
the single spins \(\sigma^x_i, \sigma^y_i\), \(i=1,\cdots,N\)
are elements in the bond algebra generated by \(A_i, B_i\), \(i=1,\cdots,N-1\).
Clearly they are not, since every one of these bonds commutes
with \(\prod_{i=1}^{N}\ \sigma^z_i\). As in Section \ref{sec3.7},
the solution to this problem is to enlarge the bond algebra by
adding generators that do not spoil the symmetry under exchange of $J$
and $h$.
A simple analysis shows that two such operators are
$\sigma^y_1$ and $\sigma^x_N$, since they
commute with almost every other bond, except
for $A_{1}$ and $B_{N-1}$,
\begin{equation}
\{A_1,\sigma^y_1\}=0,\ \ \ \ \ \ \{B_{N-1}, \sigma^x_N\}=0,
\end{equation}
and these relations (being independent of the couplings)
preserve the symmetry in $J$ and $h$.
In other words, the extended Hamiltonian
\begin{equation}\label{sdXY}
H_{\sf XY}[J,h,\tilde{h}^y,\tilde{h}^x]=\tilde{h}^y\sigma^y_1+
\tilde{h}^x\sigma^x_N+ \sum_{i=1}^{N-1}\
(J\sigma^x_i\sigma^x_{i+1}+h\sigma^z_i+ J^{-1}\sigma^y_i\sigma^y_{i+1}+
h^{-1}\sigma^z_{i+1})
\end{equation}
is self-dual under the exchange $J\leftrightarrow h$, provided that the
real constants $\tilde{h}^y$ and $\tilde{h}^x$ are kept fixed, as follows
from the isomorphism of Equation \eqref{iso_xy} extended as
\begin{equation}\label{automorphism_sdXY}
\sigma^y_1\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^y_1,\ \ \ \ \ \ \sigma^x_N\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_N.
\end{equation}
This is
a {\it hidden self-duality} that will not be apparent if we consider
only bond algebras independent of the couplings, and that will not hold
true unless \(J_x=J=1/J_y\).
In the special case with only two sites, \(N=2\), the inclusion
of \(\sigma^y_1\) in the list of generators suffices to compute
dual variables. {}From Equations \eqref{iso_xy} and \eqref{automorphism_sdXY},
we get
\begin{eqnarray}
&&\mu^y_1=\Phi_{{\sf d}}(\sigma^y_1)=\sigma^y_1 ,\nonumber\\
&&\mu^x_1=\Phi_{{\sf d}}(\sigma^x_1)=
(2Jh\sigma^x_1+(J^2-h^2)\sigma^z_1\sigma^x_2)/(J^2+h^2),\nonumber\\
&&\ \\
&&\mu^y_2=\Phi_{{\sf d}}(\sigma^y_2)=
(2(Jh)^{-1}\sigma^y_2+(J^{-2}-h^{-2})\sigma^y_1\sigma^z_2)/(J^{-2}+h^{-2}),\nonumber\\
&&\mu^x_2=\Phi_{{\sf d}}(\sigma^x_2)=\sigma^x_2.
\nonumber
\end{eqnarray}
These dual variables become identical to the original ones at the self-dual point,
as expected, since the self-duality map of Equation \eqref{iso_xy}
reduces to the identity map.
It is interesting to point out, in the light of this self-duality, an argument that
has been put forward
to show that non-Abelian self-dualities cannot possibly exist. In Reference
\cite{orland}, it is argued that self-dualities are unitary transformations
{\it that exchange the kinetic with the potential energy term} in a Hamiltonian,
and since these two terms must have different spectra for a non-Abelian theory,
a non-Abelian self-duality cannot exist. The emergent self-duality
of the XY model does not explicitly contradict this reasoning, but
suggests a way to escape its conclusion:
an emergent non-Abelian self-duality may appear as a property
of bonds that are {\it combinations of kinetic and
potential energy terms}, since such combinations can have matching spectra.
\subsection{Elimination of gauge symmetries by bond-algebraic dualities}
\label{sec3.12}
In this section we explain an extension of the notion of duality
established in Section \ref{sec3.8} that can accommodate changes in
the dimension of the state space, and show its use
to eliminate gauge symmetry constraints. In practice, however, we can potentially
eliminate any {\it local} (or in the language
of Reference \cite{tqo}, \(d=0\) gauge-like) symmetry
in this way, so it is important to keep in mind
that a local symmetry need not always be a gauge constraint that
can be disposed of. We term dualities that
eliminates gauge symmetries gauge-reducing dualities.
We start with a brief reminder of the distinction between ordinary (Wigner) and gauge
quantum symmetries, before discussing gauge-reducing dualities in detail.
In principle, the ideas that follow apply equally well to Abelian and non-Abelian
gauge theories,
but non-Abelian models present technical complications that put them at the frontier
of bond-algebraic studies, and thus beyond the scope
of this paper.
\subsubsection{Ordinary versus gauge symmetries}
Quantum symmetries are always embodied in one and the same mathematical
statement: they are unitary or anti-unitary mappings that commute with the Hamiltonian
(see the discussion at the beginning of Section \ref{sec3.9}).
But this is not to say that all symmetries have the same physical meaning,
nor the same mathematical consequences. There is a distinction
between ordinary symmetries, like rotations in space, and gauge symmetries.
Ordinary symmetries have direct physical impact, since they
can influence the
level degeneracy of a Hamiltonian (and with it, its thermal physics, see Appendix
\ref{sec3.1}) and constrain transition amplitudes to satisfy stringent selection
rules. On the other hand, gauge symmetries are {\it constraints} pointing to
a fundamental redundancy. The state space of a model
with gauge symmetries is {\it larger than physical}, meaning that it contains
states that cannot be prepared or observed by experimental means (or may even
contain states of negative norm). The
{\it sector of physical states} is precisely that sector that
is invariant under the action of all the gauge symmetries.
Similarly, Hermitian operators that do not
commute with the gauge symmetries are not observables, in the sense
that its eigenvalues do not represent measurable quantities (think, for
example, of the vector potential in QED). In a gauge
theory, an observable must be Hermitian, and commute with all the gauge
symmetries.
In perspective, gauge symmetries are better thought of as constraints,
and it may seem desirable to keep them conceptually far apart from
ordinary symmetries.
But it is unavoidable on first principles that quantum constraints
may look just like a symmetry. For suppose \(C\) is an operator representing
a quantum constraint. Then it must be that
\begin{equation}
\frac{dC}{dt}=0=i[H,C],\ \ \ \ (\mbox{consistent constrained dynamics}),
\end{equation}
to ensure that the dynamics generated by \(H\) is consistent with
the constraint. Then if \(C\) is (Hermitian)
unitary, it will look just like (the generator of) a symmetry.
In practice, physical input is required to set apart
constraints from symmetries. Take, for example,
the specification of the quantum statistics of identical particles.
Until Pauli proposed his exclusion principle, it would have
been natural to argue that the many-body Schr\"odinger equation
for indistinguishable particles had the group of permutations among its {\it symmetries}.
It took the introduction of a {\it new physical principle}
to show that these symmetries were in fact {\it constraints},
or superselection rules,
that select the fermionic or bosonic sector of Fock space as the
sector of physical states.
If a quantum gauge theory arises from the quantization of a {\it classical} gauge theory,
then there is no risk of confusing gauge and ordinary symmetries.
On the other hand, recent developments in condensed matter physics are
fostering the development of quantum models that do not show an obvious classical
limit, and that posses local symmetries that look much like gauge symmetries (see, for
example, Kitaev's honeycomb model, discussed at the end of Section \ref{sec3.6},
and Section \ref{subsec3.12.3}). Then, one is forced to face the problem of deciding
whether these should be treated as ordinary symmetries, or
as gauge symmetries (constraints), in part because the (self)-dualities available
will depend drastically on which one it is.
The problem could easily show up for effective theories
of strongly correlated systems, where emergent symmetries \cite{GJW} could
well be local.
\subsubsection{Gauge-reducing dualities}
In the light of the previous discussion, it would seem natural to assume that
a duality between a model with and a model without gauge symmetries
(a gauge-reducing duality) should
be emergent, in the sense of Section \ref{sec3.11}.
That is, it would seem that one should
project the bond algebra into the subspace of gauge invariant states first,
\(h_\Gamma \rightarrow P_{\sf GI}h_\Gamma P_{\sf GI}\), in order to figure out
the algebraic relations between physical (gauge-invariant) bonds, and then look
for a duality.
But as it turns out, the bond-algebraic approach does not require the
elimination of gauge symmetries to work, and that is why bond-algebraic
dualities are practical tools for removing gauge symmetries. If one chooses
the bond algebra of the gauge model wisely, one can find mappings that preserve all the
algebraic relations to models that do not
have any gauge symmetries.
To make these ideas more precise, let \(H_{\sf G}\) be the Hamiltonian for the gauge model,
with gauge symmetries \(G_\Gamma\), \([H_{\sf G},G_\Gamma]=0\), and let \(H_{\sf GR}\)
be the dual, completely gauge-reduced model. Then the gauge-reducing duality
maps
\begin{equation}\label{projective_unitary_I}
\Phi_{{\sf d}}(H_{\sf G})=H_{\sf GR},\ \ \ \ \ \ \mbox{and}\ \ \ \
\Phi_{{\sf d}}(G_\Gamma)=\mathbb{1},\ \ \forall \Gamma,
\end{equation}
thus rendering all the gauge symmetries trivial. Notice that \(\Phi_{{\sf d}}\) is not
an isomorphisms as for ordinary dualities, but rather an
{\it homomorphism} (homomorphisms preserve all the algebraic relations,
but need not be one-to-one). To be quantum-mechanically
meaningful, \(\Phi_{{\sf d}}\) must be implementable as an operator \(U_{{\sf d}}\)
(called a projective unitary)
that preserves the norm of gauge-invariant states, and projects other states out.
In formulas,
\begin{equation}
\Phi_{{\sf d}}(\mathcal{O})=U_{{\sf d}} \mathcal{O}U_{{\sf d}}^\dagger,
\end{equation}
with
\begin{equation}\label{projective_unitary_II}
U_{{\sf d}} U_{{\sf d}}^\dagger=\mathbb{1},\ \ \ \ \ \ U_{{\sf d}}^\dagger U_{{\sf d}}=P_{\sf GI},
\end{equation}
where \(P_{\sf GI}=P_{\sf GI}^2=P_{\sf GI}^\dagger\) is the orthogonal
projector onto the subspace of gauge invariant states \(|\psi\rangle\)
that satisfy
\begin{equation}
G_\Gamma|\psi\rangle=|\psi\rangle,\ \ \forall \Gamma.
\end{equation}
This completes the definition of a gauge-reducing duality.
Appendix \ref{appG}
describes a
concrete example of projective unitaries \(U_{{\sf d}}\).
Unlike ordinary unitaries, projective unitaries are represented by {\it rectangular
matrices}. In the particular case described in Equation
\eqref{projective_unitary_I}, \(U_{{\sf d}}\) has dimension \({\sf dim}\mathcal{H}_{\sf GR}
\times{\sf dim}\mathcal{H}_{\sf G}\), where \(\mathcal{H}_{\sf G}\)
and \(\mathcal{H}_{\sf GR}\) are the state spaces of the
gauge and gauge-reduced models, respectively.
Our bond-algebraic approach constitutes an excellent technique for detecting
gauge-reducing dualities, in part because
gauge symmetries are {\it local} in general.
This makes possible to choose a set of bond generators such that
each bond {\it individually} commutes with the gauge symmetries.
In this way we gain direct access to the physical (gauge-invariant)
algebra of interactions, and we can look for bond algebraic dualities to other
representations that show {\it no gauge symmetries}. We saw already
a duality along these lines in Section \ref{sec3.6}, when we studied
a new duality for the Abelian quantum Ising and non-Abelian Heisenberg models
in arbitrary dimension $d$. Let us study next
a simpler example, well-known in the literature \cite{fradkin_susskind,kogut},
from our new perspective.
The \(\mathbb{Z}_2\), \(d=2\) dimensional gauge model \cite{fradkin_susskind},
\begin{equation}\label{ising_gauge}
H_{\sf G}=\sum_{\bm{r}}\ (\sigma^x_{({\bm{r}},1)}+\sigma^x_{({\bm{r}},2)}+
\lambda\ B_{({\bm{r}},3)}),
\end{equation}
features spin $S=1/2$ degrees of freedom residing on links of a square
lattice, and plaquette operators $B_{({\bm{r}},3)}$ defined in Equation
\eqref{gauge_plaquette}.
Its group of gauge symmetries is generated by the unitaries
\begin{equation}
G_{\bm{r}}= \sigma^x_{({\bm{r}},1)}\sigma^x_{({\bm{r}},2)} \sigma^x_{({\bm{r}}-{\bm{e_1}},1)}
\sigma^x_{({\bm{r}}-{\bm{e_2}},2)},
\end{equation}
that not only commute with \(H_{\sf G}\), \([H_{\sf G},G_{\bm{r}}]=0\), but
commute with each one of the bonds
\begin{equation}
\sigma^x_{({\bm{r}},1)},\ \ \ \ \ \ \sigma^x_{({\bm{r}},2)}, \ \ \ \ \ \ B_{({\bm{r}},3)},
\end{equation}
individually. In other words, the bond algebra they generate is
gauge-invariant, and it is further characterized by simple relations:
(i) all the bonds square to the identity,
(ii) each spin \(\sigma^x\) anti-commutes with two adjacent plaquettes
\(B_{(.,3)}\), and
(iii) each plaquette \(B_{(.,3)}\) anti-commutes
with four spins \(\sigma^x\).
This set of relations is identical to the one found in the \(d=2\)
dimensional quantum Ising model of Equation \eqref{anyd_ising}, and
the mapping
\begin{eqnarray}
\sigma^x_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_{{\bm{r}}-{\bm{e_2}}}\sigma^z_{\bm{r}},\ \ \ \ \ \
\sigma^x_{({\bm{r}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_{{\bm{r}}-{\bm{e_1}}}\sigma^z_{\bm{r}},\ \ \ \ \ \
B_{({\bm{r}},3)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{\bm{r}},
\label{gauge_to_ising}
\end{eqnarray}
illustrated in Figure \ref{duality_gauge_ising}, shows
that the two are homomorphic. Thus \(\Phi_{{\sf d}}\)
maps $H_{\sf G}$ to $H_{\sf I}$, provided we identify the constants
$\lambda\leftrightarrow h$ and $1 \leftrightarrow J$.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.5\columnwidth]{./dgauge_ising.eps}
\end{center}
\caption{A joint (superimposed) representation of the \(d=2\) dimensional Ising and
$\mathbb{Z}_2$ (Ising) gauge models.
The bonds of the Ising model are indicated by heavy
bullets and rods; and the bonds of the $\mathbb{Z}_2$ gauge model by
crosses and dashed diamonds. The square lattices for the quantum Ising and
$\mathbb{Z}_2$ (Ising) gauge models are shown displaced relative to each
other (they are lattices dual to each other), to clarify the geometric aspects
of the duality homomorphism \(\Phi_{{\sf d}}\) of Equation \eqref{gauge_to_ising}.
}
\label{duality_gauge_ising}
\end{figure}
On the other hand, the mapping \eqref{gauge_to_ising}
is intriguing, because the Ising model does not have any gauge
symmetries. What happened to all the gauge symmetries? To answer this
question, we compute
\begin{eqnarray}
&&\Phi_{{\sf d}}(G_{\bm{r}})=\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)}\sigma^x_{({\bm{r}},2)}
\sigma^x_{({\bm{r}}-{\bm{e_1}},1)} \sigma^x_{({\bm{r}}-{\bm{e_2}},2)})=\nonumber\\
&&\sigma^z_{{\bm{r}}-{\bm{e_2}}}\sigma^z_{\bm{r}}\times\sigma^z_{{\bm{r}}-{\bm{e_1}}}\sigma^z_{\bm{r}}\times
\sigma^z_{{\bm{r}}-{\bm{e_2}}-{\bm{e_1}}}\sigma^z_{{\bm{r}}-{\bm{e_1}}}\times\sigma^z_{{\bm{r}}-{\bm{e_1}}-{\bm{e_2}}}\sigma^z_{{\bm{r}}-{\bm{e_2}}}=
\mathbb{1}.
\end{eqnarray}
Thus we see that \(\Phi_{{\sf d}}\) is a gauge-reducing duality homomorphism, and that
\(H_{\sf I}\) represents all the physics contained in \(H_{\sf G}\), but
without all the gauge redundancies.
Since a gauge-invariant bond algebra
captures the purely physical properties of the interactions, the
equivalence (duality) between the two theories is self-evident in our
formalism. The bond algebraic approach naturally includes gauge
invariant Wilson loops, Aharonov-Bohm phases, and the local fields
\(F^{\mu \nu}\) as gauge invariant quantities constructed out of the
individual bonds of the gauge theory.
Let us explain briefly how to construct
the projective unitary that implements the mapping of Equation \eqref{gauge_to_ising}.
Notice that the complete commuting set of observables
\(\sigma^x_{({\bm{r}},\nu)}\), \(\nu=1,2\), is mapped to the set
\(\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\). Thus we can write
the projective unitary \(U_{\sf d}\)
in terms of the basis \(|x\rangle\)
of simultaneous eigenstates of \(\sigma^x_{({\bm{r}},\nu)}\), and the spanning set
\(|z\rangle\) of simultaneous eigenstates of \(\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\),
\begin{eqnarray}
\sigma^x_{({\bm{r}},\nu)}|x\rangle&=&x_{({\bm{r}},\nu)}|x\rangle,
\ \ \ \ \ \ x_{({\bm{r}},\nu)}=\pm 1,\\
\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_\nu}}|z\rangle&=&z_{({\bm{r}},\nu)}|z\rangle,
\ \ \ \ \ \ z_{({\bm{r}},\nu)}=\pm 1 . \nonumber
\end{eqnarray}
The set \(\{|z\rangle\}\) is over-complete, because the same vector shows
up $m_z$ times with different labels, and it is convenient to renormalize
the \(|z\rangle\) so that
\begin{equation}
\langle z|z'\rangle=\frac{1}{m_z}\tilde{\delta}(z,z'),
\end{equation}
\(\tilde{\delta}(z,z')\) equals \(1\) if \(z\) and \(z'\) label the same
vector, and zero otherwise.
With these conventions in place, we can describe \(U_{\sf d}\) explicitly:
\begin{equation}
U_{\sf d}^\dagger=\sum_{x_{({\bm{r}},\bar{\nu})}=
z_{({\bm{r}}-\bm{e_{{\nu}}},{\nu})}} |x\rangle
\langle z|,
\end{equation}
where \(\bar{\nu}=2\) if \(\nu=1\), and \(\bar{\nu}=1\) if \(\nu=2\).
The condition \({x_{({\bm{r}},\bar{\nu})}=z_{({\bm{r}}-\bm{e_{{\nu}}},{\nu})}}\)
follows from the duality homomorphism that maps
\begin{equation}
\sigma^x_{({\bm{r}},\bar{\nu})}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_{{\bm{r}}-\bm{e_\nu}}\sigma^z_{\bm{r}}.
\end{equation}
In Appendix \ref{appG}, we describe this same construction for
finite renditions of the model that can be checked numerically.
\subsubsection{Ordinary versus gauge symmetries II: An example}
\label{subsec3.12.3}
The bond-algebraic elimination of gauge symmetries can proceed
just as easily for {\it any local symmetry}, but if the symmetry that gets
discarded is not gauge, the dual model is not a faithful representation
of the physics of the original model. This fact forces us
to reconsider critically the concept of a gauge symmetry. What sets an
ordinary local symmetry (that should not be eliminated)
apart from a gauge symmetry? Is there any {\it intrinsic} property of
\(H_{\sf G}\) and/or its symmetries that distinguish some
of them as gauge symmetries? Unfortunately, we do not know the answer to this
question, and, as we understand them know,
it seems that dualities cannot set apart gauge from ordinary
local symmetries. This is suggested by the example that we discuss next.
Consider a honeycomb lattice with spins \(S=1/2\) residing on its
vertices \({\bm{r}}\), and a dual triangular lattice with sites \({\bm{r}}^*\). As
shown in Figure \ref{hex_tr}, we can use the sites \({\bm{r}}^*\) to label the
elementary hexagons of the honeycomb lattice.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.6\columnwidth]{./latt_hex_and_tr.eps}
\end{center}
\caption{Honeycomb and triangular dual lattices with vertices ${\bm{r}}$ and
${\bm{r}}^*$, respectively. There are spins $S=1/2$ degrees of freedom placed on
the vertices of both lattices, that are labelled either by the site labels
\({\bm{r}}\) or ${\bm{r}}^*$, or by an integer \(i\) that indicates
their relative position within elementary hexagonal or triangular plaquettes.
Notice also that \({\bm{r}}\) (${\bm{r}}^*$) can be used to label these plaquettes.}
\label{hex_tr}
\end{figure}
With this convention and notation in place, we can introduce the
Hamiltonian
\begin{equation}
H_{\sf honeycomb}[h,J]= \sum_{\bm{r}}\ h\ \sigma^x_{\bm{r}}+\ \sum_{{\bm{r}}^*}\ J\
(\sigma^z_1\sigma^z_2\sigma^z_3 \sigma^z_4\sigma^z_5\sigma^z_6)_{{\bm{r}}^*}
\end{equation}
that feature plaquette interactions among the spins laying on every
single elementary hexagon.
It is ideal to illustrate the dilemma brought
about just now, because $H_{\sf honeycomb}[h,J]$ does commutes with a large
set of local unitaries
\begin{equation}\label{sym_honey}
S_{{\bm{r}}^*}=(\sigma^x_1\sigma^x_2\sigma^x_3\sigma^x_4\sigma^x_5\sigma^x_6)_{{\bm{r}}^*},
\ \ \ \ \ \ \ \ \ \ \ [S_{{\bm{r}}^*},\ H_{\sf honeycomb}]=0,
\end{equation}
but does not have an obvious classical limit. This face us with the
problem of deciding whether we should treat the \(S_{{\bm{r}}^*}\) as
gauge symmetries and try to eliminate them, or as ordinary symmetries.
In any case, there is a duality that maps all of the
operators \(S_{{\bm{r}}^*}\) to the identity. The dual model is most easily
described by placing spins \(S=1/2\) on the sites \({{\bm{r}}^*}\) of the
dual triangular lattice, as in Figure \ref{hex_tr}. Its Hamiltonian
then reads
\begin{eqnarray}
H_{\sf tr}[J,h]= \sum_{{\bm{r}}^*}\ J\ \sigma^x_{{\bm{r}}^*}+\ \sum_{{\bm{r}}}\ h\
(\sigma^z_1\sigma^z_2\sigma^z_3)_{{\bm{r}}},
\end{eqnarray}
and features plaquette interactions among the spins laying on every
single elementary triangle. It is straightforward to check that the
mapping
\begin{equation}
\sigma^x_{\bm{r}} \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ (\sigma^z_1\sigma^z_2\sigma^z_3)_{{\bm{r}}}\ ,
\ \ \ \ \ \ \ \
(\sigma^z_1\sigma^z_2\sigma^z_3 \sigma^z_4\sigma^z_5\sigma^z_6)_{{\bm{r}}^*}
\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{{\bm{r}}^*}
\end{equation}
is a homomorphism of bond algebras such that \(\Phi_{{\sf d}}(H_{\sf
honeycomb}[h,J])=H_{\sf tr}[J,h]\), {\it but \(H_{\sf tr}\) has no local
symmetries}. In particular,
\begin{equation}
\Phi_{{\sf d}}(S_{{\bm{r}}^*})=\prod_{i=1}^6\ \Phi_{{\sf d}}(\sigma^x_i)=\prod_{i=1}^6
\ (\sigma^z_1\sigma^z_2\sigma^z_3)_i=\mathbb{1},
\end{equation}
since \(\Phi_{{\sf d}}\) maps the six spins \(\sigma^x\) on the vertices of an
hexagon to the six plaquette terms \(\sigma^z_1\sigma^z_2\sigma^z_3\)
that share the center point of that hexagon, a vertex of the dual
lattice.
Whether this duality is physically meaningful or not rests on deciding
whether the symmetries of Equation \eqref{sym_honey} should be
discarded or not.
\subsubsection{Systematic determination of gauge-reduced dual models}
\label{sec3.12.4}
There is a systematic way to construct completely gauge-reduced duals
of gauge models that has the unpleasant feature of
requiring the introduction of non-local bonds in the dual model.
While there are intuitive
physical arguments to justify this (namely, Gauss' law permits to
measure the charge of a particle
by measuring its electric field on the surface of
a sphere {\it arbitrarily far away from it}), the fact is that sometimes
completely gauge-reducing duals with local bonds
are {\it readily available} (like in the example of the previous section),
and there is no need to exploit the systematic construction
(see Sections \ref{sec5.3}, \ref{sec5.4}, \ref{sec5.5}, and \ref{sec6.4}). But sometimes
the systematic construction seems to be the best we can do (see Section
\ref{sec6.3}).
In this section, we describe the systematic construction for the \(\mathbb{Z}_2\)
gauge model of Equation \eqref{ising_gauge}, to illustrate the ideas on which
it rests (these ideas are a reinterpretation and generalization
in terms of bond algebras of an scheme introduced in
Reference \cite{fradkin_susskind} to study the
\(d=3\), \(\mathbb{Z}_2\) gauge theory). The generalization to higher dimensions,
other (Abelian) gauge groups, and matter-coupled gauge theories is straightforward,
and will be illustrated in other parts of this paper (see for example
Section \ref{sec6.4}).
Given $H_{\sf G}$, we are looking for a model \(H_{\sf GR}\),
and a mapping \(\Phi_{{\sf d}}\), \(\Phi_{{\sf d}}(H_{\sf G})=H_{\sf GR}\), that satisfy some
stringent conditions. The bonds of the dual model \(H_{\sf GR}\) are going to be
the operators
\begin{equation}
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)}),\ \ \ \ \ \Phi_{{\sf d}}(\sigma^x_{({\bm{r}},2)}),
\ \ \ \ \ \Phi_{{\sf d}}(B_{({\bm{r}},3)}),
\end{equation}
that satisfy
\begin{equation}\label{byebye_gauge}
\Phi_{{\sf d}}(G_{\bm{r}})=\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)})\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},2)})
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-{\bm{e_1}},1)})\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-{\bm{e_2}},2)})=\mathbb{1},
\end{equation}
and our job is to determine them (the splitting \( \Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)}\sigma^x_{({\bm{r}},2)}
\sigma^x_{({\bm{r}}-{\bm{e_1}},1)}\sigma^x_{({\bm{r}}-{\bm{e_2}},2)})=\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)})\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},2)})
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-{\bm{e_1}},1)})\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-{\bm{e_2}},2)})\) is
correct because the \(\sigma^x_{({\bm{r}},\nu)}\) are gauge-invariant operator, see
next section, Section \ref{no_dual_vars}).
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.65\columnwidth]{./2dstring.eps}
\end{center}
\caption{The gauge-reduced dual of the \({\mathbb{Z}}_2\) gauge theory constructed
in this section features spins \(S=1/2\) on the {\it horizontal} links of
a square lattice (denoted by crosses on those links),
but {\it not} on the vertical links. Also it features a non-local bond,
the string operator \(s_{\bm{r}}\) of Equation \eqref{bye_gauge_string}, constructed
as an infinite product of all the spins \(\sigma^x_{({\bm{r}}',1)}\) that lie
below the link \(({\bm{r}},2)\), and to the immediate left and right of the straight line
containing it (some of them are shown in red).
In spite of being non-local, \(s_{\bm{r}}\) commutes with most other bonds. For example,
it commutes with \(\sigma^z_{({\bm{r}},1)}\sigma^z_{({\bm{r}}-{\bm{e_2}},1)}\), shown in green
in the figure.}
\label{string2dop}
\end{figure}
Now, since \(\Phi_{{\sf d}}(\sigma^{x\ \ 2}_{({\bm{r}},\nu)})=\Phi_{{\sf d}}(\sigma^{x}_{({\bm{r}},\nu)})^2=\mathbb{1}\),
Equation \eqref{byebye_gauge} shows that \(\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},2)})\) satisfies
a recurrence relation,
\begin{equation}
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},2)})=\left(\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-{\bm{e_1}},1)})
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)})\right)\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-{\bm{e_2}},2)}),
\end{equation}
so that
\begin{equation}\label{bye_gauge_string}
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},2)})=\prod_{m=0}^\infty\ \left(\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-m{\bm{e_2}}-{\bm{e_1}},1)})
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}}-m{\bm{e_2}},1)})\right)\equiv s_{({\bm{r}},2)}.
\end{equation}
The form of the string operator \(s_{({\bm{r}},2)}\)
suggests that we should look for a representation of
\(\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)})\) and \(\Phi_{{\sf d}}(B_{({\bm{r}},3)})\) in terms
of \(\sigma^x_{({\bm{r}},1)}\) and \(\sigma^z_{({\bm{r}},1)}\) {\it alone},
because then the dual model will feature only {\it half as many}
degrees of freedom (or equivalently, will act on a state space
exponentially smaller). It follows that
\begin{equation}\label{bye_gaugeII}
\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},1)})=\sigma^x_{({\bm{r}},1)},\ \ \ \ \ \
\Phi_{{\sf d}}(B_{({\bm{r}},3)})=\sigma^z_{({\bm{r}},1)}\sigma^z_{({\bm{r}}+{\bm{e_2}},1)},
\end{equation}
and the dual, completely gauge-reduced model reads
\begin{equation}\label{systematic_gr}
H_{\sf GR}=\sum_{\bm{r}}\ (\sigma^x_{({\bm{r}},1)}+ s_{({\bm{r}},2)}+
\lambda\ \sigma^z_{({\bm{r}},1)}\sigma^z_{({\bm{r}}+{\bm{e_2}},1)})=H_0+\sum_{\bm{r}}\ s_{({\bm{r}},2)}.
\end{equation}
It is interesting to notice that \(H_0\) describes a bundle of
horizontal quantum Ising chains that do not interact with each other.
The interactions between chains are carried exclusively
by the term \(\sum_{\bm{r}}\ s_{({\bm{r}},2)}\) that contains all the
non-local bonds.
The traditional approach to the duality between the \(\mathbb{Z}_2\)
gauge model and the Ising model relies more or less implicitly on the
construction we have just described \cite{kogut}, because the
Hamiltonian of Equation \eqref{systematic_gr} can be mapped to
\(H_{\sf I}\) in terms of dual variables, while \(H_{\sf G}\) cannot
(this will be explained in detail in the next section). Let us show how
this works. The duality (which is now an ordinary, unitarily implementable duality)
between \( H_{\sf GR}\) and \(H_{\sf I}\) is established by the Isomorphism
\begin{equation}
\label{em_ising_isoIII}
\sigma^x_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_{{\bm{r}}-{\bm{e_2}}}\sigma^z_{{\bm{r}}},\ \ \ \ \ \
\sigma^z_{({\bm{r}},1)}\sigma^z_{({\bm{r}}+{\bm{e_2}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{\bm{r}}
\end{equation}
Since \(s_{({\bm{r}},2)}\) is a function of \(\sigma^x_{({\bm{r}},1)}\), the relation
\begin{equation}
s_{({\bm{r}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_{{\bm{r}}-{\bm{e_1}}}\sigma^z_{{\bm{r}}}
\end{equation}
follows from the ones just listed. The dual variables that follow
from the isomorphism of Equation \eqref{em_ising_isoIII} are
\begin{equation}
\mu^x_{({\bm{r}},1)}=\sigma^z_{{\bm{r}}-{\bm{e_2}}}\sigma^z_{{\bm{r}}},\ \ \ \ \ \
\mu^z_{({\bm{r}},1)}=\prod_{m=0}^\infty\ \sigma^x_{{\bm{r}}+m{\bm{e_2}}},
\end{equation}
that clearly satisfy the correct Pauli algebra, and are essentially the
same as the dual variable used in Reference \cite{kogut}.
\subsubsection{Dual variables in gauge-reducing dualities}\label{no_dual_vars}
The bond-algebraic approach to dualities
of gauge models is remarkably simpler than the traditional approach
\cite{fradkin_susskind,kogut},
and it is also unrelated, because the homomorphism \(\Phi_{{\sf d}}\) cannot
be extended to define dual variables. To see this, suppose to the
contrary that we can extend the action
of the gauge-reducing homomorphism
\(\Phi_{{\sf d}}\) defined in Equation \eqref{gauge_to_ising} to include
the spins \(\sigma^z_{({\bm{r}},\nu)}\), \(\nu=1,2\), so that
\begin{equation}
\sigma^z_{({\bm{r}},\nu)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \Phi_{{\sf d}}(\sigma^z_{({\bm{r}},\nu)})
\end{equation}
becomes a meaningful operation. Then, on one hand, this would
imply that
\begin{equation}
\Phi_{{\sf d}}([\sigma^z_{({\bm{r}},\nu)},G_{\bm{r}}])=[\Phi_{{\sf d}}(\sigma^z_{({\bm{r}},\nu)}),\mathbb{1}]=0.
\end{equation}
But, on the other hand,
\begin{equation}
[\sigma^z_{({\bm{r}},\nu)},G_{\bm{r}}]=-2\sigma^z_{({\bm{r}},\nu),{\bm{r}}}G_{\bm{r}}
\end{equation}
which, together with the previous equation, implies that
\begin{equation}
\Phi_{{\sf d}}(\sigma^z_{({\bm{r}},\nu)})=0.
\end{equation}
Since this is in contradiction with the fact that \(\Phi_{{\sf d}}(B_{({\bm{r}},3)})\neq 0\),
we see that we cannot extend the action of \(\Phi_{{\sf d}}\)
to the spins \(\sigma^z_{({\bm{r}},\nu)}\) in a consistent way. In other words,
\(\Phi_{{\sf d}}\) cannot be used to define dual spins \(\mu^z\).
This conclusion
seems paradoxical because \(\Phi_{{\sf d}}\) must be projectively unitarily implementable. Then,
it would seem natural to define
\begin{equation}\label{dual_zz}
\Phi_{{\sf d}}(\sigma^z_{({\bm{r}},\nu)})\equiv U_{{\sf d}}\sigma^z_{({\bm{r}},\nu)}U_{{\sf d}}^\dagger
\stackrel{?}{=}\mu^z_{({\bm{r}},\nu)}.
\end{equation}
Notice, however, that this extension of \(\Phi_{{\sf d}}\) to operators
that are not gauge-invariant is {\it not multiplicative}.
In other words, the relation
\begin{equation}\label{multiplicative_on_gi}
U_{{\sf d}}\mathcal{O}\mathcal{O}'U_{{\sf d}}^\dagger =
U_{{\sf d}}\mathcal{O}U_{{\sf d}}^\dagger\ U_{{\sf d}}\mathcal{O}'U_{{\sf d}}^\dagger\ \ \ \ \ \
\ \ \ (\mathcal{O},\ \mathcal{O}'\ \mbox{gauge invariant}),
\end{equation}
only holds true if \(\mathcal{O}\) and \(\mathcal{O}'\) commute
with all the gauge symmetries. This means that
we should not expect the operator of Equation \eqref{dual_zz} to satisfy
the correct anti-commutation relations with \(\Phi_{{\sf d}}(\sigma^x_{({\bm{r}},\nu)})\),
and also explains the paradoxical conclusion of the last paragraph
(\(\Phi_{{\sf d}}(\sigma^z_{({\bm{r}},\nu)})=0\)),
that was derived precisely on the assumption that Equation
\eqref{multiplicative_on_gi} holds true {\it in general}.
The fact that Equation \eqref{multiplicative_on_gi} only works for
gauge-invariant operators follows from the general
relations of Equations \eqref{projective_unitary_I} and
\eqref{projective_unitary_II}. First notice that
\begin{eqnarray}
P_{\sf GI}(\mathbb{1}-P_{\sf GI})&=&U_{{\sf d}}^\dagger U_{{\sf d}}(\mathbb{1}-P_{\sf GI})=0\
\ \ \rightarrow \ \ \ U_{{\sf d}}(\mathbb{1}-P_{\sf GI})=0,\\
(\mathbb{1}-P_{\sf GI})P_{\sf GI}&=&(\mathbb{1}-P_{\sf GI})U_{{\sf d}}^\dagger U_{{\sf d}}=0
\ \ \ \rightarrow \ \ \ (\mathbb{1}-P_{\sf GI})U_{{\sf d}}^\dagger=0
\end{eqnarray}
(since \(U_{{\sf d}} U_{{\sf d}}^\dagger=\mathbb{1}\)). Then, it follows from
the decomposition
\begin{equation}
\mathcal{O}=P_{\sf GI}\mathcal{O}P_{\sf GI}+P_{\sf GI}\mathcal{O}
(\mathbb{1}-P_{\sf GI})+
(\mathbb{1}-P_{\sf GI})\mathcal{O}P_{\sf GI}+(\mathbb{1}-P_{\sf GI})
\mathcal{O}(\mathbb{1}-P_{\sf GI}) ,
\end{equation}
that
\begin{equation}
U_{{\sf d}}\mathcal{O}\mathcal{O}'U_{{\sf d}}^\dagger=
U_{{\sf d}} P_{\sf GI}\mathcal{O}\mathcal{O}'P_{\sf GI}U_{{\sf d}}^\dagger.
\end{equation}
But, if and only if, \(\mathcal{O}\) and \(\mathcal{O}'\) commute
with all the gauge symmetries, then they also commute with \(P_{\sf GI}\)
as well, so that we can further write
\begin{equation}
U_{{\sf d}} P_{\sf GI}\mathcal{O}\mathcal{O}'P_{\sf GI}U_{{\sf d}}^\dagger=
U_{{\sf d}}\mathcal{O}P_{\sf GI}\mathcal{O}'U_{{\sf d}}^\dagger=
U_{{\sf d}}\mathcal{O}U_{{\sf d}}^\dagger U_{{\sf d}}\mathcal{O}'U_{{\sf d}}^\dagger.
\end{equation}
In summary, \(\Phi_{{\sf d}}(\mathcal{O})=U_{{\sf d}}\mathcal{O}U_{{\sf d}}^\dagger\) is defined
on any operator, {\it but it acts as an algebra
homomorphism only on gauge-invariant operators}.
\subsection{Unifying classical and quantum dualities}
\label{classical&quantum}
This section explains one of the most important new results
in this paper: a bond-algebraic approach to dualities in classical
statistical mechanics \cite{con} (classical dualities described in
Section \ref{sec2.1} and Appendix \ref{appA}).
To our knowledge, this section's results
have been completely overlooked in the literature,
possibly because the key fact that dualities are unitary or projective unitary
mappings \cite{con}, has been unclear or overlooked up to now as well.
Some advantages of the bond-algebraic over
the traditional approach of Appendix \ref{appA} to classical dualities
will be discussed in detail
in Section \ref{sec8}, but we present
here for illustration a toy example, a {\it new self-duality}
for the \(D=1=d+1\) classical Ising chain in an external magnetic field.
Elementary as it is, this duality is noteworthy because the traditional
approach to classical dualities described in Appendix \ref{appA}
{\it fails in the presence of a minimally coupled external field}, and could not
have been used to derive it.
The bond-algebraic
approach to classical dualities is only rigorously applicable to models that admit a transfer
matrix formulation. While this covers a very wide range of interesting models,
of finite of infinite size extent,
there are dualities for models outside this category, most famously, the
duality of the solid-on-solid (SoS) model to a coulomb gas \cite{bookEPTCP}.
We explain these dualities in Appendix \ref{poisson}.
We associate bond algebras to partition functions of classical models through
the transfer matrix. The transfer matrix formalism
\cite{onsager,baxter} permits to recast partition functions $\cal Z$ as traces of
linear operators, and can come in several different flavors. For example,
row-to-row transfer matrices permit to write
\begin{equation}\label{rowtorow}
\mathcal{Z}={{\sf Tr}\ } (T_1\cdots T_s)^N ,
\end{equation}
where \(N\) is an integer related to the number of sites
in one of the lattice directions, that
one can think of as the {\it Euclidean time direction}, and
\(T_1, \cdots, T_s\) contain information about the directions transverse to the
time direction, and how constant-time sections of the lattice
are connected from one time to the next.
Written as in Equation \eqref{rowtorow}, \(\mathcal{Z}\) must satisfy periodic boundary
conditions in the time direction. In contrast, corner transfer matrices \cite{baxter}
permit to write
\begin{equation}\label{cmt}
\mathcal{Z}={{\sf Tr}\ } C_1\cdots C_s ,
\end{equation}
where \(s\) is fixed, the size of the lattice is completely encoded in
the \(C_i\), and the BCs are not fixed by the structure
of Equation \eqref{cmt}.
This paper will only consider row-to-row transfer matrices,
so from now on we focus on Equation \eqref{rowtorow},
keeping in mind though that it is possible to extend our formalism
to other types of transfer matrices. The general arguments of Section
\ref{sec3.2} concerning the additive bond structure of physical Hamiltonians
can be repeated {\it verbatim} for transfer matrices, with the only difference
that transfer matrices display a {\it multiplicative} rather than
additive bond structure:
\begin{equation}
T_i=\prod_\Gamma\ T_{i\Gamma}
\end{equation}
(\(\Gamma\) stands for a general index). The
\(\{T_{i\Gamma}\}_{\Gamma, i=1,\cdots,s}\) are now the bonds of interest,
and the definition of bond algebra proceeds as before, see Section \ref{sec3.2}.
Suppose next that you have an isomorphic representation of the bond algebra
\(\mathcal{A}\{T_{i\Gamma}\}_{\Gamma, i=1,\cdots,s}\), generated by a set
of dual bonds \(\{T_{i\Gamma}^D\}_{\Gamma, i=1,\cdots,s}\). The dual
transfer matrices \(T_i^D={\cal U}_{\sf d } T_i {\cal U}^\dagger_{\sf d }=
\prod_\Gamma\ T^D_{i\Gamma}\) will define,
through Equation \eqref{rowtorow}, a partition function \(\mathcal{Z}^D\) that may
look very different from \(\mathcal{Z}\). However,
\begin{equation}
\mathcal{Z}^D={{\sf Tr}\ } (T_1^D\cdots T_s^D)^N=
{{\sf Tr}\ } (\mathcal{U}_{{\sf d}} T_1\cdots T_s\mathcal{U}^\dagger_{{\sf d}})^N=\mathcal{Z}.
\end{equation}
We call this relation between partition functions, obtained in this way,
a classical bond-algebraic duality. Much as with quantum bond-algebraic dualities,
we would like to show that:
\begin{itemize}
\item{classical bond-algebraic dualities include traditional classical dualities
(at least for models that admit a transfer matrix) \cite{con}, and that}
\item{classical bond-algebraic dualities are useful.}
\end{itemize}
Both points will be discussed at length in Section \ref{sec8} where
we derive classical dualities, old and new, by bond-algebraic methods.
We would like, however, to advance here some support for the second point,
by illustrating our ideas with a new self-duality {\it that is elementary, and
yet cannot be derived by traditional (Fourier-based) means} (Appendix \ref{appA}).
The partition function of the periodic Ising chain of length $N$
($\sigma_{i+N}\equiv\sigma_i$) is
\begin{equation}\label{classical_ichain}
\mathcal{Z}_{\sf I}(K,\tilde{h})=\sum_{\{\sigma_i\}}\ \exp \left [ \sum_{i=1}^{N}
(K\sigma_{i}\sigma_{i+1}+\tilde{h}\sigma_i)\right ]=
{{\sf Tr}\ } (T_1T_2)^N,
\end{equation}
where
\begin{equation}
\label{tsfl}
T_1=e^K+e^{-K}\ \sigma^x,\ \ \ \ \ \
T_2=e^{\tilde{h}\ \sigma^z}=\cosh(\tilde{h})+\sinh(\tilde{h})\ \sigma^z,
\end{equation}
and we assume that \(K,\tilde{h}\geq0\). In this $D=1$ case, the
row-to-row transfer matrix connects rows that feature one Ising degree of freedom only.
The bond algebra of \(T_1\) and \(T_2\) is symmetric under the exchange
\(\sigma^x\leftrightarrow\sigma^z\) (this is just a rotation in spin space).
It follows that
\begin{eqnarray}\label{dualt0dising}
T_1^D=e^K+e^{-K}\ \sigma^z=A\ e^{\tilde{h}^*\sigma^z},
\ \ \ \ \ \
T_2^D=e^{\tilde{h}\ \sigma^x}=B(e^{K^*}+ e^{-K^*}\ \sigma^x),
\end{eqnarray}
provided the dual couplings \(\tilde{h}^*, K^*\) satisfy
\begin{equation}\label{dc_icf}
\sinh(2K)\sinh(2\tilde{h}^*)=1,\ \ \ \ \ \
\sinh(2K^*)\sinh(2\tilde{h})=1,
\end{equation}
and
\begin{equation}
A^2=1/(2\sinh(2\tilde{h}^*)),\ \ \ \ \ \ B^2=2\sinh(2\tilde{h}).
\end{equation}
Then, Equations \eqref{tsfl} and \eqref{dualt0dising} put together
define the bond-algebraic classical duality
\begin{equation}\label{ZKTH}
\frac{\mathcal{Z}_{\sf I}(K,\tilde{h})}{(2\sinh(2\tilde{h}))^{N/2}}
=\frac{{{\sf Tr}\ }(T_1T_2)^N}{(2\sinh(2\tilde{h}))^{N/2}}=
\frac{{{\sf Tr}\ }(T_2^DT_1^D)^N}{(2\sinh(2\tilde{h}))^{N/2}}=
\frac{\mathcal{Z}_{\sf I}(K^*,\tilde{h}^*)}{(2\sinh(2\tilde{h}^*))^{N/2}},
\end{equation}
where we have used the cyclic property of the trace.
Notice how {\it linear} bond algebraic operations
(the exchange \(\sigma^x\leftrightarrow\sigma^z\)) produce
highly {\it non-linear} relations between
classical couplings.
The classical self-dual line determined by \(\tilde{h}^*=\tilde{h}\) and \(K^*=K\)
is characterized by
\begin{equation}
\sinh(2K)\sinh(2\tilde{h})=1.
\end{equation}
This self-dual line is only {\it critical} when $ \tilde{h}=0$ and
$K \rightarrow \infty$, i.e., at zero temperature.
To our knowledge, the classical self-duality embodied in Equations \eqref{ZKTH}
and \eqref{dc_icf}, valid for any $N$, has not been written before in the literature
despite its simplicity.
It has, however, one remarkable feature: it is a duality for a model {\it
in a minimally-coupled external field}, and dualities for such models are
beyond the traditional approach as described in Appendix \ref{appA}, and
References \cite{wegner,wu_wang,savit,malyshev}.
The reason is that the standard approach
relies on the Fourier transform technique for establishing dualities, {\it and},
for these models with two-body interactions,
on having individual Boltzmann weights that define {\it circulant matrices}
\cite{aldrovandi}. But the Boltzmann weights in \(\mathcal{Z}_{\sf I}(K,\tilde{h})\)
cannot be chosen to be circulant except if \(\tilde{h}=0\). Therefore, the self-duality
of Equation \eqref{ZKTH} is not attainable by the Fourier transform method.
We think that this indicates that our bond-algebraic approach to classical
dualities may push its scope beyond the standard paradigm of
Fourier transforms, perhaps even to include classical non-Abelian dualities,
though this is a matter under study.
There is a different albeit related way to connect bond-algebraic quantum
dualities to classical dualities. It exploits the well-known relation between
partition functions of classical problems in $D=d+1$ dimensions and
quantum Hamiltonian problems in $d$ dimensions.
Quantum mechanical problems in Euclidean time (or equivalently, at finite temperature)
can be mapped to a classical partition function problem by use of
Feynman's path integral
for the case of quantum particles \cite{schulman} and fields \cite{rivers},
or by use of the closely related Suzuki-Trotter-Lie (STL) decomposition for
quantum lattice models \cite{suzuki,bookEPTCP}.
This {\it quantum-classical} mapping takes the general form \cite{bookEPTCP}
\begin{equation}\label{quantum_to_classical}
\mathcal{Z}(K)={{\sf Tr}\ } e^{-H[\lambda]} ,
\end{equation}
where \(\mathcal{Z}(K)\) stands for the path integral/partition function, and the
classical, $K$, and quantum, $\lambda$, couplings typically
connected by non-linear functional relations.
In general, this quantum-classical mapping
takes quantum problems $H$ in \(d\) dimensions into classical problems
$\mathcal{Z}$ in
\(D=d+1\) dimensions where the extra dimension, because of the
construction, attains periodic BCs \cite{bookEPTCP}.
Now we can translate duality properties of \(H\) into properties of \(\mathcal{Z}(K)\).
Suppose that
the Hamiltonian \(H_1[\lambda]\) is dual to another Hamiltonian
\(H_2[\lambda^*]\) as in Equations \eqref{H12H2},
$H_2[\lambda^*]=\mathcal{U}_{{\sf d}} H_1[\lambda]\mathcal{U}_{{\sf d}}^\dagger$.
Then, with the aid of the identity \({{\sf Tr}\ }(AB)={{\sf Tr}\ }(BA)\),
\begin{equation}
{{\sf Tr}\ } e^{-H_1[\lambda]}={{\sf Tr}\ } \!(\mathcal{U}_{{\sf d}}^\dagger e^{-H_2[\lambda^*]}\mathcal{U}_{{\sf d}})
= {{\sf Tr}\ } e^{-H_2[\lambda^*]},
\end{equation}
which implies that
\begin{equation}\label{classd_from_q}
\mathcal{Z}_1(K)=A(K,K^*) \mathcal{Z}_2(K^*),
\end{equation}
with an analytic proportionality factor $A(K,K^*)$ that is model specific.
Thus, {\it every quantum duality translates into a classical duality} \cite{con}.
Notice that it is crucial that $\mathcal{U}_{{\sf d}}$ is
a unitary or projective unitary, a fact that was overlooked in the past \cite{con}.
Due to the results of Section \ref{sec3.7}, the duality
relation \eqref{classd_from_q} can be made exact for {\it any} finite
size system $N$, and {\it not} just for the thermodynamic limit
$N \rightarrow \infty$.
Rigorously speaking this is always the case, with quantum and classical
dualities representing two sides of the same coin. On the other hand,
there is a useful
practical algorithm, that will be presented in Section \ref{sec8}, that makes
use of the STL decomposition and where the
thermodynamic limit is, in principle, required \cite{bookEPTCP}.
Typically, the use of the STL decomposition implies that the derived
classical model presents infinitesimally weak and infinitely-strong couplings,
and so the classical dualities derived in this way are less clearly established.
It is also important to keep in mind that the classical model \(\mathcal{Z}\) of Equation
\eqref{quantum_to_classical} depends on the quantum couplings \(\lambda\) not just
through \(K\), in the sense that different quantum couplings
can relate to qualitatively very different classical models (and not just
the same \(\mathcal{Z}\) at different values of \(K\)). For example,
the STL decomposition maps
a single quantum spin \(H=h_x\sigma^x+h_z\sigma^z\) to the model
of Equation \eqref{classical_ichain} (in the limit \(N\rightarrow\infty\),
{\it provided \(h_x<0\)}), and not otherwise. Similar constraints apply
elsewhere. As dualities relate different regions in the coupling spaces of
quantum problems, it is essential to keep track of exactly which classical
problems may correspond to a given quantum system for each set of
couplings.
\section{Quantum self-dualities by example: Lattice models}
\label{sec4}
\subsection{Self-dualities in the Potts, vector Potts, and
\(\mathbb{Z}_p\) clock models}
\label{sec4.1}
In this section, we will study two self-dual generalizations of the
\(d=1\) dimensional quantum Ising chain: the quantum Potts (P) and {\it vector}
Potts (VP) models (also known as \(p\)-clock model \cite{bookEPTCP,potts}).
These models are just two special examples of
a large class of \(\mathbb{Z}_p\) clock models \cite{cardy}, that we discuss
too, but very briefly. The fact that the family of \(\mathbb{Z}_p\) clock models
contains so many self-dual members is remarkable because \(\mathbb{Z}_p\) clock models
have {\it non-Abelian} symmetries {\it for \(p\geq3\)}. In other words, this
section discusses important examples of, inappropriately called, {\it non-Abelian self-dualities}
(see Section \ref{sec3.6} for a critical discussion of the notion
of non-Abelian duality). To the best of
our knowledge, the fact that these models display non-Abelian symmetries has been
overlooked up to now.
Potts models feature spins confined to a plain that can exist in any one of
$p\geq 2$ different states, and if $p=2$, they reduce to the Ising model.
We know already that in that case, the unitary implementing the self-duality squares
to one. But if $p\geq 3$ the self-duality is the square root of a {\it non-trivial}
discrete symmetry. This illustrates some ideas discussed
in Section \ref{sec3.9}.
We will first discuss the VP model in detail, since then the
self-dual properties of the P
and general \(\mathbb{Z}_p\) clock models will easily follow.
\subsubsection{The vector Potts model} \label{sec4.1.1}
The VP model is a popular test ground for {\it exotic} critical
behavior, such as phase transitions without long-range order
\cite{taroni,ortiz}. The configurations of
the classical, $D=2$ VP model are specified by a set of discretized angles
$\theta_{{\bm{r}}}$
\begin{equation}
\theta_{\bm{r}}=\theta_{r^1,r^2}=\frac{2\pi s_{\bm{r}}}{p},\ \ \ \ \ \ \ \ \ \
s_{\bm{r}}=0,1,\cdots,p-1,
\end{equation}
situated at the sites ${\bm r}=(r^1,r^2)$ of a square
lattice, and its partition function reads
\begin{equation}\label{classicalVP}
\mathcal{Z}_{\sf VP}=\sum_{\{\theta_{\bm{r}}\}}\ \exp\left[\sum_{{\bm{r}}}\sum_{\mu=1,2}\
K_\mu\cos(\theta_{{\bm{r}}+\bm{e_\mu}}-\theta_{\bm{r}})\right].
\end{equation}
The statistical mechanics of this model has been the subject of
research for many years, but there are still open problems. For
example, the VP model is known to exhibit a Kosterlitz-Thouless (KT)
phase transition
\cite{kosterlitz_thouless}, {\it for sufficiently large $p$}
\cite{frohlich}, yet to determine the smallest $p$ required for a
KT transition is a difficult problem, currently under debate \cite{ortiz}.
Reference \cite{ortiz} discusses the phase diagram, the nature of the
transitions and topological excitations of the VP model, and unveils the
U(1) symmetry, that emerges for $p\geq 5$, associated
with the appearance of discrete vortices and KT transitions.
In preparation for
writing the \(d=1\) quantum model corresponding to the classical
VP model by the quantum-classical mapping \cite{bookEPTCP},
we introduce some basic facts about the Weyl group algebra
\cite{schwinger}.
Its generators $U$ and
$V$ are operators characterized by the relations
\begin{equation}\label{weyl_group_algebra}
VU=\omega UV,\ \ \ \ \ \ V^p=1=U^p,
\end{equation}
where $ \omega=e^{2\pi i/p}$ is a $p$th root of unity. Equations
(\ref{weyl_group_algebra}) completely determine the irreducible, finite
dimensional, representations of $U$ and $V$. A $(p\times p)$ matrix
representation is given by
\begin{eqnarray}
{V}=
\begin{pmatrix}
0& 1& 0& \cdots& 0\\
0& 0& 1& \cdots& 0\\
\vdots& \vdots& \vdots& & \vdots \\
0& 0& 0& \cdots& 1\\
1& 0& 0& \cdots& 0\\
\end{pmatrix} \ , \mbox{ and } {U}={\sf
diag}(1,\omega,\omega^2,\cdots,\omega^{p-1}).
\label{VsandUs}
\end{eqnarray}
$V$ is sometimes called the fundamental circulant (unitary) matrix because
it generates the (commutative) algebra of circulant matrices
(meaning that any circulant matrix \(C\) is of the form \(C=\sum_{j=0}^{p-1}a_jV^j,\
a_i\in\mathbb{C}\) \cite{aldrovandi}). \(U\) and \(V\) together generate the full algebra of
$(p\times p)$ complex matrices, that we continue to call the Weyl group algebra,
to emphasize that we are working with a distinguished set of generators.
The Weyl group algebra admits a unitary automorphism $\Phi$ in the form
of a discrete Fourier transform \(F\) \cite{schwinger} that,
essentially, interchanges the two types of operators ($U,V$ and their
Hermitian conjugates) of the Weyl algebra. A direct calculation reveals that
the unitary and symmetric Fourier matrix \(F_{m
n}^\dagger=\omega^{mn}/\sqrt{p}\) with $m,n=0,1,\cdots,p-1$ maps
\begin{equation}\label{wga_aut}
\Phi(U)=V^{\dagger}=FUF^\dagger,\ \ \ \ \ \ \Phi(V)=U=FVF^\dagger.
\end{equation}
With these notations in place, we can proceed to introduce the
{\it quantum VP model} \cite{elitzur},
\begin{equation}\label{infinite_vector_potts}
H_{\sf VP}[\lambda]= -\frac{1}{2}\sum_i\
(V_i+\ \lambda\ U_iU_{i+1}^{\dagger}+ {\sf h.c.})
\end{equation}
(the link to its classical counterpart is worked out Section \ref{sec8.3}),
and to study its bond algebra \(\mathcal{A}_{\sf VP}\) generated by
\begin{eqnarray}
U_i U_{i+1}^{\dagger},\ \ U_i^{\dagger} U_{i+1},\ \ V_i,\ \
V_i^{\dagger}; \ \ \ \ \ i\in\mathbb{Z}.
\end{eqnarray}
It is readily
verified, either from Equation \eqref{weyl_group_algebra} or Equation
\eqref{wga_aut}, that the mapping
\begin{eqnarray}\label{infinite_sd_vector_potts}
U_{i-1}^{\dagger}U_i\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}\
V_i^{\dagger}, \ \ \ \ \ \ \ \
V_i^{\dagger} \ &\stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}& \
U_i^{\dagger}U_{i+1},
\end{eqnarray}
(together with the corresponding relations for the Hermitian conjugate
bonds) defines a self-duality isomorphism of \(\mathcal{A}_{{\sf
VP}}\) that maps \(\Phi_{{{\sf d}}}(H_{\sf VP}[\lambda])=\lambda H_{\sf
VP}[1/\lambda]\).
The existence of the self-duality automorphism $\Phi_{\sf d}$ is
intimately connected to (i) the {\it
connectivity} of the system, as determined by the structure of
interactions (specifically, in this case, by the non-commutativity
amongst bonds), and (ii) the local degrees
of freedom posses a local symmetry which, in this case, is captured by
the automorphism $\Phi$ of Equation \eqref{wga_aut}.
The fact that these two ingredients can be combined consistently
renders \(H_{\sf VP}\) self-dual. Although not all (self-) dualities
follow from this pattern, many do. An
interesting exception was presented in Section \ref{sec3.11}.
As explained in Section \ref{sec3.5}, $\Phi_{\sf d}$ determines the dual
variables for the problem. Since $U_i$ can we written in terms of bonds
as
\begin{equation}\label{UVUtVt}
U_i=\cdots (U_{i-3}^{\dagger}U_{i-2}) (U_{i-2}^{\dagger}U_{i-1})
(U_{i-1}^{\dagger}U_{i})=\prod_{j\leq i}
(U_{j-1}^{\dagger}U_j),
\end{equation}
the dual variables are
\begin{eqnarray}
\label{UVUtVt1}
\hat{V}_i&&\equiv \Phi_{\sf d}(V_i)=U_i^{\dagger}U_{i+1},\\
\hat{U}_i&&\equiv \Phi_{\sf d}(U_i)=\Phi_{\sf d}(\prod_{j\leq i}
U_{j-1}^{\dagger}U_j)=\prod_{j\leq i}\Phi_{\sf d}(
U_{j-1}^{\dagger}U_j)=\prod_{j\leq i} V_j^{\dagger}.\nonumber
\end{eqnarray}
The fact that \(\hat{U}_i\), \(\hat{V}_i\) satisfy the
same algebra as the \(U_i\), \(V_i\) affords a useful independent check
of the correctness of $\Phi_{{{\sf d}}}$ as a bond algebra isomorphism.
The product \( \Gamma_i =
U_i\tilde{U}_{i-1}^\dagger\) of a direct degree of freedom \(U_i\) and its
neighbor dual \(\tilde{U}_{i-1}^\dagger\) satisfy the non-local algebra
\begin{equation}\label{P_order_disorder}
\Gamma_i\Gamma_j=\omega\Gamma_j\Gamma_i,\ \ \ \ \ \ \mbox{if}\ \ i\neq
j, \ \ \ \mbox{and}\ \ \ \Gamma_i\Gamma_i^\dagger=\mathbb{1}.
\end{equation}
This suggests that if \(p>2\), the excitations of the model are
governed by the parafermionic statistics that were described in Reference
\cite{rajabpour_cardy} for classical \(p\)-state models, see the discussion
at the end of Section \ref{sec3.10}. For \(p=2\), Equation \eqref{P_order_disorder}
reduces to the fermionic algebra associated with the Ising model.
Next we show
that the group of symmetries of the VP model is non-Abelian for
\(p\geq3\). The first step
is to introduce two new operators \(C_{0 i}, C_{1 i}\) that
act on the basis states \(|s_i\rangle, s_i=0,\cdots, p-1\) of the
VP model at site \(i\) as follows:
\begin{equation}\label{defc0c1}
C_{0 i}|s_i\rangle=|-s_i\rangle, \ \ \ \ \ \ \ \
C_{1 i}|1-s_i\rangle=|1-s_i\rangle
\end{equation}
The arithmetic in these definitions is modular, modulo \(p\).
For example, if \(p=7\), then \( C_{0 i}|0\rangle=|-0\rangle=|0\rangle\),
\( C_{0 i}|1\rangle=|-1\rangle=|6\rangle\), and so on.
If \(p=2\), \( C_{0 i}=\mathbb{1}\) and \(C_{1 i}=\sigma^x\), so we
assume in what follows that \(p\geq3\).
Let us define \(\mathcal{C}_0=\prod_i C_{0 i}\) and
\(\mathcal{C}_1=\prod_i C_{1 i}\). \(\mathcal{C}_0\) is known
in the literature as the charge conjugation operator \cite{Henkel},
but, to the best of our knowledge, \(\mathcal{C}_1\) has not been
discussed before. Both \(\mathcal{C}_0\) and
\(\mathcal{C}_1\) are unitary {\it and} Hermitian, see Equation
\eqref{polyhedral1} below. Since the operators \(U_i\)
and \(V_i\) act basis states \(|s_i\rangle\) as
\begin{eqnarray}
U_i|s_i\rangle=\omega^{s_i}|s_i\rangle,\ \ \ \ \ \ \ \ \ \
V_i|s_i\rangle=|s_i-1\rangle,
\end{eqnarray}
it follows from Equation \eqref{defc0c1} that
\begin{eqnarray}\label{c0c1}
\mathcal{C}_0V_i\mathcal{C}_0&=&V_i^\dagger,\ \ \ \ \ \ \ \ \ \
\mathcal{C}_1V_i\mathcal{C}_1= V_i^\dagger, \\
\mathcal{C}_0U_i\mathcal{C}_0&=&U_i^\dagger,\ \ \ \ \ \ \ \ \ \
\mathcal{C}_1U_i\mathcal{C}_1=\omega U_i^\dagger.
\end{eqnarray}
Toghether with their Hermitian conjugates,
the relations of Equation \eqref{c0c1} can be used to show that
\(\mathcal{C}_0=\prod_i C_{0 i}\) and
\(\mathcal{C}_1=\prod_i C_{1 i}\) commute with \(H_{\sf VP}\).
Moreover,
\begin{equation}\label{polyhedral1}
\mathcal{C}_0^2=\mathcal{C}_1^2=(\mathcal{C}_0\mathcal{C}_1)^p=\mathbb{1}.
\end{equation}
This means \cite{rotman} that the group of symmetries of the VP model
generated by \(\mathcal{C}_0\), \(\mathcal{C}_1\)
provides a representation of the {\it non-Abelian} polyhedral group
\(P(2,2,p)\). The unitary \(\mathcal{C}_0\mathcal{C}_1\) generates
the well-known \(\mathbb{Z}_p\) subgroup of symmetries, since
\(\mathcal{C}_0\mathcal{C}_1=\prod_iV_i\).
As it turns out, many \(\mathbb{Z}_p\) models,
including the \(\mathbb{Z}_p\) gauge theories discussed in Sections
\ref{sec5.3} and \ref{sec6.3}, have this group among its
symmetries.
Our discussion of the VP model presented above
has the disadvantage that the dual
variables of Equation \eqref{UVUtVt1} are not strictly speaking
well defined operators (see Section \ref{sec3.5}). We remedy this by
considering a finite-size chain with self-dual BCs
\begin{equation}\label{finitesdvp}
H_{\sf VP}^N[\lambda]= -\frac{1}{2}\sum_{i=1}^N\
V_i-\frac{1}{2}\sum_{i=1}^{N-1}\ \lambda\ U_iU_{i+1}^{\dagger}-
\frac{\lambda}{2}U_N
+{\sf h.c.}\ \ .
\end{equation}
The first step in constructing the finite self-duality isomorphism
$\Phi_{\sf d}$ is to match the bonds at the boundaries of the chain
\begin{equation}\label{aut_finite_VP1}
U_N \ \stackrel{\Phi_{\sf d}}{\longrightarrow}\ \Phi(U_1)= V^{\dagger}_1,
\ \ \ \ \ \ \ V^{\dagger}_1 \ \stackrel{\Phi_{\sf d}}{\longrightarrow}\
\Phi(V^{\dagger}_N)=U_N^{\dagger} ,
\end{equation}
where \(\Phi\) was defined in Equation \eqref{wga_aut}. Next, we
extend the mapping to the remaining bonds,
\begin{eqnarray}
U_iU_{i+1}^{\dagger}\ &\stackrel{\Phi_{\sf d}}{\longrightarrow}& \
\Phi(U_{r(i)})=V^{\dagger}_{r(i)},\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
\ \ \ \ \ i=1,\cdots,N-1 \label{aut_finite_VP2}\nonumber \\
V^{\dagger}_{i}\ &\stackrel{\Phi_{\sf d}}{\longrightarrow}& \
\Phi(V^{\dagger}_{r(i)})U_{r(i)+1}=U^{\dagger}_{r(i)}U_{r(i)+1},
\ \ \ \ \ \ i=2,\cdots,N,
\label{uiui}
\end{eqnarray}
guided by the fact that we
must preserve the connectivity of the interactions and exploit the local
symmetry of the Weyl group algebra.
In Equation \eqref{uiui}, the reflection map $r$ of site $i$ is defined
as in Equation \eqref{inversion_map_eq} (i.e., $r(i)=N+1-i$).
Notice that the unitary \(\mathcal{U}_{{{\sf d}}}\) that implements
the mapping \(\Phi_{{\sf d}}\) is not
just the discrete Fourier transform $F^\dagger=\prod_{i=1}^N
F_i^\dagger$. The latter maps $H_{\sf VP}^N$ into $\tilde{H}_{\sf
VP}=F^\dagger {H}_{\sf VP}^N F$,
\begin{equation}
\tilde{H}_{\sf VP}^N[\lambda]= -\frac{1}{2}\sum_{i=1}^N\
U_i-\frac{1}{2}\sum_{i=1}^{N-1}\ \lambda\ V_iV_{i+1}^{\dagger}+ {\sf
h.c.}\ \ \ .
\end{equation}
Finally, we use \(\Phi_{{{\sf d}}}\) to compute well defined dual variables,
\begin{eqnarray}
\hat{V}^{\dagger}_{1} &\equiv& \Phi_{{\sf d}}(V^{\dagger}_1)=U^{\dagger}_N\nonumber\\
\hat{V}^{\dagger}_i &\equiv& \Phi_{{\sf d}}(V_i^{\dagger})=
U^{\dagger}_{r(i)}U_{r(i)+1}, \ \ \ \ \ \ i=2,\cdots,N,\\
\hat{U}_N &\equiv& \Phi_{{\sf d}}(U_N)=V_1^{\dagger},\nonumber\\
\hat{U}_i &\equiv& \Phi_{\sf d}(U_i)=V^{\dagger}_{r(i)}V^{\dagger}_{r(i)-1}\cdots
V^{\dagger}_{2}V^{\dagger}_{1},\ \ \ \ \ \
i=1,\cdots,N-1.\nonumber
\end{eqnarray}
According to the general ideas of Section \ref{sec3.9}, the square of the
self-duality unitary \(\mathcal{U}_{{{\sf d}}}\) commutes with the
Hamiltonian. One of the advantages of the bond algebraic approach is
that we do not need to compute \(\mathcal{U}_{{{\sf d}}}\) explicitly to figure
out the action of $\mathcal{U}^2_{{\sf d}}$ as an operator.
Since, by construction,
conjugation by $\mathcal{U}^2_{{\sf d}}$ amounts to applying $\Phi_{\sf d}$
twice, it follows from Equations \eqref{aut_finite_VP1} and
\eqref{aut_finite_VP2} that
\begin{equation}
\mathcal{U}^2_{{\sf d}}\ V_i\ \mathcal{U}^{2\ \dagger}_{\sf
d}=V^{\dagger}_i, \ \ \ \ \ \ \mathcal{U}^2_{{\sf d}}\ U_i\
\mathcal{U}^{2\ \dagger}_{{\sf d}}= U^{\dagger}_i,\ \ \ \
\ \ \ \ i=1,\cdots, N.
\label{UVC}
\end{equation}
Thus comparing with Equation \eqref{c0c1}, we see that
\(\mathcal{U}^2_{{\sf d}} =\mathcal{C}_0\), where \(\mathcal{C}_0\) is the
charge conjugation symmetry of the VP model as discussed above.
Notice also
that the self-dual BCs of Equation \eqref{finitesdvp} spoil the
\(\mathcal{C}_1\) symmetry of the (infinite) VP model. We could have used
self-dual BCs that preserve both symmetries, in agreement with the techniques
of Section \ref{sec3.7} and Appendix \ref{appG}.
\subsubsection{The Potts model and general \(\mathbb{Z}_p\) clock models}
\label{potts}
The Potts (P) model
\begin{eqnarray}\label{Pclassical}
\mathcal{Z}_{\sf P}=\sum_{\{\theta_{\bm{r}}\}}\ \exp\left[\sum_{{\bm{r}}}\sum_{\mu=1,2}\
K_\mu \delta(s_{{\bm{r}}}, s_{{\bm{r}}+\bm{e_\mu}})\right],
\ \ \ \ \ \ s_{\bm{r}}=0,1,\cdots,p-1,
\end{eqnarray}
in \(D=2\) dimensions is yet another
\(p\)-state generalization of the Ising model that, unlike
the VP model of the previous section, has a very well understood
statistical behavior \cite{martin} (\(\delta(s,s')=1\) if \(s=s'\),
and \(\delta(s,s')=0\) otherwise). The P model has a group of
global {\it non-Abelian} symmetries (if \(p>2\)),
the group \(\mathfrak{S}_p\) of permutations of \(p\) elements.
The \(d=1\) quantum rendition of the P
model \cite{solyom} reads
\begin{equation}
H_{\sf P}[\lambda]=\sum_i\sum_{m=0}^{[p/2]} \left[(V_i^{m}+V_i^{\dagger m})
+\lambda\ (U_i^m U_{i+1}^{\dagger m}+U_i^{\dagger m} U_{i+1}^{m})\right],
\end{equation}
where \([p/2]\) denotes the integer part of \(p/2\), that is,
the largest integer \(\leq p/2\). It follows that the P and VP model
coincide for \(p=2,3\). But even for arbitrary \(p\), since the bonds
in \(H_{\sf P}\) are powers of the bonds in \(H_{\sf VP}\), it is not
hard to see that the self-duality of the VP model, Equation
\eqref{infinite_sd_vector_potts}, implies the self-duality of
the P model.
The form of the Hamiltonians \(H_{\sf P}\) and \(H_{\sf VP}\) suggests
introducing a family of models that generalize them and interpolates between
them, the \(\mathbb{Z}_p\) clock models. They are defined by the Hamiltonians
\begin{equation}
H_{\sf GP}[\{\alpha_m\},\{\lambda_m\}]=
\sum_i\sum_{m=0}^{[p/2]} \left[\alpha_m(V_i^{m}+V_i^{\dagger m})
+\lambda_m\ (U_i^m U_{i+1}^{\dagger m}+U_i^{\dagger m} U_{i+1}^{m})\right],
\end{equation}
with \(\alpha_m,\lambda_m\in\mathbb{R}\), \(m=0,\cdots, [p/2]\), arbitrary
real coupling constants. The duality properties of some of these models
where studied in References \cite{rittenberg}, and the duality properties
of their classical counterparts where studied in Reference \cite{cardy}.
The \(\mathbb{Z}_p\) clock models do not have in general as many symmetries as their
special case the P model, {\it but they are all non-Abelian},
as the operators \(\mathcal{C}_0,\mathcal{C}_1\) of Equation \eqref{c0c1}
commute with \(H_{\sf GP}\) for any value of its couplings.
Next we can use our experience with the P and VP models
to identify immediately the \(\mathbb{Z}_p\) clock models that are
self-dual: The Hamiltonian \(H_{\sf GP}\) is self-dual
in those regions in coupling space
where \(\alpha_m\) vanishes if and only if \(\lambda_m\) vanishes as well.
Clearly, the P and VP model belong to two such regions.
\subsection{Dualities in some limits and related approximations}
\label{sec4.3}
It is standard and useful practice to think of some models as related through
a limit of some parameter. The VP model of Equation \eqref{classicalVP} is a good case
in point. It is intuitively reasonable to expect that its behavior
must approach that of the XY model as the discrete angle
\(\theta_{\bm{r}}\) becomes very dense in the limit \(p\rightarrow\infty\), and
in fact this is know to be the case in many respects \cite{frohlich}.
Dualities, however, can show some counter-intuitive behavior with respect
to limits. This is the topic of this section, with the VP model
as main illustration.
Weyl's group algebra describes states on a finite, equidistant set of points on
a unit circle. If we think of the roots of unity
\(\omega^0,\omega^1,\cdots, \omega^{p-1}\) as these points, then $U$ and
$U^\dagger$ will play the role of position operators, while \(V^\dagger\)
(\(V\)) acts as a clockwise (counter-clockwise) rotation from
any of the \(p\) roots to one of its neighbors. Following Schwinger
\cite{schwinger}, one can define position $\hat{\bf q}$, and momentum
$\hat{\bf p}$ Hermitian operators, such that
\begin{eqnarray}
{U}=e^{i \epsilon \hat{\bf q}} \ , \ {V}=e^{i \epsilon \hat{\bf p}} ,
\end{eqnarray}
with eigenvalues ${q}_r({p}_r)=0,\epsilon, 2\epsilon,
\cdots,(p-1)\epsilon$, and $\epsilon^2=2\pi/p$. In matrix form,
$\hat{\bf q}=\epsilon \, {\sf diag}(0,1,2,\cdots,p-1)$, and
\begin{eqnarray}
\hat{\bf p}=\epsilon \begin{pmatrix}
\frac{p-1}{2}& {\cal P}^*_1& {\cal P}^*_2& \cdots& {\cal P}^*_{p-1}\\
{\cal P}_1& \frac{p-1}{2}& {\cal P}^*_1& \cdots& {\cal P}^*_{p-2}\\
\vdots& \vdots& \vdots& & \vdots \\
{\cal P}_{p-2}& {\cal P}_{p-3}& {\cal P}_{p-4}& \cdots& {\cal P}^*_1\\
{\cal P}_{p-1}& {\cal P}_{p-2}& {\cal P}_{p-3}& \cdots& \frac{p-1}{2}\\
\end{pmatrix} = {F}^\dagger \hat{\bf q} {F},
\end{eqnarray}
with ${\cal P}_m=\frac{1}{p}\sum_{n=1}^{p-1} n \, \omega^{m n}$. In the
$p \rightarrow \infty$ limit, Weyl's algebra relates to the {\it
continuous} circle and it can be shown that $\epsilon \hat{\bf q}
\rightarrow \theta \in [0,2\pi)$, and $ \hat{\bf p} \rightarrow -i
\epsilon \partial/\partial \theta$. In this subtle limit, \(H_{\sf
VP}\) of Equation \eqref{infinite_vector_potts} converges to the
quantum version of the classical $D=2$ XY model \cite{bookEPTCP} (up to
irrelevant constants)
\begin{equation}\label{KT_H}
H_{\sf KT}=\sum_i\ (\frac{1}{2}L_i^2- \lambda \cos(\theta_{i+1}
-\theta_i) ),
\end{equation}
where \(L_j=-i \partial/\partial\theta_j\). Albeit being the $p \to
\infty$ limit of the self-dual VP model limit, $H_{\sf KT}$ {\it is not
self-dual}, but rather dual to the quantum SoS model (see
Section \ref{sec5.1}). It follows that {\it the limit of a sequence of
self-dual models needs not be self-dual}.
There is a profound difference between compact theories (such as those
of variables defined on a circle or the compact $U(1)$ fields of
electromagnetism) and non-compact theories (e.g., a theory whose fields
are defined on a line). An example of the latter which shares some
relation to $H_{\sf KT}$ is a linear chain of coupled harmonic
oscillators representing acoustic phonons \cite{kittel}
\begin{eqnarray}\label{phonons_H}
H_{\sf Ph}=\sum_i\ (\frac{p_i^{2}}{2m} +\ \frac{1}{2}m\omega^2
({x}_{i+1}-{x}_i)^2)\ , \ \ \ \ \ \ \ \ [x_i,p_j]=\ i\delta_{i,j} .
\end{eqnarray}
However, unlike $H_{\sf KT}$ the Hamiltonian \(H_{\sf Ph}\)
{\it is} self-dual under the exchange \(m\omega^2\leftrightarrow1/m\),
as the isomorphism
\begin{eqnarray}\label{phonons_aut}
{x}_{i+1}-{x}_{i}\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow} \ {p}_{i+1},
\ \ \ \ \ \ \ \ \
{p}_i\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow} \ {x}_{i+1}-{x}_{i}
\end{eqnarray}
shows. This self-duality arises due an interplay between the connectivity of
the model and the automorphism \(\Phi\) of the Heisenberg algebra
\begin{equation}
x\ \stackrel{\Phi}{\longrightarrow}\ p,\ \ \ \ \ \ \ \ \
p\ \stackrel{\Phi}{\longrightarrow}\ -x ,
\end{equation}
and generates the set of dual variables
\begin{eqnarray}\label{phonons_dual_var}
\hat{p}_i={x}_{i+1}-{x}_{i},\ \ \ \ \ \ \ \ \ \ \ \ \ \
\hat{x}_i=\sum_{m=-\infty}^i\ p_m.
\end{eqnarray}
The ground state of \(H_{\sf Ph}\) does not show a phase transition
at the self-dual point \(m\omega=1\), or anywhere else for that matter.
Thus we see that
while self-dualities can constrain phase transitions greatly, they
cannot guarantee by themselves that a phase transition will occur.
\subsection{The Xu-Moore model}
\label{sec4.4}
The \(d=2\) Xu-Moore (XM) Hamiltonian,
\begin{equation}\label{XM}
H_{\sf XM}[J,h]= -\sum_{{\bm{r}}}\ (J\square\sigma^z_{\bm{r}} + h\sigma^x_{\bm{r}}),
\end{equation}
with
\begin{equation}\label{plaquetteXM}
\square\sigma^z_{\bm{r}} =\sigma^z_{\bm{r}} \sigma^z_{\bm{r+e_2}}
\sigma^z_{\bm{r-e_1+e_2}} \sigma^z_{\bm{r-e_1}},
\end{equation}
was introduced in Reference \cite{XM}
as an effective model to study ordering in
arrays of Josephson-coupled \(p\pm ip\)
superconducting grains. \(H_{\sf XM}\) looks similar
to the \(\mathbb{Z}_2\) gauge theory studied Section \ref{sec3.12},
but the fact that the spins \(S=1/2\) are
now located at the {\it vertices} of a square lattice (see Figure
\ref{notation_links}) rather than at its links, makes the symmetries
and properties of the two models very different. In fact,
unlike the \(\mathbb{Z}_2\) gauge theory, the XM model
is self-dual \cite{XM}, and displays \(d=1\) dimensional
gauge-like symmetries \cite{tqo},
\begin{equation}
G_{r^1}=\prod_m\ \sigma^x_{r^1,m},\ \ \ \ \ \ G_{r^2}=\prod_m\ \sigma^x_{m,r^2},
\end{equation}
\([G_{r^i},\ H_{\sf XM}]=0\), that make the model a toy example
of topological quantum order and dimensional reduction \cite{tqo}
(see also Section \ref{6.8}).
Its self-duality is the subject of this section.
The bond algebra generated by the set of bonds \(\{\square\sigma^z_{\bm{r}},\sigma^x_{\bm{r}}\}_{\bm{r}}\)
is characterized by three relations:
Every bond (i) squares to one, (ii) anti-commutes with
four other neighboring bonds, and (iii) commutes with every other bond.
It follows that there is an isomorphism
\begin{eqnarray}\label{sdXM}
\square\sigma^z_{\bm{r}}\ \xrightarrow{\Phi_{\sf d}}\ \sigma^x_{\bm{r-
e_1+e_2}},
\ \ \ \ \ \ \ \
\sigma^x_{\bm{r}}\ \xrightarrow{\Phi_{\sf d}}\ \square\sigma^z_{\bm{r}},
\end{eqnarray}
illustrated in Figure \ref{sd_inf_XM}, that establishes the
self-duality of the XM model.
As it should, $\Phi_{{\sf d}}^2$ is a symmetry of the model,
\begin{equation}
\square\sigma^z_{\bm{r}}\ \xrightarrow{\Phi_{\sf d}^2}\
\square\sigma^z_{\bm{r-e_1+e_2}}, \ \ \ \ \ \ \ \ \ \ \sigma^x_{\bm{r}}\
\xrightarrow{\Phi_{\sf d}^2}\ \sigma^x_{\bm{r-e_1+e_2}},
\end{equation}
but it shows also unsettling features that are typical
of working directly in the limit of infinite size.
Since formally \(\mathbb{1}=\prod_m\ \square\sigma^z_{r^1,m}=
\prod_m\ \square\sigma^z_{m,r^2}\), it seems that one could argue that
\(\Phi_{{\sf d}}\) is in fact a multivalued mapping,
\begin{equation}
\Phi_{{\sf d}}(\mathbb{1})=\mathbb{1},\ \ \mbox{or}\ G_{r^1},\ \ \mbox{or}\ G_{r^2}.
\end{equation}
This is not a problem, however, in the light of the general discussion
of Section \ref{sec3.7}.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.5\columnwidth]{./sd_infiniteXM.eps}
\end{center}
\caption{ In this representation of the XM model,
the heavy crosses stand for the bonds \(\sigma^x_{\bm{r}}\) and
the dashed squares for the bonds \(\square\sigma^z_{\bm{r}}\). The self-duality
isomorphism \(\Phi_{{\sf d}}\) of Equation \eqref{sdXM} exchanges the
two types of bonds. Notice that the net result of applying the mapping
\(\Phi_{{\sf d}}\) {\it twice} amounts to a translation by \(\bm{d}=-{\bm{e_1}}+{\bm{e_2}}\),
a symmetry of the model. }
\label{sd_inf_XM}
\end{figure}
To make contact with the traditional approach described in Section
\ref{sec2.1} and exploited in Reference \cite{XM},
we need to employ the ideas of Section \ref{sec3.5}
to compute dual variables. Since
\begin{eqnarray}
\sigma^z_{\bm{r}}=\ \prod_{m^1 \leq r^1,\ r^2\leq m^2}
\ \square\sigma^z_{\bm{m}},
\end{eqnarray}
where $\bm{m}=m^1\bm{ e_1}+m^2\bm{e_2}$, it follows that
\begin{eqnarray}\label{dualXM}
\mu^x_{\bm{r}} \equiv \Phi_{\sf d}(\sigma^x_{\bm{r}})=\square\sigma^z_{\bm{r}}
,\ \ \ \ \ \ \ \mu^z_{\bm{r}} \equiv \Phi_{\sf d}(\sigma^z_{\bm{r}})=\
\prod_{m^1\leq r^1-1,\ r^2+1\leq m^2 }\ \sigma^x_{\bm{m}}.
\end{eqnarray}
This completes the calculation.
Let us discuss next self-dual, open BCs for the
XM Hamiltonian. Consider an square portion of the infinite model,
featuring \(N^2\) spins,
\begin{equation}\label {XM_H_finite}
H_{\sf XM}^o=-J\sum_{r^1=1}^{N-1}\sum_{r^2=0}^{N-2} \,
\square\sigma^z_{r^1,r^2} -h\sum_{r^1,r^2=0}^{N-1}\ \sigma^x_{r^1,r^2}
\end{equation}
($r^1,r^2 = 0,1,\cdots, N-1$).
The goal is to determine boundary corrections to make \(H_{\sf
XM}^o\) self-dual, and as usual, this could be accomplished by a systematic
study of \(H_{\sf XM}^o\)'s bond algebra. But this task grows increasingly
harder with dimension, so it is important to realize that
there is a simple recipe to construct self-dual boundary terms that
works well in general (but not always), and that we illustrate
next with the XM model. The starting point is the
self-duality mapping for the infinitely large model.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.55\columnwidth]{./finiteXM.eps}
\end{center}
\caption{One possible self-dual BC for the
the finite-size XM model on a \(3\times3\) portion of the
lattice ($N=3$). The bonds shown on the upper and left edge
(indicated with broken lines), and on the left upper corner (indicated
with a broken circle) render the the finite-size XM Hamiltonian self-dual,
but also break the \(d=1\)-dimensional gauge-like symmetries of the model.
There are other self-dual BCs that preserve some or all of
the gauge-like symmetries, but are not as convenient for computing dual
variables.}
\label{f_XM}
\end{figure}
So imagine that we try to apply the self-duality isomorphism of
Equation \eqref{sdXM} to the bonds \(\sigma^x_{\bm{r}}\) in \(H_{\sf XM}^o\).
The problem is that,
as defined in Equation \eqref {XM_H_finite}, \(H_{\sf XM}^o\) does
not have any plaquettes to which we can map the spins \(\sigma^x_{0,r^2}\)
at the left edge and the spins \(\sigma^x_{r^1,N-1}\) at the top edge (see
Figure \ref{f_XM}). This suggests that to restore self-duality, we
should re-introduce the missing plaquettes as boundary corrections,
or maybe incomplete versions of these plaquettes, in order to
preserve the number of spins.
As it turns out, this idea works perfectly, so it is convenient to
introduce some notation to describe complete and incomplete plaquettes
in a unified manner. We
write \(\square^I\sigma^z_{r^1,r^2}\)
for the plaquette, or {\it incomplete} plaquette, that results from simplifying the
standard plaquette of Equation \eqref{plaquetteXM} by discarding the
spins {\it outside} the square region encompassed by \(H^o_{\sf XM}\). So, for example,
\begin{equation}
\square^I\sigma^z_{0,0}=\sigma^z_{0,0}\sigma^z_{0,1},\ \ \ \
\square^I\sigma^z_{1,0}=\square\sigma^z_{1,0},\ \
\mbox{and}\ \ \ \square^I\sigma^z_{0,N-1}=\sigma^z_{0,N-1}.
\end{equation}
Our proposal is that
\begin{equation}\label{fselfdXM}
\tilde{H}_{\sf XM}^o=-\sum_{r^1,r^2=0}^{N-1}\ (J\square^I\sigma^z_{r^1,r^2}+
h\sigma^x_{r^1,r^2})
\end{equation}
is a finite self-dual rendition of the XM model (\(\tilde{H}_{\sf XM}^o\)
is illustrated in Figure \ref{f_XM}, for \(N=3\)). To prove this,
we must construct the self-duality isomorphism, starting by noticing that
\(\square^I\sigma^z_{0,N-1}=\sigma^z_{0,N-1}\) can only map to
\(\sigma^x_{N-1,0}\). The full isomorphism follows right away,
\begin{equation}\label{iso_finite_XM}
\square^I\sigma^z_{r^1,r^2}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{r^2,r^1},
\ \ \ \ \ \ \sigma^x_{r^1,r^2}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \square^I\sigma^z_{r^2,r^1}.
\end{equation}
Geometrically, it is a reflection along the main diagonal, and
$\Phi_{\sf d}^2=\mathbb{1}$.
The self-dual BC illustrated in Figure \ref{f_XM}
breaks the global \({\mathbb{Z}}_2\), and all the gauge-like \cite{tqo} symmetries
of the XM model, and one could have chosen among several other ones
that preserve some or all of these symmetries (see Section \ref{sec3.7}).
In particular, if we
just replace in Equations \eqref{fselfdXM} and \eqref{iso_finite_XM} the
incomplete plaquettes \(\square^I\sigma^z\) with standard, complete ones
\(\square\sigma^z\), the resulting model is self-dual under the same
mapping, and moreover has all the global and gauge-like symmetries
of the XM model. On the other hand, the self-dual BCs
that we chose guarantee that all of the spins
$\sigma^z_{(r^1,r^2)}$ contained in the square region encompassed by
$\tilde{H}_{\sf XM}^o$ are elements in its bond algebra. Thus we
can use the finite isomorphism just defined in
Equation \eqref{iso_finite_XM} to compute finite dual variables.
It is convenient to do so, at least for small system size, to check
the algebraic consistency of Equation \eqref{iso_finite_XM}. For \(N=3\),
\begin{eqnarray}
\mu^x_{0,2}&=&\square\sigma^z_{2,0},\ \ \ \ \ \ \ \ \ \ \ \
\mu^z_{0,2}=\sigma^x_{2,0},\nonumber\\
\mu^x_{0,1}&=&\square\sigma^z_{1,0},\ \ \ \ \ \ \ \ \ \ \ \
\mu^z_{0,1}=\sigma^x_{1,0} \sigma^x_{2,0},\nonumber\\
\mu^x_{0,0}&=&\sigma^z_{0,0}\sigma^z_{0,1},\ \ \ \ \ \ \ \ \ \
\mu^z_{0,0}=\sigma^x_{0,0}\sigma^x_{1,0} \sigma^x_{2,0},\nonumber\\
\mu^x_{1,2}&=&\square\sigma^z_{2,1},\ \ \ \ \ \ \ \ \ \ \ \
\mu^z_{1,2}=\sigma^x_{2,0}\sigma^x_{2,1},\\
\mu^x_{1,1}&=&\square\sigma^z_{1,1},\ \ \ \ \ \ \ \ \ \ \ \
\mu^z_{1,1}=\sigma^x_{1,0}\sigma^x_{1,1}\sigma^x_{2,1}\sigma^x_{2,0},
\nonumber\\
\mu^x_{1,0}&=&\sigma^z_{0,1}\sigma^z_{0,2},\ \ \ \ \ \ \ \ \ \
\mu^z_{1,0}=\sigma^x_{0,0}\sigma^x_{1,0}\sigma^x_{2,0}
\sigma^x_{0,1}\sigma^x_{1,1} \sigma^x_{2,1},\nonumber\\
\mu^x_{2,2}&=&\sigma^z_{1,2} \sigma^z_{2,2},\ \ \ \ \ \ \ \ \ \
\mu^z_{2,2}=\sigma^x_{2,0}\sigma^x_{2,1}\sigma^x_{2,2},\nonumber\\
\mu^x_{2,1}&=&\sigma^z_{0,2} \sigma^z_{1,2},\ \ \ \ \ \ \ \ \ \
\mu^z_{2,1}=\sigma^x_{1,0}\sigma^x_{1,1}\sigma^x_{1,2}
\sigma^x_{2,0}\sigma^x_{2,1}\sigma^x_{2,2},\nonumber\\
\mu^x_{2,0}&=&\sigma^z_{0,2},\ \ \ \ \ \ \ \ \ \ \ \ \ \ \
\mu^z_{2,0}=\sigma^x_{0,0}\sigma^x_{0,1} \sigma^x_{0,2}
\sigma^x_{1,0}\sigma^x_{1,1}\sigma^x_{1,2}
\sigma^x_{2,0}\sigma^x_{2,1}\sigma^x_{2,2}.\nonumber
\end{eqnarray}
\section{Quantum dualities by example: Lattice models}
\label{sec5}
\subsection{{\rm \bf XY}/solid-on-solid models}
\label{sec5.2}
The best developed approach to the Kosterlitz-Thouless phase transition
of the classical $D=2$ XY model exploits two duality transformations
(see for example Reference \cite{bookEPTCP}). First one maps the XY model,
via a Villain approximation, to the SoS model, to proceed
afterwards to map the SoS model to a classical $D=2$ Coulomb gas.
The two steps combined leads to the mapping between the XY model and a
$D=2$ Coulomb gas for which deconfinement of charges can be shown to occur at
sufficiently high temperatures.
In this section, we will study the first of these dualities to the
SoS model from a quantum, bond-algebraic perspective. The $D=2$
SoS model is specified by the partition function
\begin{equation}
\mathcal{Z}_{\sf SS}=\sum_{\{m_{\bm{r}}\}}\
\exp\left[\sum_{\bm{r}}\sum_{\nu=1,2}\ \
K_\nu(m_{{\bm{r}}+\bm{e_\nu}}-m_{\bm{r}})^2\right],\ \ \ \ \ \ \ \ m_{\bm{r}} \in
\mathbb{Z},
\label{SOSpart}
\end{equation}
where the classical degrees of freedom $m_{\bm{r}}$ reside on the vertices of
a square lattice (see Figure \ref{notation_links}).
The $d=1$ quantum version of the SoS model must have states
labelled by integers \(\{\vert m\rangle\}\). Three operators are then
required to describe its quantum dynamics: a position operator \(X\), and
left/right shift operators \(R/R^\dagger\),
\begin{equation}
X\vert m\rangle=\ m\, \vert m\rangle,\ \ \ \ \ \ \ \
R\vert m\rangle=\ \vert m-1\rangle,\ \ \ \ \ \ \ \
R^\dagger \vert m\rangle=\ \vert m+1\rangle,
\end{equation}
that satisfy
\begin{equation}\label{Z}
[X,R^\dagger]=R^\dagger,\ \ \ \ \ \ \ \ \ \ [X,R]=-R.
\end{equation}
\(X,\ R,\) and \(R^\dagger\) generate an algebra isomorphic to
the one satisfied by the
elementary degree of freedom associated to Equation \eqref{KT_H},
$[L,e^{\pm i\theta}]=\pm e^{\pm i\theta}$.
In particular, \(X\) and \(L\) have identical
spectra.
The duality mapping between the quantum versions of the XY and the
SoS models can be derived by comparing the bond
algebra generated by
\begin{equation}
L_i,\ \ \ \ \ \ \ \ e^{i(\theta_{i+1}-\theta_{i})},
\ \ \ \ \ \ \ \ e^{-i(\theta_{i+1}-\theta_{i})},
\end{equation}
which is characterized by the relations
\begin{equation}
[L_i,\ e^{\pm i(\theta_{j+1}-\theta_{j})}]=\ (\pm \delta_{i, j+1} \mp
\delta_{i, j})\ e^{\pm i(\theta_{j+1}-\theta_{j})},
\end{equation}
and the one generated by the SoS operators
\begin{eqnarray}
[(X_{i+1}-X_i),
R_{j+1}]&=&(\delta_{i,j+1}-\delta_{i,j})R_{j+1},\nonumber\\
{[}(X_{i+1}-X_i),
{R^\dagger_{j+1}}]&=&(-\delta_{i,j+1}+\delta_{i,j})R^\dagger_{j+1}.
\end{eqnarray}
Then, the isomorphism
\begin{equation} \label{iso_XY_ss}
L_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ (X_{i+1}-X_i),\ \ \ \ \ \ \ \
e^{i(\theta_{i+1}-\theta_{i})}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
R_{i+1}
\end{equation}
establishes a duality between the two models (a related mapping in terms of
less well-defined operators was described in Reference \cite{mattis}).
The dual form of \(H_{\sf KT}\) reads
\begin{eqnarray}
H_{\sf SS}= \frac{1}{2}\sum_i\ \left (- \lambda(R_i +
R^\dagger_i)+(X_{i+1}-X_i)^2\right )\ .
\end{eqnarray}
which to our knowledge has not been reported in the literature before.
One can verify, using the standard quantum-classical mapping
\cite{bookEPTCP}, that the classical counterpart of \(H_{\sf SS}\) is
indeed the classical $D=2$ SoS model of Equation
\eqref{SOSpart}.
This duality is similar, in spirit, to the self-duality of the quantum
Ising chain as seen through the dual variables
\begin{eqnarray}
L_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \widehat{L}_i=X_{i+1}-X_i, \ \
\ \ \ \ \ \ \ \ e^{i\theta_i}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
\widehat{e^{i\theta_i}} =\prod_{j<i}\ R_j\ .
\end{eqnarray}
The inverse \(\Phi_{{\sf d}}^{-1}\) of the isomorphism of Equation \eqref{iso_XY_ss}
defines a reciprocal set of dual variables.
\subsection{Xu-Moore/planar orbital compass models}
\label{sec5.1}
The planar orbital compass (POC) model of orbital ordering
\cite{orbital_compass},
\begin{equation}
H_{\sf POC}[J_x,J_z]=\ \sum_{\bm{r}}\ (J_x\sigma^x_{\bm{r}}\sigma^x_{{\bm{r}}+\bm{e_1}}+
J_z\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+\bm{e_2}}),
\label{3dOC}
\end{equation}
features spins $S=1/2$ residing on the vertices of a square
lattice (see Figure \ref{notation_links}), and interacting
in such a way that strongly correlates directions in real and spin space.
The POC model is dual to the XM model studied in Section \ref{sec4.3}
\cite{NF} (for another duality see \cite{vidal_Thomale_dusuel_schmidt}),
as follows from the duality isomorphism
\begin{equation}
\label{xme}
\square\sigma^z_{\bm{r}}\ \stackrel{\Phi_{\sf d}}{\longrightarrow} \
\sigma^x_{\bm{r-e_1}}\sigma^x_{\bm{r}},\ \ \ \ \ \ \
\ \sigma^x_{\bm{r}} \ \stackrel{\Phi_{\sf d}}{\longrightarrow} \
\sigma^z_{\bm{r-e_2}}\sigma^z_{\bm{r}}\ ,
\end{equation}
so that \( \Phi_{\sf d}(H_{\sf XM}[J,h])=H_{\sf POC}[J,h] \).
If follows that {\it the POC model is self-dual}.
The inverse duality transformation reads
\begin{equation}
\label{xme-}
\sigma^x_{\bm{r}} \sigma^x_{\bm{r+e_1}}\
\stackrel{\Phi_{\sf d}^{-1}}{\longrightarrow} \
\square\sigma^z_{\bm{r+e_1}}, \ \ \ \ \ \ \ \
\sigma^z_{\bm{r}} \sigma^z_{\bm{r+e_2}} \
\stackrel{\Phi_{\sf d}^{-1}}{\longrightarrow} \
\sigma^x_{\bm{r+e_2}},
\end{equation}
and both $\Phi_{{{\sf d}}}$ and $\Phi_{{{\sf d}}}^{-1}$ define their own set
of variables. {}From Equation \eqref{xme},
\begin{eqnarray}
\mu^x_{\bm{r}} =\sigma^z_{\bm{r-e_2}}\sigma^z_{\bm{r}},
\ \ \ \ \ \
\mu^z_{\bm{r}}=\Phi_{{\sf d}} \left(\prod_{m^1\leq
r^1,r^2\leq m^2}\square\sigma^z_{\bm{m}}\right)=
\prod_{n=0}^{\infty}\ \sigma^x_{\bm{r}+n\bm{e_2}},
\nonumber
\end{eqnarray}
so that $H_{\sf XM}(\mu)= H_{\sf POC}(\sigma)$ ($H(\cdot)$ means the
Hamiltonian $H$ written in term of the variables inside the brackets). Similarly, from
Equation \eqref{xme-},
\begin{eqnarray}
\tau^z_{\bm{r}}&=& \Phi_{\sf d}^{-1}(\sigma^z_{\bm{r}})= \Phi_{{\sf d}}
^{-1}\left(\prod_{n=0}^{\infty}\sigma^z_{\bm{r}-(n+1)\bm{e_2}}
\sigma^z_{\bm{r}-n\bm{e_2}}\right)=\prod_{n=0}^\infty\
\sigma^x_{\bm{r}-n\bm{e_2}},\\
\tau^x_{\bm{r}}&=& \Phi_{\sf d}^{-1}(\sigma^x_{\bm{r}})= \Phi_{{\sf d}}
^{-1}\left(\prod_{n=0}^{\infty}\sigma^x_{\bm{r}-(n+1)\bm{e_1}}
\sigma^x_{\bm{r}-n\bm{e_1}}\right)=\prod_{n=0}^{\infty}\
\square\sigma^z_{\bm{r}-n\bm{e_1}}=\sigma^z_{\bm{r}}
\sigma^z_{\bm{r+e_2}},\nonumber
\end{eqnarray}
so that $H_{\sf POC}(\tau)=H_{\sf XM}(\sigma)$.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.85\columnwidth]{./XM_2dOC.eps}
\end{center}
\caption{The POC model (left panel) restricted to a finite square
section of its lattice is dual to the XM model (right panel)
restricted to the same square section, provided both models are endowed
with suitable BCs. On the left panel (POC model), the
broken circles represent spins \(\sigma^x_{0,i}, \ i=0,1,2\), and the
crosses on the lower edge represent spins \(\sigma^z_{i,0}, \ i=0,1,2\), while
the broken (solid) links represent bonds \(\sigma^x_{i,j} \sigma^x_{i+1,j}\)
(\(\sigma^z_{i,j} \sigma^z_{i,j+1}\)). The right panel was explained in Figure
\ref{f_XM}.}
\label{XM_2dOC}
\end{figure}
Boundary corrections are
required to preserve the duality for finite lattices. One possible set
of boundary terms is shown in Figure \ref{XM_2dOC}, for a $3\times 3$
lattice. Both models, including the boundary corrections, have now
nine $\sigma^z$-like and nine $\sigma^x$-like bonds. The duality
isomorphism is given below,
\begin{alignat}{3}
\square\sigma^z_{1,0}&\ \leftrightarrow\ \sigma^x_{0,0}\sigma^x_{1,0},&\qquad
\square\sigma^z_{2,0}\ & \leftrightarrow\ \sigma^x_{1,0}\sigma^x_{2,0}, &\qquad
\square\sigma^z_{1,1}\ & \leftrightarrow\ \sigma^x_{0,1}\sigma^x_{1,1},\nonumber\\
\square\sigma^z_{1,2}&\ \leftrightarrow \ \sigma^x_{1,1}\sigma^x_{1,2},&\qquad
\sigma^z_{1,2}\sigma^z_{2,2}\ &\leftrightarrow\ \sigma^x_{1,2}\sigma^x_{2,2},&\qquad
\sigma^z_{0,2}\sigma^z_{1,2}\ & \leftrightarrow\ \sigma^x_{0,2}\sigma^x_{1,2}, \nonumber\\
\sigma^z_{0,2}&\ \leftrightarrow \ \sigma^x_{0,2}, &\qquad
\sigma^z_{0,1}\sigma^z_{0,2} \ &\leftrightarrow\ \sigma^x_{0,1},&\qquad
\sigma^z_{0,0}\sigma^z_{0,1}\ &\leftrightarrow\ \sigma^x_{0,0},\nonumber \\
\sigma^x_{0,0}&\ \leftrightarrow \ \sigma^z_{0,0},&\qquad
\sigma^x_{1,0}\ &\leftrightarrow\ \sigma^z_{1,0},&\qquad
\sigma^x_{2,0}\ & \leftrightarrow\ \sigma^z_{2,0},\nonumber\\
\sigma^x_{0,1}&\ \leftrightarrow \ \sigma^z_{0,0}\sigma^z_{0,1},&\qquad
\sigma^x_{1,1}\ & \leftrightarrow\ \sigma^z_{1,0}\sigma^z_{1,1},&\qquad
\sigma^x_{2,1}\ & \leftrightarrow\ \sigma^z_{2,0}\sigma^z_{2,1},\nonumber\\
\sigma^x_{0,2}&\ \leftrightarrow \ \sigma^z_{0,1}\sigma^z_{0,2},&\qquad
\sigma^x_{1,2}\ & \leftrightarrow\ \sigma^z_{1,1}\sigma^z_{1,2},&\qquad
\sigma^x_{2,2}\ & \leftrightarrow\ \sigma^z_{2,1}\sigma^z_{2,2}.
\label{longtra}
\end{alignat}
In Equation \eqref{longtra}, the bonds to the left of the double arrow
are those of the XM model and those to the right denote bonds of
the orbital compass model. Albeit being tedious, it is straightforward
to extend Equation \eqref{longtra} to an $N \times N$ square lattice.
This explicit transformation enables the computation of all dual
variables.
\subsection{Two-dimensional \({\mathbb{Z}}_p\) gauge/vector Potts models}
\label{sec5.3}
In this section we study a gauge-reducing duality, along the
lines of Section \ref{sec3.12}.
The $d=2$ dimensional \(\mathbb{Z}_p\) gauge theory \cite{horn},
\begin{equation}
H_{\sf G}=\frac{1}{2}\sum_{\bm{r}}\ \left(V_{({\bm{r}},1)}+V_{({\bm{r}},2)}+\
\lambda\ B_{({\bm{r}},3)}\ +\ {\sf h.c.}\right),
\end{equation}
with
\begin{equation}\label{B_plaquette}
B_{({\bm{r}},3)}\equiv\ U_{({\bm{r}},1)}U_{({\bm{r}}+\bm{e_1},2)}U^\dagger_{({\bm{r}}+\bm{e_2},1)}
U^\dagger_{({\bm{r}},2)},
\end{equation}
is a generalization of the Ising gauge model studied in Section \ref{sec3.12},
Equation \eqref{ising_gauge}.
The operators \(U_{({\bm{r}},\nu)}, V_{({\bm{r}},\nu)}\) (\(\nu=1,2\)), located at
the links (see Figure \ref{notation_links}) of a square lattice, commute
on different links, and satisfy the algebra described
in Section \ref{sec4.1.1} otherwise. As the name of the model suggests,
\(H_{\sf G}\) displays a gauge \(\mathbb{Z}_p\) symmetry
as realized by the local symmetry operators
\begin{equation}
G_{\bm{r}}=V_{({\bm{r}},1)}V_{({\bm{r}},2)}V^\dagger_{({\bm{r}}-{\bm{e_1}},1)}V^\dagger_{({\bm{r}}-{\bm{e_2}},2)}.
\end{equation}
However, the {\it global} symmetries of the model are {\it non-Abelian},
as can be seen by folowing the discussion in Section \ref{sec4.1.1}
of the symmetries of the VP model.
We want to find a transformation to a dual Hamiltonian which is free of
gauge symmetries. The $d=2$ dimensional quantum VP model (a
higher-dimensional version of Equation \eqref{infinite_vector_potts})
\begin{equation}
H_{\sf VP}= \frac{1}{2}\sum_{\bm{r}}\ \left(\lambda V_{\bm{r}}+\ U_{\bm{r}}
U_{{\bm{r}}+{\bm{e_1}}}^\dagger+ U_{\bm{r}} U_{{\bm{r}}+{\bm{e_2}}}^\dagger\ +\ {\sf h.c.}\right),
\end{equation}
is a natural candidate, because it is a
generalization of the $d=2$ Ising model in a transverse field in
terms of Weyl group algebra operators (defined on the vertices of the
lattice, see Figure \ref{notation_links}). It is not hard to
find a bond algebra homomorphism that confirms the duality,
\begin{eqnarray}\label{I_G_aut}
B_{({\bm{r}},3)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ V_{\bm{r}},\ \ \ \ V_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ U_{{\bm{r}}-{\bm{e_2}}}U^\dagger_{\bm{r}},\ \ \ \
V_{({\bm{r}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ U_{{\bm{r}}-{\bm{e_1}}}^\dagger U_{\bm{r}} ,
\end{eqnarray}
{\it up to the complete elimination of gauge symmetries},
\begin{eqnarray}\label{G_I_trivial}
\Phi_{{\sf d}}(G_{\bm{r}})= U_{{\bm{r}}-{\bm{e_2}}}U_{\bm{r}}^\dagger\times U_{{\bm{r}}-{\bm{e_1}}}^\dagger U_{\bm{r}}
\times U_{{\bm{r}}-{\bm{e_2}}-{\bm{e_1}}}^\dagger U_{{\bm{r}}-{\bm{e_1}}}\times
U_{{\bm{r}}-{\bm{e_1}}-{\bm{e_2}}}U_{{\bm{r}}-{\bm{e_2}}}^\dagger =\mathbb{1}.
\end{eqnarray}
The homomorphism of Equation \eqref{I_G_aut} affords a simple and
conceptually clarifying proof that \(H_{\sf VP}\) encodes the observable, gauge-invariant,
physics of \(H_{\sf G}\). On the other hand, it
cannot be used to define dual variables, see Section \ref{no_dual_vars}.
\subsection{Two-dimensional compact QED and {\rm \bf XY} models}
\label{sec5.4}
The elimination of gauge symmetries has a slightly different flavor
for models that feature continuous (or merely infinite) degrees of freedom,
essentially because it becomes more convenient to work
with the Hermitian generators of the gauge symmetries, rather than with the
unitary symmetries themselves. This section presents two examples of this kind,
a gauge-reducing duality for \(d=2\) compact QED to the \(d=2\)
SoS model, and a duality from a gauge SoS model to
the \(d=2\) XY model. These {\it quantum} dualities are new to the best of our
knowledge, but a classical (\(D=3\)) relative (performed by way of the Villain
approximation) of the duality for the \(d=2\) XY model was first
used in Reference \cite{peskin}.
The Hamiltonian for \(d=2\) compact QED that follows from
Wilson's lattice QED \cite{wilson} can be worked out along the
lines of Reference \cite{kogut_susskind},
\begin{equation}\label{LEMh}
H_{\sf LEM}=\sum_{\bm{r}}\ (\frac{1}{2}\ L_{({\bm{r}},1)}^2+\
\frac{1}{2}\ L_{({\bm{r}},2)}^2-\lambda\ \cos \Theta_{({\bm{r}},3)}),
\end{equation}
It features continuous angular variables $\theta_{({\bm{r}},\nu)} \in
[0,2\pi)$ defined on the links of a square lattice, with
\begin{equation}
L_{({\bm{r}},\nu)}=-i\frac{\partial}{\partial\theta_{({\bm{r}},\nu)}}\ ,
\ \ \ \ \ \ \ \ \Theta_{({\bm{r}},3)}\ =\ \theta_{({\bm{r}},1)}+\theta_{({\bm{r}}+{\bm{e_1}},2)}-
\theta_{({\bm{r}}+{\bm{e_2}},1)}-\theta_{({\bm{r}},2)},
\end{equation}
and where the elementary degrees of freedom satisfy the operator algebra
\begin{eqnarray}
[L_{({\bm{r}},\mu)},e^{\pm i \theta_{({\bm{r}}',\nu)}}]=\pm \delta_{{\bm{r}},{\bm{r}}'} \delta_{\mu,\nu}
e^{\pm i \theta_{({\bm{r}}',\nu)}}.
\label{CQEDcomr}
\end{eqnarray}
The gauge symmetries of Wilson's action translates in the Hamiltonian
language to the fact that
\(H_{\sf LEM}\) commutes with all the star operators
\begin{equation}
g_{\bm{r}}=L_{({\bm{r}},1)}+L_{({\bm{r}},2)}-L_{({\bm{r}}-{\bm{e_1}},1)}-L_{({\bm{r}}-{\bm{e_2}},2)}.
\end{equation}
These are the infinitesimal generators of gauge symmetries.
The gauge-reducing duality that we describe next is a hybrid between the dualities
presented in Sections \ref{sec5.2} and \ref{sec5.3}. It allow
us to recast the gauge invariant information contained in \(H_{\sf LEM}\) in
a form that is free of gauge redundancies. The dual, completely gauge
reduced model is the SoS model
\begin{equation}
H_{\sf SS}=\frac{1}{2}\sum_{\bm{r}}\ (-\lambda\ (R_{\bm{r}}+R_{\bm{r}}^\dagger)+
(X_{{\bm{r}}+\bm{e_1}}-X_{\bm{r}})^2+\
(X_{{\bm{r}}+\bm{e_2}}-X_{\bm{r}})^2),
\end{equation}
that features elementary degrees of freedom placed on the sites \({\bm{r}}\) of
a square lattice, and can be connected to the classical
\(D=3\) SoS model through a STL decomposition (see
Section \ref{classical&quantum}). The operators \(R_{\bm{r}}, R_{\bm{r}}^\dagger, X_{\bm{r}}\)
at site \({\bm{r}}\) satisfy
the relations of Equation \eqref{Z}, and commute with the operators on
other sites. Notice that \(H_{\sf SS}\) possesses global symmetries only.
The connection between \(H_{\sf LEM}\) and
\(H_{\sf SS}\) is established through the homomorphism of bond algebras
\begin{eqnarray}
e^{i\Theta_{({\bm{r}},3)}}\ & \stackrel{\Phi_{{\sf d}}}{\longrightarrow}&\ R_{\bm{r}},
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \
e^{-i\Theta_{({\bm{r}},3)}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow} \ R_{\bm{r}}^\dagger,
\\
L_{({\bm{r}},2)} \
&\stackrel{\Phi_{{\sf d}}}{\longrightarrow}& \ (X_{\bm{r}}-X_{{\bm{r}}-{\bm{e_1}}}),\ \ \ \ \ \ \ \ \ \ \ \ \,
L_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow} \ -(X_{\bm{r}}-X_{{\bm{r}}-{\bm{e_2}}}).
\nonumber
\end{eqnarray}
That the gauge symmetries of \(H_{\sf LEM}\) are
trivialized (eliminated) by the duality follows from the computation
\begin{eqnarray}
\Phi_{{\sf d}}(g_{\bm{r}})&=&-(X_{\bm{r}}-X_{{\bm{r}}-{\bm{e_2}}})+(X_{\bm{r}}-X_{{\bm{r}}-{\bm{e_1}}})+\nonumber\\
&&\ \ \ \ (X_{{\bm{r}}-{\bm{e_1}}}-X_{{\bm{r}}-{\bm{e_2}}-{\bm{e_1}}})-(X_{{\bm{r}}-{\bm{e_2}}}-X_{{\bm{r}}-{\bm{e_1}}-{\bm{e_2}}})=0.
\end{eqnarray}
Next we consider a duality that is in some sense complementary to the
previous one. The $d=2$ XY model
\begin{equation}
H_{\sf KT}=\sum_{\bm{r}}\ (\frac{1}{2}L_{\bm{r}}^2\ -\lambda
\cos(\theta_{{\bm{r}}+{\bm{e_1}}}-\theta_{\bm{r}}) - \lambda \cos(\theta_{{\bm{r}}+{\bm{e_2}}}-\theta_{\bm{r}})
)
\end{equation}
is the completely gauge-reduced version of a gauge SoS
model,
\begin{equation}
H_{\sf GSS}=\frac{1}{2}\sum_{\bm{r}}\ (-\lambda
(R_{({\bm{r}},1)}+R_{({\bm{r}},1)}^\dagger)- \lambda
(R_{({\bm{r}},2)}+R_{({\bm{r}},2)}^\dagger)+ b_{({\bm{r}},3)}^2),
\end{equation}
where \(b_{({\bm{r}},3)}=X_{({\bm{r}},1)} +X_{({\bm{r}}+{\bm{e_1}},2)}-X_{({\bm{r}}+{\bm{e_2}},1)}-X_{({\bm{r}},2)}\).
The generators of the group of gauge symmetries of \(H_{\sf GSS}\) are
\begin{equation}
G_{\bm{r}}=R_{({\bm{r}},1)}R_{({\bm{r}},2)}R^\dagger_{({\bm{r}}-{\bm{e_1}},1)}R^\dagger_{({\bm{r}}-{\bm{e_2}},2)},
\ \ \ \ \ \ \ \ \ \ \ \ \mbox{ and } \ G_{\bm{r}}^\dagger,
\end{equation}
but these are not infinitesimal. The gauge-reducing duality to the
XY model is established by the homomorphism of bond algebras
\begin{eqnarray}
&&X_{({\bm{r}},1)}+X_{({\bm{r}}+{\bm{e_1}},2)}-X_{({\bm{r}}+{\bm{e_2}},1)}-X_{({\bm{r}},2)} \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ L_{\bm{r}},\\
&&R_{({\bm{r}}+{\bm{e_2}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ e^{-i(\theta_{{\bm{r}}+{\bm{e_2}}}-\theta_{\bm{r}})}, \ \ \ \
R^\dagger_{({\bm{r}}+{\bm{e_1}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ e^{-i(\theta_{{\bm{r}}+{\bm{e_1}}}-\theta_{\bm{r}})},\nonumber
\end{eqnarray}
that satisfies
\begin{equation}
\Phi_{{\sf d}}(G_{\bm{r}})=e^{-i(\theta_{\bm{r}}-\theta_{{\bm{r}}-{\bm{e_2}}})}e^{i(\theta_{\bm{r}}-\theta_{{\bm{r}}-{\bm{e_1}}})}
e^{i(\theta_{{\bm{r}}-{\bm{e_1}}}-\theta_{{\bm{r}}-{\bm{e_2}}-{\bm{e_1}}})}e^{-i(\theta_{{\bm{r}}-{\bm{e_2}}}-\theta_{{\bm{r}}-{\bm{e_1}}-{\bm{e_2}}})}=
\mathbb{1}.
\end{equation}
\subsection{Toric code/\(\mathbb{Z}_2\) Higgs models}
\label{sec5.5}
Over the last fifteen years quantum computation has become a
well developed theoretical discipline, fostering a
paradigm-breaking new understanding of computational complexity and
quantum mechanics \cite{mermin}. In contrast, the technological
problem of building a quantum computer remains essentially unsolved,
and one of the biggest challenges is the realization of quantum memories.
Kitaev's toric code (TC) model \cite{toric_code} (see Equation
\eqref{etch}) is an excellent example of the virtues and pitfalls of one
of the most popular approaches to the problem of storing quantum
information: the use of topological quantum order \cite{tqo}.
While the TC model is a good example of topological quantum
order, it fails as a quantum memory at any finite temperature
\cite{tqo}. This surprising result, known as {\it thermal fragility}
\cite{tqo,3detc}, could be proved and probed in detail thanks to the
realization \cite{tqo} that the TC model is
(exactly) dual to two decoupled Ising chains in zero magnetic field,
a duality established by arguments that are direct precursors of the
bond-algebraic machinery.
Whether or not topological quantum order turns out to be key to the
implementation of quantum memories or the quantum computer, it is clear
by now that, as an order of matter that goes beyond the Landau
symmetry-breakdown paradigm, it is worth studying in its own right.
In this section we show that two popular models to study
topological quantum order, the {\it extended} toric code (ETC) model
in two \cite{extended_tc} and three space dimensions, are
dual to the \(\mathbb{Z}_2\) Higgs model \cite{fradkin_shenker}. While
the duality in $d=2$ dimensions is already exploited in
Reference \cite{extended_tc} to help the numerical simulation
of the ETC model,
the duality in $d=3$ dimensions is one of the most interesting
new dualities reported in this paper. Both dualities are in fact special
cases of a general gauge-reducing duality for the \(\mathbb{Z}_2\) Higgs model
{\it that works in any number \(d\) of space dimensions}, and that
is also special in that it does not necessitates the introduction of
non-local string operators (recall the discussion of Section \ref{sec3.12.4}).
Other aspects of this duality will be discussed in Section \ref{sec6.3}.
Section \ref{sec_tqo_dim} presents a broader discussion of the role of dualities
in the study of topological quantum order.
The \(\mathbb{Z}_2\) Higgs model in \(d\) spatial dimensions
\begin{eqnarray}\label{z2_higgs}
H_{\sf dH}=-\sum_{\bm{r}} ( J_x\sigma^x_{\bm{r}} +
\sum_{\nu=1}^d (h_z\sigma^z_{{\bm{r}}}
\sigma^z_{({\bm{r}},\nu)}\sigma^z_{{\bm{r}}+\bm{e_\nu}}+h_x\sigma^x_{({\bm{r}},\nu)} )
+ J_z\sum_{\nu<\mu}\ B_{({\bm{r}},\nu \mu)}),
\end{eqnarray}
features spin \(1/2\) degrees of freedom on the sites and links
of a hyper-cubic lattice. It can be thought of as a lattice,
two-state approximation to a complex Higgs field \(\phi\) of fixed modulus
\(\phi\phi^*=1\) (or in its broken symmetry phase \cite{savit}), in interaction
with electromagnetism, and it represents one of the best understood models of
confinement in dimensions \(d=2\) and \(d=3\).
Its gauge symmetries are
\begin{equation}
G_{\bm{r}}=\sigma^x_{\bm{r}}\ \prod_{\nu=1}^d\
\sigma^x_{({\bm{r}},\nu)}\sigma^x_{({\bm{r}}-\bm{e_\nu},\nu)} \equiv \sigma^x_{\bm{r}}
A_{\bm{r}},\ \ \ \ \ \ \ \ \ \ [H_{\sf dH},G_{\bm{r}}]=0.
\end{equation}
The goal of this section is to find a completely gauge-reducing duality to take care
of this gauge redundancy.
The structure of \(H_{\sf dH}\)s bond algebra suggests the model
\begin{equation}
\label{KTCMd3}
H_{\sf dGRH}=-\sum_{\bm{r}} (J_x A_{\bm{r}}+
\sum_{\nu=1}^d\ ( h_z\sigma^z_{({\bm{r}},\nu)} + h_x\sigma^x_{({\bm{r}},\nu)})
+J_z\sum_{\nu<\mu}B_{({\bm{r}},\nu \mu)} ),
\end{equation}
for a gauge-reduced dual. \(H_{\sf dGRH}\) shows no local symmetries,
due to the couplings to external magnetic fields, and
the bond algebra homomorphism that connects it to \(H_{\sf dH}\)
follows naturally,
\begin{eqnarray}
\sigma^x_{\bm{r}} &\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & A_{\bm{r}},\ \ \ \ \ \ \ \ \ \
\sigma^z_{{\bm{r}}}\sigma^z_{({\bm{r}},\nu)}\sigma^z_{{\bm{r}}+\bm{e_\nu}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
\sigma^z_{({\bm{r}},\nu)}\nonumber\\
\ \nonumber\\
B_{({\bm{r}},\nu\mu)}&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & B_{({\bm{r}},\nu\mu)},\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \,
\sigma^x_{({\bm{r}},\nu)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{({\bm{r}},\nu)}.
\end{eqnarray}
To check that \(\Phi_{{\sf d}}\) is indeed gauge reducing, we compute
\begin{equation}
\Phi_{{\sf d}}(G_{\bm{r}})=\Phi_{{\sf d}}(\sigma^x_{\bm{r}}) \Phi_{{\sf d}}(A_{\bm{r}})=A_{\bm{r}} A_{\bm{r}}=\mathbb{1}.
\end{equation}
Notice that the completely gauge-reduced model \(H_{\sf dGRH}\) features only
degrees of freedom on the links of the lattice, and local bonds.
The duality just described works in any space dimension $d$. In
$d=1,2,3$, the models \(H_{\sf dGRH}\) have well known physical
meanings. The $d=1$ \(\mathbb{Z}_2\) Higgs model
\begin{equation}
H_{\sf 1H}=-\sum_i (J_x\sigma^x_i+
h_z\sigma^z_i\sigma^z_{(i,1)}\sigma^z_{i+1}+h_x\sigma^x_{(i,1)})
\end{equation}
is dual to
\begin{equation}
H_{\sf 1GRH}=-\sum_i (J_x \sigma^x_{(i-1,1)}\sigma^x_{(i,1)}
+h_z\sigma^z_{(i,1)}+h_x\sigma^x_{(i,1)} ),
\end{equation}
which is just an Ising chain in the presence of transverse and
longitudinal fields \cite{fradkin_shenker}. This means that \(H_{\sf
1H}\) has no phase transition when \(h_x\neq0\). We see that the gauge
field has opened a mass gap in the model.
In $d=2$ dimensions, the gauge reduced form of the \(\mathbb{Z}_2\)
Higgs model reads
\begin{equation}\label{etch}
H_{\sf 2GRH}=H_{\sf ETC}=-\sum_{\bm{r}} (J_x A_{\bm{r}}+
\sum_{\nu=1}^2 ( h_x\sigma^x_{({\bm{r}},\nu)}+h_z\sigma^z_{({\bm{r}},\nu)} )
+J_zB_{({\bm{r}},3)})
\end{equation}
(\(B_{({\bm{r}},3)}\) was defined in Equation \eqref{gauge_plaquette}).
This is exactly the ETC model of Reference \cite{extended_tc}
(if we further set \(h_x=h_z=0\), we recover Kitaev's TC model).
The duality maps the Coulomb phase \cite{fradkin_shenker} of the \(\mathbb{Z}_2\) Higgs
model to the topological quantum ordered state of the ETC model.
In $d=2$ dimensions, the \(\mathbb{Z}_2\) Higgs model is self-dual \cite{horn_yankielowicz},
which implies that the ETC model is self-dual as well \cite{extended_tc,con}. Let us check
this in the latter model. The self-duality isomorphism \cite{con} reads
\begin{equation}\label{ETC_aut}
\sigma^{x,z}_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}
\ \sigma^{z,x}_{({\bm{r}}+\bm{e_1},2)},\ \ \ \ \ \ \ \
\sigma^{x,z}_{({\bm{r}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}
\ \sigma^{z,x}_{({\bm{r}}+\bm{e_2},1)}
\end{equation}
that exchanges $J_x$ with $J_z$, and simultaneously $h_x$ with $h_z$ in
\(H_{\sf ETC}\). This is one of the rare instances where a self-duality
mapping is {\it local in the spins}. Other interesting dualities for
the ETC model are reported in \cite{con} (and its supplemental material).
The gauge-reduced version of \(H_{\sf 3H}\) is again intimately
connected to a $d=3$ dimensional generalization of the ETC model
\begin{eqnarray}
H_{\sf 3GRH}&=&-\sum_{\bm{r}}
\sum_{\nu=1}^3\ (h_x\sigma^x_{({\bm{r}},\nu)}+h_z\sigma^z_{({\bm{r}},\nu)} )
-\nonumber\\
&&\ \ \,
\sum_{\bm{r}}\ (J_x A_{\bm{r}}+ J_z (B_{({\bm{r}},1)}+\ B_{({\bm{r}},2)}+\ B_{({\bm{r}},3)})),
\end{eqnarray}
studied in Reference \cite{3detc} in the case of vanishing
magnetic fields. The phase diagram of the full model,
that we call the $d=3$ ETC model, can be obtained from the
literature on the Higgs model (see, for example, Reference
\cite{fradkin_shenker}, Figure 2).
\section{Bond-algebraic dualities in quantum field theory}
\label{sec6}
Over the years, some of the most interesting and ambitious
dualities have been conjectured in the context of quantum field
theory (QFT)
\cite{montonen_olive, AdS-review},
and any progress in the theory of non-Abelian dualities
should be tested against QCD (quantum chromodynamics).
The functional approach to QFT \cite{rivers} puts QFTs in a
language that resembles closely that of classical statistical mechanics.
Therefore there have been some attempts \cite{thooftI} at dualities
for path integrals of QFTs that resemble that of Kramers and Wannier
introduced in Section \ref{sec2.1}. However,
the progress in this direction has been limited (see though
References \cite{alvarez,non_Abelian,lozano}).
The situation improves considerably if the (Euclidean) path integral for the QFT of
interest is regularized by replacing the continuum for a lattice \cite{wilson}
(see Appendix \ref{appE}).
Then, for Abelian theories, one can use the machinery of Appendix \ref{appA} to
construct systematically regularized dual (Euclidean) field theories \cite{savit}.
But, as discussed in Section \ref{classical&quantum}, and illustrated
in Section \ref{sec8}, {\it this path-integral based, lattice approach to
dualities in QFT is covered, and in fact simplified} by the bond-algebraic techniques
of this paper \cite{con}. In this light, many of the dualities we have seen already
can be interpreted as bond-algebraic dualities for QFTs.
In contrast, this section aims to explore the extension of bond algebraic techniques
to operator-based quantizations of field theories. This is perfectly feasible
for some QFTs, but in general we do not know yet how to construct
a complete operator quantization of an {\it interacting} field theory, and
so in many cases the {\it bond algebra of an interacting QFT is not
well defined}. This means for instance that we could have trouble
deciding whether two operators in a QFT should commute or not, as
exemplified by the Schwinger term in QED \cite{schwinger_term}:
The canonical quantization of electromagnetism in interaction with the Dirac electron field
dictates the charge density operator should commute with the current, but in
fact this is inconsistent with the requirement that the theory should have
a ground state \cite{schwinger_term}. It follows that this
commutation relation must be changed relative to its canonical form.
There is an approximate approach to the operator quantization
of field theories that is specially compatible with bond-algebraic
techniques, and was popularized in the literature on confinement
under the name of {\it lattice Hamiltonian formalism}
\cite{kogut_susskind, kogut}.
The idea is to discretize the {\it classical} field theory first {\it only
in space} (that is, to
approximate it with a classical mechanical model featuring degrees of freedom
on a spatial lattice), and then proceed to quantization, that can now
be carried through by standard means (see Appendix \ref{appE}).
The resulting many-body quantum theory typically features non-relativistic bosons,
fermions, or rigid rotators in interaction, and contains a new parameter, the lattice
spacing \(a\). Of course, it lacks the symmetries characteristic
of the continuum, {\it but has in exchange a well defined operator content}, and
typically internal and gauge symmetries are well represented.
In what follows, we study the duality properties of
several QFTs in this approximation. But when possible,
we work also directly in the continuum, and show that the two approaches give
compatible results when a naive continuum limit \(a\rightarrow 0\) is taken.
\subsection{One-dimensional free and externally coupled bosonic field, and
the Kibble model}
\label{sec6.1}
The free, massless bosonic field in $d=1$ (i.e., 1+1 space-time dimensions)
affords the simplest example of a self-dual QFT \cite{witten,con}.
In dimensions \(d\geq 2\), a complete operator quantization is always available
for {\it free} fields \cite{takahashi}. This is not the case in dimension \(d=1\).
In particular, the Green's function for the massless boson field is too singular to
be interpreted in the sense of distributions. Still, its bond-algebra based on
canonical quantization
reflects its self-dual properties, and we can check them in the lattice Hamiltonian
approximation. In this approximation, the bosonic field reduces to a self-dual model of one
dimensional phonons, Section \ref{sec4.3}, and the phononic lattice dual variables
converge to the bosonic dual variables in the continuum.
Next we consider two simple extensions, the $d=1$ Kibble model in
Section \ref{1dKibblemodel}, and a multiplet of bosonic fields in interaction
with an external driving forces.
\subsubsection{Free, Massless bosonic field}
\label{triv_boson_sub}
The massless, free bosonic field in \(1+1\) dimensions, ($\mu,\nu=0,1$)
is described by the action
\begin{equation}
\label{1+1sb}
S_{\sf FB}=\int d^2x\ \frac{1}{2}\eta^{\mu \nu}
\partial_\mu\phi\partial_\nu\phi ,
\end{equation}
where \(\eta={\sf diag}[1,-1]\),
$x^{0} = t$ stands for the time coordinate, and
$x^1= x$ for the spatial coordinate.
Its canonical quantization proceeds by defining
the Hamiltonian
\begin{equation}\label{boson_H}
H_{\sf FB}=\int dx\ \left
(\frac{1}{2}\pi^2+\frac{1}{2}\left(\partial_1 {\phi}\right)^2 \right ),
\end{equation}
together with the equal-time commutation relations
\begin{equation}
[\phi(x),\pi(y)] =\ i\delta(x-y).
\end{equation}
One may think of $H_{\sf FB}$ as the quantum theory of a continuous
elastic line (see Reference \cite{kittel}, Chapter 2).
{}From the perspective of bond algebras, the new feature is
that we have an uncountable infinity of bonds, two bonds
\(\pi^2(x),\ \left(\partial_1 {\phi}\right)^2(x)\) per space point.
It is easier to characterize the bond algebra
in terms of \(\pi(x)\) and \(\partial_1\phi(x)\),
\begin{equation}
[\partial_1\phi(x),\pi(y)]=\ i\delta'(x-y),
\label{can_com}
\end{equation}
where $\delta'(x-y)=\partial_1 \delta(x-y)=\partial_x \delta(x-y)$ is
the spatial derivative of the Dirac delta function.
It is apparent from this relation that
\begin{equation}\label{boson_aut}
\partial_1\phi(x)\ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}\
\pi(x),\ \ \ \ \ \ \ \
\pi(x) \ \stackrel{\Phi_{{{\sf d}}}}{\longrightarrow}\ \partial_1\phi(x),
\end{equation}
is a self-duality isomorphism, since
\begin{equation}
[\Phi_{{{\sf d}}}(\partial_1\phi(x)),\Phi_{{{\sf d}}}(\pi(y))]=
[\pi(x),\partial_1\phi(y)]= -i\delta'(y-x)=
i\delta'(x-y),
\end{equation}
and \(\Phi_{{\sf d}}(H_{\sf FB})= H_{\sf FB}\).
Next we use Equation \eqref{boson_aut} to compute dual variables.
Since
\begin{equation}
\label{phi_int}
\phi(x)=\int_{-\infty}^x dy\ \partial_y\phi(y),
\end{equation}
the dual field variables are
\begin{eqnarray}\label{boson_dual_var}
\pi(x)&&\stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
\hat{\pi}(x)=\ \partial_1\phi(x),\\
\phi(x)&&\stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
\hat{\phi}(x)=\int_{-\infty}^x dy\ \widehat{\partial_1\phi}(y)=
\int_{-\infty}^x dy\ \pi(y). \nonumber
\end{eqnarray}
We may regard these dual fields as toy examples of Mandelstam variables
\cite{mandelstam_variables}, the variables that appear in the bosonization
of $d=1$ theories \cite{bosonization}.
Let us compare next the calculations in the continuum to the
predictions of the lattice Hamiltonian approach. If we discretize
in space the action of Equation \eqref{1+1sb}, and quantize afterwards,
we get a quantum model specified by
\begin{equation}
H_{\sf LFB}=\frac{1}{2a}\sum_i\ (\pi_i^2+(\phi_{i+1}-\phi_i)^2),\ \ \ \ \ \
\ \ [\pi_m,\phi_n]=i\delta_{m,n},
\end{equation}
with \(a\) the lattice spacing.
This model is essentially identical to the self-dual model of phonons studied
in Section \ref{sec4.3}, and thus self-dual as well (notice that while
\(\pi_i\) is dimensionless, \(\pi(x)\) has dimensions of \(1/a\)).
Its self-duality mapping can be read from Equation \eqref{phonons_aut},
\begin{equation}
\frac{\pi_i}{a}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \frac{\phi_{i+1}-\phi_i}{a},\ \ \ \
\frac{\phi_{i+1}-\phi_i}{a}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \frac{\pi_{i+1}}{a},
\end{equation}
and it clearly converges to the corresponding mapping in the continuum,
Equation \eqref{boson_aut}, in the naive continuum limit \(a\rightarrow 0\).
\subsubsection{The one-dimensional Kibble model}
\label{1dKibblemodel}
The self-duality mapping investigated in the previous section
is readily applicable to the Kibble model \cite{aitchison},
\begin{eqnarray}
H_{\sf K}=\frac{1}{2}\int dx\ \left(\pi^2+(\partial_1\phi)^2\right)
+ \frac{1}{2}\int dx\ dy\ \ \pi(x)V(x-y)\pi(y),
\end{eqnarray}
in \(d=1\) dimension. The $d=3$ version of this
model with Coulomb potential $V(\bm{x}) =e^2/|\bm{x}|$ has been
studied as an example of a model that violates Goldstone theorem
\cite{bookEPTCP}, due to the long-range nature of the Coulomb interaction.
The Kibble model has the unusual feature that the momenta $\pi(x)$
participate in the interaction term. In \(d=1\) this is remedied
by the duality of the previous section. The mapping of Equation
\eqref{boson_aut} shows that \(H_{\sf K}\) admits a dual representation
\begin{align}
H_{\sf K}^D= \frac{1}{2} \int dx\ \left(\pi^2+(\partial_1\phi)^2\right)
+\frac{1}{2}\int dx\ dy\ \ \partial_1\phi(x)\ V(x-y)\ \partial_1\phi(y),
\end{align}
so that now the interaction term involves only spatial gradients.
\subsubsection{Massless bosonic field coupled to classical external sources}
The self-duality of the free boson field survives even after we
have coupled it to classical, external sources \(A(x^0,x^1)\) and \(J(x^0,x^1)\),
\begin{equation}\label{boson_H}
H_{\sf B}=\int dx\ \left
(\frac{1}{2}(\pi-A)^2+\frac{1}{2}\left(\partial_1 {\phi}\right)^2+J\phi \right),
\end{equation}
that vanish outside some finite interval.
While the coupling \(J\phi\) is standard, the coupling \(\pi-A\) is not,
because \(A\) is not the vector potential for an EM field (for one thing,
\(\phi=\phi^\dagger\) is not charged). We introduce it anyways, because the
self-duality we are going to describe exchanges \(J\) and \(A\).
The next step is to apply to \(H_{\sf B}\) the self-duality mapping for
the free boson field, Equation \eqref{boson_dual_var}. After some
rearrangements (that include discarding a boundary term, since
$A$ has compact support), the resulting Hamiltonian reads
\begin{equation}
H_{\sf B}^D=\int dx\ \left
(\frac{1}{2}(\pi-A^D)^2+\frac{1}{2}\left(\partial_1
{\phi}\right)^2+J^D\phi \right)+ \frac{1}{2}\int dx\ \left(A^2-A^{D 2}
\right),
\end{equation}
with
\begin{equation}
A^{D}(x^0,x^1)=-\int_{x^1}^{\infty}dy\ J(x^0,y),\ \ \ \ \ \
J^D(x^0,x^1)=\frac{\partial A}{\partial x^1}(x^0,x^1).
\end{equation}
Since \(H_{\sf B}^D\) has the same structure as \(H_{\sf B}\) (up to
an additive c-number), we see that \(H_{\sf B}\) is still self-dual as
in the free case. Notice that \(\Phi_{{\sf d}}^2=\mathbb{1}\), since
\(A^{DD}=A\) and \(J^{DD}=J\).
\subsection{The Luttinger model}
\label{sec6.7}
Next we describe a duality for fermions in one dimension.
The Luttinger model describes a $d=1$ dimensional interacting
many-electron system in a box of size $\ell$. Its {\it Fermi surface}
consists of only two points, $k= \pm k_{F}$, corresponding to two types
of electrons moving to the right/left.
Henceforth, we denote the right
and left moving electrons by the fields $\psi_{1}$ and $\psi_{2}$
(which anti-commute) and construct a two component field $\psi^{\dagger}
= (\psi_{1}^{\dagger}, \psi_{2}^{\dagger})$. In the vicinity of the two
Fermi points $\pm k_F$ (i.e., for small $|q|$ and small $|k-k_{F}|$ for
$k>0$ or small $|k+k_{F}|$ for $k<0$), the free electron dispersion may
be linearised to read $\epsilon_{k+q} - \epsilon_{k} = \pm v_{F} q$ with
$v_{F}$ the Fermi velocity. The (spinless) Luttinger model is defined
by the Hamiltonian
\begin{equation}
\label{lutt_model}
H_L=\int_0^\ell dx \, \psi^\dagger\sigma^z\left(-i\frac{\partial}
{\partial x}\right)\psi + \int_0^\ell dx dy\ \psi^\dagger_1(x)
\psi_1(x)V(x-y)\psi^\dagger_2(y)\psi_2(y) .
\end{equation}
In Equation \eqref{lutt_model}, the first term describes the free
electron dispersion (with $v_{F}$ set to unity). This is augmented, in
the second term of Equation \eqref{lutt_model}, by density-density
interactions that couple left ($k<0$) and right ($k>0$) movers. By
explicitly constructing unitary transformations, Luttinger showed that
$H_L$ is unitarily equivalent to a non-interacting model, and thus it is
exactly solvable (see, for example, \cite{mattis_lieb} and references
therein). It has, however, the non-physical characteristic that in the
thermodynamic limit it displays an infinite reservoir of negative
energy states.
We will next re-derive this unitary equivalence within the bond algebraic
framework. The Hamiltonian \(H_L\) commutes with particle number
operators for each fermion species (left and right movers), and can be
written in a first-quantized form as
\begin{equation}\label{luttinger_first_quantized}
H_L=\sum_{m=1}^M\ p_m -\sum_{n=1}^N\ P_n + \sum_{m=1}^M\sum_{n=1}^N
V(x_m-y_n),
\end{equation}
where
\((p_m=-i\partial/\partial x_m, x_m)\) and \((P_n=-i\partial/\partial
y_n, y_n)\) are the momenta and positions associated with the right (a
total of $M$) and left (a total of $N$) movers. The wave functions on
which $H_L$ operates must be totally antisymmetric because of Fermi
statistics. Luttinger's result amounts to the statement that \(H_L\) is
dual to
\begin{equation}
H^D_L=\sum_{m=1}^M\ p_m -\sum_{n=1}^N\ P_n.
\end{equation}
To prove this result from a bond-algebraic perspective, we introduce
the bonds
\begin{equation}
A_m= p_m+\frac{1}{2}\sum_{n=1}^N\ V(x_m-y_n),
\ \ \ \ \ \ B_n=P_n-\frac{1}{2} \sum_{m=1}^M V(x_m-y_n),
\end{equation}
so that
\begin{eqnarray}
H_L=\sum_{m=1}^M\ A_m - \sum_{n=1}^N B_n.
\end{eqnarray}
It is immediate that
\begin{equation}
[A_m, A_{m'}]=0,\ \ \ \ \ \ \ [B_n, B_{n'}]=0, \ \ \ \ \ \ \ \mbox{and }
\ \ [A_m, B_n]=0,
\end{equation}
since
\begin{equation}
[A_{m}, B_{n}]= - \frac{1}{2}\left[p_{m}, \sum_{m'=1}^M\
V(x_{m'}-y_{n})\right]
+\frac{1}{2}\left[\sum_{n'=1}^N\ V(x_{m}-y_{n'}), P_{n}\right]=0.
\end{equation}
Thus, putting all the pieces together, we establish the duality
isomorphism
\begin{equation}
A_m\ \stackrel{\Phi_{\sf d}}{\longrightarrow}\ p_m,\ \ \ \ \ \
B_n\ \stackrel{\Phi_{\sf d}}{\longrightarrow} P_n.
\end{equation}
The above demonstration illustrates that Luttinger's assertion holds
for arbitrary interactions \(V(x-y)\) in Equation \eqref{lutt_model}.
\subsection{QED without sources, compact QED, and \({\mathbb{Z}}_p\) gauge theories}
\label{sec6.2}
\subsubsection{QED without sources}\label{noncompactQED}
The quantization of the EM field suffers from well known
complications due to gauge invariance, and very different from the
complications that arise in the quantization of the \(d=1\) free boson.
They are much less of a problem from a bond algebraic perspective, because
in the end, at least in the absence of sources, we do know how to construct
a full operator quantization of the EM field. The resulting vacuum QED is the
starting point to quantum optics \cite{glauber},
and is self-dual under the exchange of the quantum
electric and magnetic field operators.
We start by setting up the version of QED that we are going to work with.
The starting point is the gauge-invariant action for the vector potential,
\begin{equation}
S_{\sf EM}=-\frac{1}{4}\int dx^0d^3x\ (\partial_\mu A_\alpha-\partial_\alpha A_\mu)
(\partial_\nu A_\beta-\partial_\beta A_\nu)\eta^{\mu\nu}\eta^{\alpha\beta}
\end{equation}
(the connection of the vector potential to the electric \(\vec{E}\)
and magnetic \(\vec{B}\) fields was described in Section \ref{sec2.1}).
To proceed with canonical quantization, we need to partially fix the gauge.
If we choose the condition $A_{0}=0$, called the axial gauge, we can
complete the canonical quantization prescription easily. The resulting $d=3$ QFT
reads
\begin{equation}\label{QED_H}
H_{\sf EM}=\int d^3x\ \left ( \frac{1}{2}\ \vec{\Pi}^{2}({\bm{x}})\ +\
\frac{1}{2}(\nabla\times \vec{A}({\bm{x}}))^2 \right ),
\end{equation}
together with
\begin{equation}\label{comm_EM}
[A_\mu({\bm{x}}),\ \Pi_\nu({\bm{x}}')]=\ i\ \delta_{\mu,\nu}\ \delta({\bm{x}}-{\bm{x}}'),\ \ \ \ \
\ \mu,\nu=1,2,3,
\end{equation}
and the \(A_\mu({\bm{x}}),\ \Pi_\nu({\bm{x}}')\) can be realized as well defined
operators (i.e., operator-valued distributions) action on a Hilbert
state space. Notice that in the axial gauge we are using,
\(\vec{E}=-\partial_t\vec{A}=-\vec{\Pi}\). There is, however, and issue
left from the remaining gauge symmetry of the theory defined by
Equations \eqref{QED_H} and \eqref{comm_EM}. The state space is
larger-than-physical, and only the states \(\vert \Psi \rangle\)
that satisfy the Gauss constraint
\begin{equation}\label{gauss_constraint}
\nabla\cdot\vec{\Pi}\ \ \vert \Psi \rangle=0,
\end{equation}
can be prepared and observed by experimental means. The reason is
that \(\nabla\cdot\vec{\Pi}\) is the generator of the residual
gauge symmetries that where not fixed by the axial condition
$A_{0}=0$, and so Equation \eqref{gauss_constraint} amounts to
the statement that only gauge invariant states are physical.
Let us consider next the bond algebra characterized by
\begin{equation}\label{bonda_em}
[(\nabla\times \vec{A})_\mu({\bm{x}}),\ \Pi_\nu({\bm{x}}')]=
i\delta_{\mu,\nu}(\nabla\times \delta({\bm{x}}-{\bm{x}}'))_\mu.
\end{equation}
It is easy to check that the mapping
\begin{equation}\label{sd_emcont}
\Pi_\mu({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ (\nabla\times \vec{A})_\mu({\bm{x}}),\ \ \ \ \ \
(\nabla\times \vec{A})_\mu({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\Pi_\mu({\bm{x}}),
\end{equation}
preserves the relations of Equation \eqref{bonda_em}, but there is
a subtlety of interpretation. \(\Phi_{{\sf d}}\) maps \(\nabla\times \vec{A}\),
that is automatically divergenceless, to \(\vec{\Pi}\), that is not.
However, \(\vec{\Pi}\) is divergenceless in the gauge invariant
subspace of physical states, and so \(\Phi_{{\sf d}}\) is a true isomorphism
of bond algebras in that subspace. In the language of Section
\ref{sec3.11}, and for this particular quantization of the EM field,
{\it its self-duality is an emergent one}. In other words, the
EM duality is truly a
duality between the observable physical electric and magnetic fields, and
not a more general property of the vector potential.
We can use the mapping \(\Phi_{{\sf d}}\) of Equation \eqref{sd_emcont} to
compute dual variables on the subspace of physical states,
\begin{eqnarray}
&&\vec{\Pi}({\bm{x}}) \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
\widehat{\vec{\Pi}}({\bm{x}})=\nabla\times\vec{A}({\bm{x}}),\nonumber\\
&&\vec{A}({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
\widehat{\vec{A}}({\bm{x}})=-\frac{1}{4\pi}\nabla\times\int d^3x'\
\frac{\vec{\Pi} ({\bm{x}}')}{\vert {\bm{x}}-{\bm{x}}'\vert}.
\end{eqnarray}
This completes our
discussion of the self-duality of QED in the continuum. The next step
is to check it against the lattice Hamiltonian formalism. This analysis
will facilitate later our discussions of duality in other
Hamiltonian lattice gauge field theories that are closely connected to
QED and to the dynamics of center vortices in QCD \cite{fradkin_susskind, horn}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.55\columnwidth]{./plaquette.eps}
\end{center}
\caption{Gauge field degrees of freedom $A_{({\bm{r}},\nu)}$ ($\nu=1,2,3$) for
$d=3$ non-compact lattice QED are placed on the links of a cubic
lattice (of lattice spacing \(a\)), as indicated by the crosses.
The plaquette variable $\Theta_{({\bm{r}},3)}=A_{({\bm{r}},1)}+A_{({\bm{r}}+{\bm{e_1}},2)}-
A_{({\bm{r}}+{\bm{e_2}},1)}-A_{({\bm{r}},2)}$
resides in the plaquette indicated in the $({\bm{e_1}},{\bm{e_2}})$ plane. Other
symbols' meaning are defined in Figure \ref{notation_links}.}
\label{plaquette}
\end{figure}
If we discretize the EM field as we discretized the free boson field near the
end of Section \ref{triv_boson_sub}, we get
\begin{equation}
H_{\sf LEM}=\sum_{\bm{r}}\ \left ( \frac{1}{2a^3}\ \vec{\Pi}_{\bm{r}}^2\ +\
\frac{1}{2}\ a\vec{\Theta}_{\bm{r}}^2 \right ),
\end{equation}
where $\vec{\Pi}_{\bm{r}}=(\Pi_{({\bm{r}},1)}, \Pi_{({\bm{r}},2)},\Pi_{({\bm{r}},3)})$, and
$\vec{\Theta}_{\bm{r}}$ stands for the discretized curl,
\begin{eqnarray}
\label{theta_fields}
\Theta_{({\bm{r}},1)}&=&A_{({\bm{r}},2)}+A_{({\bm{r}}+{\bm{e_2}},3)}-A_{({\bm{r}}+{\bm{e_3}},2)}-A_{({\bm{r}},3)},
\nonumber\\
\Theta_{({\bm{r}},2)}&=&A_{({\bm{r}},3)}+A_{({\bm{r}}+{\bm{e_3}},1)}-A_{({\bm{r}}+{\bm{e_1}},3)}-A_{({\bm{r}},1)},
\label{discrete_curl}\\
\Theta_{({\bm{r}},3)}&=&A_{({\bm{r}},1)}+A_{({\bm{r}}+{\bm{e_1}},2)}-A_{({\bm{r}}+{\bm{e_2}},1)}-A_{({\bm{r}},2)}
\nonumber
\end{eqnarray}
(see Figure \ref{plaquette}). The fact that the theory in the continuum
features a vector field \(\vec{A}\) is reflected in that the lattice
degrees of freedom reside {\it on links}, with commutation relations
\begin{eqnarray}
[A_{({\bm{r}},\mu)}\ ,\ \Pi_{({\bm{r}}',\nu)}]=\ i\ \delta_{\mu,\nu}\
\delta_{{\bm{r}},{{\bm{r}}'}}.
\label{AP_comm}
\end{eqnarray}
Discretizing the theory spoils its space-time symmetries, but gauge
symmetries remain almost unchanged. The generators of gauge symmetries are
\begin{equation}
g_{\bm{r}}=\sum_{\nu=1}^{3}\ (\Pi_{({\bm{r}},\nu)}-\Pi_{({\bm{r}}-\bm{e_\nu},\nu)}),\ \ \
\ \ \ \ \ \ \ [g_{\bm{r}}, H_{\sf LEM}]=0,
\end{equation}
and physical states are characterized as before as \(g_{\bm{r}}|\Psi\rangle=0, \
\forall {\bm{r}}\).
\begin{figure}[b]
\begin{center}
\includegraphics[width=0.55\columnwidth]{./pi_to_delta}
\end{center}
\caption{The effect of the exchange duality $\Phi_{\sf d}$ of Equation
\eqref{EM_duality_auto} on the three $\Pi$ fields at site ${\bm{r}}$. The
directions $x, y, z$ are associated to the unit vectors ${\bm{e_1}},{\bm{e_2}},{\bm{e_3}}$,
respectively. Notice that the $\Pi$ fields, although associated to the
vertex ${\bm{r}}$, reside on the links of the lattice.}
\label{pi_to_delta}
\end{figure}
\(H_{\sf LEM}\) is self-dual, just as its counterpart in the continuum.
The mapping
\begin{eqnarray}
\label{EM_duality_auto}
\Pi_{({\bm{r}},1)}&\stackrel{\Phi_{{\sf d}}}{\longrightarrow}&\ \Theta_{({\bm{r}},1)},\ \
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \Theta_{({\bm{r}},1)}
\stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\Pi_{({\bm{r}}-{\bm{e_1}}+{\bm{e_2}}+{\bm{e_3}},1)},\nonumber\\
\Pi_{({\bm{r}},2)}&\stackrel{\Phi_{{\sf d}}}{\longrightarrow}& \
\Theta_{({\bm{r}}-{\bm{e_1}}+{\bm{e_2}},2)}, \ \ \ \ \ \ \ \ \ \ \, \Theta_{({\bm{r}},2)}
\stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\Pi_{({\bm{r}}+{\bm{e_3}},2)},\nonumber\\
\Pi_{({\bm{r}},3)}& \stackrel{\Phi_{{\sf d}}}{\longrightarrow}& \
\Theta_{({\bm{r}}-{\bm{e_1}}+{\bm{e_3}},3)},\ \ \ \ \ \ \ \ \ \ \, \Theta_{({\bm{r}},3)}
\stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\Pi_{({\bm{r}}+{\bm{e_2}},3)}.
\end{eqnarray}
defines a self-duality isomorphism {\it in the subspace of physical states},
that exchanges the lattice electric \(\vec{\Pi}_{\bm{r}}\) and magnetic \(\vec{\Theta}_{\bm{r}}\)
operators, see Figures \ref{pi_to_delta} and \ref{delta_to_pi}.
\(\Phi_{{\sf d}}\) is not well defined outside the subspace of physical states,
because as it stands in Equation \eqref{EM_duality_auto}, it can be shown to be a
many-valued mapping, \(\mathbb{1}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ g_{\bm{r}},\ \forall {\bm{r}}\). It is easy to check
that the lattice self-duality converges exactly to its counterpart in the continuum
in the naive limit \(a\rightarrow 0\).
\begin{figure}[h]
\begin{center}
\includegraphics[width=.6\columnwidth]{./delta_to_pi.eps}
\end{center}
\caption{The effect of the exchange duality $\Phi_{\sf d}$ of Equation
\eqref{EM_duality_auto} on the three $\Theta$ fields at site ${\bm{r}}$. The
directions $x, y, z$ are associated to the unit vectors ${\bm{e_1}},{\bm{e_2}},{\bm{e_3}}$,
respectively. Notice that the $\Pi$ fields associated to the indicated
vertices reside on the links of the lattice, while the $\Theta$ fields
reside on the corresponding plaquettes. Each colored plaquette map to
the $-\Pi$ at the correspondingly colored site.}
\label{delta_to_pi}
\end{figure}
It is remarkable that, in the end, the
EM duality has the same origin as the self-duality of the
Ising model: a symmetry of a bond algebra.
\subsubsection{Compact QED and \({\mathbb{Z}}_p\) gauge theories}
The self-dual lattice rendition of QED that we studied in the previous section
is non-standard. In
contrast, the standard Hamiltonian lattice field theory for QED
\cite{kogut,kogut_susskind} (see, for example, Section 6 of
\cite{kogut}),
\begin{equation}
\label{lem}
H_{\sf CEM}=\sum_{\bm{r}}\sum_{\nu=1}^3\ \left (
\frac{1}{2}L_{({\bm{r}},\nu)}^2-\lambda \cos \Theta_{({\bm{r}},\nu)} \right ),
\end{equation}
arises from the quantization \cite{kogut} of {\it compact} (lattice) QED as
defined by Wilson \cite{wilson}. Here, the plaquette term
$\Theta_{({\bm{r}},\nu)}$ is formally defined as in Equation \eqref{theta_fields}
up to the replacement $A_{({\bm{r}},\nu)} \rightarrow \theta_{({\bm{r}},\nu)}$,
and the latter satisfy the commutation relations of Equation
\eqref{CQEDcomr}.
The Hamiltonians \(H_{\sf LEM}\) and \(H_{\sf CEM}\) exhibit radically
different phase diagrams, simply because \(H_{\sf LEM}\) describes a system of
harmonic oscillators, while \(H_{\sf CEM}\) features plane rotors in interaction.
In particular, \(H_{\sf CEM}\) {\it is not self-dual}. On the other hand,
one can use the techniques of Sections \ref{sec4.3} and \ref{sec5.4} to set up
a dual gauge model in terms of integer valued degrees of freedom,
\begin{equation}
H_{\sf CEM}^D=\sum_{\bm{r}}\sum_{\nu=1}^3\ \left (
\frac{1}{2}b_{({\bm{r}},\nu)}^2-\frac{\lambda}{2}
(R_{({\bm{r}},\nu)}+R_{({\bm{r}},\nu)}^\dagger)\right )
\end{equation}
(the plaquette term
$b_{({\bm{r}},\nu)}$ is formally defined as in Equation \eqref{theta_fields}
up to the replacement $A_{({\bm{r}},\nu)} \rightarrow X_{({\bm{r}},\nu)}$, the operators
\(X,\ R,\ R^\dagger\) where introduced in Section \ref{sec4.3}). A close classical
relative of this duality to integer valued fields
was found of great use in the latest comprehensive
Monte-Carlo simulation of \(H_{\sf CEM}\) \cite{panero}.
We can also use the mathematics introduced in Sections \ref{sec4.1} and
\ref{sec4.3} to write down a self-dual \(p\)-state approximation
to \(H_{\sf CEM}\)
\begin{equation}\label{N_state_gauge_H}
H_{\sf G}= \frac{1}{2}\sum_{\bm{r}} \sum_{\nu=1}^{3}\ (
V_{({\bm{r}},\nu)}\ +\ \lambda B_{({\bm{r}},\nu)}+\ {\sf h.c.}),
\end{equation}
that features generators \(V_{({\bm{r}},\nu)}\) and \(U_{({\bm{r}},\nu)}\)
of Weyl's group at each link, and
\begin{eqnarray}
B_{({\bm{r}},1)}&=&U_{({\bm{r}},2)}U_{({\bm{r}}+{\bm{e_2}},3)}U^\dagger_{({\bm{r}}+{\bm{e_3}},2)}
U_{({\bm{r}},3)}^\dagger , \nonumber\\
B_{({\bm{r}},2)}&=&U_{({\bm{r}},3)}U_{({\bm{r}}+{\bm{e_3}},1)}U_{({\bm{r}}+{\bm{e_1}},3)}^\dagger
U_{({\bm{r}},1)}^\dagger , \label{3d_Uplaquettes}\\
B_{({\bm{r}},3)}&=&U_{({\bm{r}},1)}U_{({\bm{r}}+{\bm{e_1}},2)}U_{({\bm{r}}+{\bm{e_2}},1)}^\dagger
U_{({\bm{r}},2)}^\dagger. \nonumber
\end{eqnarray}
The self-duality mapping for this model was discussed in
Reference \cite{con}, and is given by
\begin{eqnarray}
V_{({\bm{r}},1)}&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ &B_{({\bm{r}},1)},\ \ \ \ \ \ \ \ \
\ \ \ \ \ \ \ \ \ B_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
V^\dagger_{({\bm{r}}-{\bm{e_1}}+{\bm{e_2}}+{\bm{e_3}},1)},\nonumber\\
V_{({\bm{r}},2)}&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & \ B_{({\bm{r}}-{\bm{e_1}}+{\bm{e_2}},2)}, \ \
\ \ \ \ \ \ \ \, B_{({\bm{r}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ V^\dagger_{({\bm{r}}+{\bm{e_3}},2)},\label{map_gauge_zp}\\
V_{({\bm{r}},3)}&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & \ B_{({\bm{r}}-{\bm{e_1}}+{\bm{e_3}},3)},\ \
\ \ \ \ \ \ \ \, B_{({\bm{r}},3)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ V^\dagger_{({\bm{r}}+{\bm{e_2}},3)}.\nonumber
\end{eqnarray}
(and similarly for Hermitian conjugate bonds).
Similar to the VP model
of Section \ref{sec4.1}, the squared duality isomorphism \(\Phi_{{\sf d}}^2\)
is a non-trivial, discrete charge conjugation symmetry \(\mathcal{C}\)
of the model,
that maps \(V_{({\bm{r}},\nu)}\) and \(U_{({\bm{r}},\nu)}\) to
\(V_{({\bm{r}},\nu)}^\dagger\) and \(U_{({\bm{r}},\nu)}^\dagger\), up to a
translation by \(\bm{d}= -{\bm{e_1}}+{\bm{e_2}}+{\bm{e_3}}\). When \(p\rightarrow\infty\),
the phase structure of \(H_{\sf G}\) approaches that of \(H_{\sf LEM}\)
\cite{frohlich_gauge}, even though \(H_{\sf LEM}\) is not self-dual
(see Section \ref{sec4.3}).
In this section, we introduced \(H_{\sf G}\) in connection to QED.
Actually this model has been intensively studied in the literature
\cite{yoneya,green,horn} in connection to confinement in QCD. This is so
as \(H_{\sf G}\) for \(p\) states affords a simple effective theory to
study center vortices in QCD with \(p\) flavors. This was discussed, to
some extent, by 't Hooft \cite{thooft} where the relevance of center
vortices to confinement was first elucidated, but the specific Hamiltonian
\(H_{\sf G}\) was proposed in Reference \cite{horn}. {}From this point of view,
however, this lattice model suffers from its inability to incorporate
magnetic monopoles. Other aspects of \(H_{\sf G}\) will be discussed in
Section \ref{classical_zpgauge}.
\subsection{QED without a vector potential}
\label{sec6.3}
In classical physics, the EM vector potential $A_\mu$
is a technical advantage, but otherwise expendable, essentially because
the interaction of classical charged particles with the EM field can be
described purely in terms of \(\vec{E}\) and
\(\vec{B}\). In contrast, the vector potential is unavoidable
at the quantum level, and
the best illustrations of this fact is the Aharonov-Bohm effect
\cite{aharonov_bohm} that is non-local in \(\vec{E}\) and
\(\vec{B}\). But even disregarding the interaction to
charges, it is difficult to set up the quantum mechanics of the EM field
alone without introducing a vector potential, though Mandelstam
\cite{mandelstam} managed to put forward a consistent scheme.
The standard quantizations of the EM field, however, rely on
the unobservable $A_\mu$, and suffer from well-known difficulties \cite{das},
that depend on the gauge fixing condition of choice. In the axial gauge
of the previous section, the state space that emerges is a Hilbert space,
but it is redundant, due to the presence of gauge symmetries inherited
upon quantization.
On the other hand, as explained in Section \ref{sec3.12}, we can
use bond-algebraic dualities to find a dual representation that features
no gauge redundancies.
In this section, we illustrate these ideas for compact QED, because it is
the model of greatest relevance in numerical studies of QED.
Ideally, we would like to find a gauge reducing duality for the
Hamiltonian \(H_{\sf LEM}\) of Equation \eqref{lem} to a model
that features local bonds. We saw in Section \ref{sec5.4}
that this is possible in \(d=2\), but it does not seem to be
possible in \(d=3\). Hence we are left with the systematic approach
of Section \ref{sec3.12.4}, that is known to introduce non-local bonds
in the dual model. The starting point is to recognize the generators
of gauge symmetries, in this case
\begin{equation}\label{gtbk}
g_{\bm{r}}=\sum_{\nu=1}^3\ (L_{({\bm{r}},\nu)}-L_{({\bm{r}}-\bm{e_\nu},\nu)}).
\end{equation}
Now we can proceed with the general techniques described in
Section \ref{sec3.12.4}. The gauge-reducing duality should satisfy
\begin{equation}
\Phi_{{\sf d}}(g_{\bm{r}})=0=\sum_{\nu=1}^3\ (\Phi_{{\sf d}}(L_{({\bm{r}},\nu)})-
\Phi_{{\sf d}}(L_{({\bm{r}}-\bm{e_\nu},\nu)})),
\end{equation}
so that in the completely gauge-reduced dual the bonds \(\Phi_{{\sf d}}(L_{({\bm{r}},3)})\)
{\it are not independent}, but we can write instead
\(
\Phi_{{\sf d}}(L_{({\bm{r}},3)})=\ \sum_{n\geq1}\
(\Phi_{{\sf d}}(L_{({\bm{r}}+n{\bm{e_3}},1)})+\Phi_{{\sf d}}(L_{({\bm{r}}+n{\bm{e_3}},2)}) -\Phi_{{\sf d}}(L_{({\bm{r}}+n{\bm{e_3}}-{\bm{e_1}},1)})
-\Phi_{{\sf d}}(L_{({\bm{r}}+n{\bm{e_3}}-{\bm{e_2}},2)}))
\). {}From this point on, the reasoning follows through just as in Section
\ref{sec3.12.4}. The gauge-reducing duality homomorphism reads
\begin{eqnarray}
&&L_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ L_{({\bm{r}},1)}, \ \ \ L_{({\bm{r}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ L_{({\bm{r}},2)},\ \ \
L_{({\bm{r}},3)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ {\cal E}_{({\bm{r}},3)}, \nonumber\\
&&e^{i\Theta_{({\bm{r}},1)}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ e^{i(\theta_{({\bm{r}}+{\bm{e_3}},2)}-\theta_{({\bm{r}},2)})},\ \
\ e^{i\Theta_{({\bm{r}},2)}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
e^{i(\theta_{({\bm{r}}+{\bm{e_3}},1)}-\theta_{({\bm{r}},1)})},\\
&&e^{i\Theta_{({\bm{r}},3)}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ e^{i\Theta_{({\bm{r}},3)}}\nonumber
\end{eqnarray}
(the plaquette angle \(\Theta_{({\bm{r}},\nu)}\) was defined right
after Equation \eqref{lem}). The electric string operator \({\cal E}_{({\bm{r}},3)}\) is defined as
\begin{equation}
{\cal E}_{({\bm{r}},3)} \equiv
\sum_{n\geq1} (L_{({\bm{r}}+n{\bm{e_3}},1)}+L_{({\bm{r}}+n{\bm{e_3}},2)} -
L_{({\bm{r}}+n{\bm{e_3}}-{\bm{e_1}},1)}-L_{({\bm{r}}+n{\bm{e_3}}-{\bm{e_2}},2)}),
\end{equation}
and carries the full weight of the non-locality that seems to be unavoidable
in \(d=3\), if gauge constraints are to be eliminated. The completely gauge-reduced
dual Hamiltonian reads
\begin{eqnarray}
&&H_{\sf GRCEM}=H^0+\sum_{{\bm{r}}}\ \left({\cal E}_{({\bm{r}},3)}-\lambda \cos\Theta_{({\bm{r}},3)}
\right),\\
&&H^0=\sum_{\bm{r}} \sum_{\nu=1,2}\ \left(\frac{1}{2}L_{({\bm{r}},\nu)}-\lambda
\cos(\theta_{({\bm{r}}+{\bm{e_3}},\nu)}-\theta_{({\bm{r}},\nu)})\right).\nonumber
\end{eqnarray}
It is remarkable that \(H^0\) describes a stalk of non-interacting (independent)
\(d=1\), XY models. The idea that the physics of gauge fields in \(D=4\) is
closely connected to the physics of spin models in \(D=2\) has been put
forward many times over the years (see \cite{fradkin_susskind}, and references
therein). The duality just presented is a new indication/realization of this
connection. Moreover, it afford us {\it a theory of QED
without a vector potential}, and consequently, without gauge symmetries
or unwanted non-physical states. Stated differently, \(H_{\sf GREM}\) is
the quantum rendition of Maxwell's dynamics purely in terms of the
electric and magnetic fields. The cost to be paid is the introduction
of non-local bonds.
{}From a practical point of view, it is important to notice that this
gauge-reducing duality can be
restricted to finite systems, and the advantage for numerical
simulations of the gauge reduced rendition of the theory could be
enormous. For one thing, since there are no degrees of freedom associated with the
links along the \({\bm{e_3}}\) direction, the Hilbert space of \(H_{\sf
GRCEM}\) reduces to
\begin{equation}
\mathcal{H}_{\sf GRCEM}=\bigotimes_{\bm{r}}\ \left({\cal L}^2(S^1)_{({\bm{r}},1)}\otimes
{\cal L}^2(S^1)_{({\bm{r}},2)}\right).
\end{equation}
that is much smaller than the one for \(H_{\sf LEM}\) (\({\cal L}^2(S^1)\) denotes
the Hilbert space of square integrable functions on the circle). Physically speaking,
\(\mathcal{H}_{\sf GRCEM}\) exhibits two ``states of polarization" per lattice site.
\subsection{Abelian and $\mathbb{Z}_p$ Higgs models}
\label{sec6.4}
In this section we study the duality properties of the Abelian Higgs (AH)
QFT {\it in its broken symmetry state}.
The AH model is complex
enough that an operator treatment in the continuum is not well
defined, so we proceed directly to the lattice Hamiltonian formalism.
We uncover a new duality in \(d=3\) to a {\it local}, completely
gauge-reduced model, introduce new \(p>2\)-state approximations to the lattice
AH model that we call \(\mathbb{Z}_p\) Higgs models,
and discuss their (self-)dual properties in \(d=2\) and \(d=3\)
(the \(p=2\) case \cite{fradkin_susskind}
is of importance to the theory of topological quantum order and storage of quantum
information, see Section \ref{sec5.5}).
We start with some general comments to put the AH model in perspective.
Both in condensed matter and high energy physics, the success of
QFTs in describing interactions
hinges to a large extent on the principle of
gauge invariance and the Higgs mechanism. The reason for the
latter is that, if the gauge group is compact, gauge invariance requires
gauge fields to be {\it massless}, restricting
in principle their applicability to the
description of {\it long-range} interactions. The Higgs mechanism affords
a way out of this restriction, since it is a
process by which gauge fields acquire mass through the spontaneous
breakdown of a continuous symmetry with no Goldstone bosons. In this way,
gauge fields become capable of describing {\it short-range} interactions as
well, at the expense of introducing a {\it Higgs field}.
The AH model,
\begin{equation}\label{AH}
S_{\sf AH}= \int d^4x\ \left ( -\frac{1}{4}F_{\mu \nu}F^{\mu \nu}+
\frac{1}{2}(D_\mu\phi)(D^\mu\phi)^*+
\lambda(\phi\phi^*-v^2)^2 \right),
\end{equation}
features a complex scalar Higgs field $\phi$ of charge \(q\), in
interaction with the EM field ($F_{\mu\nu}=\partial_\mu A_\nu - \partial_\nu A_\mu$
is the EM field, and \(D_\mu\phi\equiv(\partial_\mu-iqA_\mu)\phi\)), and it
is the simplest field theory that
combines both the principle of gauge invariance and the Higgs
mechanism. When $v^2$ is positive,
the (classical) potential energy is minimized by setting $\phi = v e^{i
\theta}$, and the ground state breaks
(spontaneously) the \(U(1)\) symmetry. The reader familiar with
superconductivity will recognize the resulting action
as the starting point for the phenomenological
Ginzburg-Landau theory of superconductivity (wherein \(\phi\) represents
the superconducting order parameter). In the light of QFT, however, the
spontaneous breakdown translates into
a particle spectrum containing one {\it massive photon}
and one {\it massive, real} scalar (see, for instance, \cite{das}).
This is the Higgs mechanism of mass generation: the would be Goldstone
boson associated to the spontaneous symmetry breakdown is reabsorbed as
an extra degree of freedom of the gauge field $A_\mu$. It is {\it the} mechanism
of mass generation in the standard model of particle physics, but it
is not the only possible one, as we will see in the next section when we
study the St\"uckelberg model.
The lattice Hamiltonian for the AH model in its broken symmetry phase reads
\begin{eqnarray}
H_{\sf LAH}=&&\sum_{\bm{r}}\sum_{\nu=1}^3 (\ \frac{1}{2}L_{({\bm{r}},\nu)}^2
- \lambda\ \cos(\Theta_{({\bm{r}},\nu)})) \nonumber \\
+&& \sum_{\bm{r}} (\ \frac{1}{2}L_{{\bm{r}}}^2-\sum_{\nu=1}^3\
\kappa\ \cos(\theta_{{\bm{r}}+\bm{e_\nu}}-\theta_{\bm{r}}-q\theta_{({\bm{r}},\nu)})),
\label{hlah}
\end{eqnarray}
with the notation of Section \ref{sec6.3} for $L_{({\bm{r}},\nu)}$ and
$\Theta_{({\bm{r}},\nu)}$. The Higgs, or matter, field is represented
by degrees of freedom $L_{\bm{r}}$ and
$\theta_{{\bm{r}}}$ on the sites of a cubic lattice, while the gauge
field is represented by degrees of freedom $\theta_{({\bm{r}}, \nu)}$,
$L_{({\bm{r}},\nu)}$ on its links. \(H_{\sf LAH}\) follows from the classical
lattice action introduced in Reference \cite{fradkin_shenker}, treated
according to the quantization techniques of Reference \cite{kogut_susskind}.
Notice that \(q\) is now constrained to take integer values.
As discussed in Reference
\cite{fradkin_shenker}, the physics (phase diagram) of the
AH model depends strongly upon whether \(q=1\) or has some other value, and
in fact \(q\) is an explicit parameter in the duality mappings to be discussed.
\(H_{\sf LAH}\) defines a gauge theory, with gauge symmetries generated
by
\begin{equation}
g_{\bm{r}} = -qL_{\bm{r}} +\sum_{\nu=1}^3\ (L_{({\bm{r}},\nu)}-L_{({\bm{r}}-\bm{e_\nu},\nu)}).
\end{equation}
It is a remarkable feature that these gauge symmetries are completely
eliminated by a duality to a {\it local} model,
\begin{eqnarray}
H_{\sf GRLAH}= &&\frac{1}{2} \sum_{\bm{r}} \sum_{\nu=1}^3
(X_{({\bm{r}},\nu)}^2-
\lambda\ (B_{({\bm{r}},\nu)}+B_{({\bm{r}},\nu)}^\dagger)) \nonumber \\
+&&\frac{1}{2} \sum_{\bm{r}}\ (\frac{1}{q^2}A_{\bm{r}}^2-
\sum_{\nu=1}^3 \kappa\ q(R_{({\bm{r}},\nu)}+R_{({\bm{r}},\nu)}^\dagger)) ,
\end{eqnarray}
that features integer-valued degrees of freedom and local bonds.
The plaquette operators \(B_{({\bm{r}},\nu)}\) are defined just as
in \eqref{3d_Uplaquettes}, up to the replacement \(U_{({\bm{r}},\nu)}
\rightarrow R_{({\bm{r}},\nu)}\), and \(
A_{\bm{r}}=\sum_{\nu=1}^3\ (X_{({\bm{r}},\nu)}-X_{({\bm{r}}-\bm{e_\nu},\nu)})\)
(the operators \(R\), \(R^\dagger\) and \(X\) were introduced in
Section \ref{sec5.2}). It follows that \(H_{\sf GRLAH}\) does not
feature any degrees of freedom on the {\it sites} of the lattice.
The duality homomorphism reads
\begin{eqnarray}
L_{\bm{r}}\ &\stackrel{\Phi_{{\sf d}}}{\longrightarrow}&\ \frac{1}{q}A_{{\bm{r}}},
\ \ \ \ \ \ \ \ \ \ \ \
e^{i(\theta_{{\bm{r}}+\bm{e_\nu}}-\theta_{\bm{r}}-q\theta_{({\bm{r}},\nu)})}
\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ qR_{({\bm{r}},\nu)}, \nonumber\\
L_{({\bm{r}},\nu)}\ &\stackrel{\Phi_{{\sf d}}}{\longrightarrow}&\ X_{({\bm{r}},\nu)},
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \,
e^{-i\Theta_{({\bm{r}},\nu)}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\
B_{({\bm{r}},\nu)},
\end{eqnarray}
and it is straightforward to check the trivialization of the
infinitesimal generators of gauge symmetries,
\begin{equation}
\Phi_{{\sf d}}(g_{\bm{r}})=-A_{\bm{r}}+\sum_{\nu=1}^3\ (\Phi_{{\sf d}}(L_{({\bm{r}},\nu)})
-\Phi_{{\sf d}}(L_{({\bm{r}}-\bm{e_\nu},\nu)})=0.
\end{equation}
\(H_{\sf GRLAH}\) has no local symmetries. Let us point out
without elaborating the details that the
AH model admits a {\it local gauge-reducing duality} along these
lines {\it in any dimension \(d\)}.
We can take a different approach to the study of the AH model,
and write \(p\)-state approximations to the
Hamiltonian of Equation \eqref{hlah}, in terms of generators
of the Weyl group algebra introduced in Sections \ref{sec4.1} (this is
similar in spirit to the approximation of the \(d=1\) XY model by a
VP model). We call these $p$-state
quantum models $\mathbb{Z}_p$ Higgs models, and to our knowledge they
have not been studied before, for \(p\geq3\)
(the $p=2$ case \cite{fradkin_shenker,horn_yankielowicz,lamont}
was discussed in Section \ref{sec5.5} from the perspective of
topological quantum order).
The $\mathbb{Z}_p$ Higgs models admit completely gauge-reducing
dualities, and the dual models that arise are natural generalizations
of the ETC model of Section \ref{sec5.5},
to any number of states $p$ and
any number of dimensions \(d\). Moreover, in \(d=2\), they define a
{\it new class of self-dual models}. Let us focus on this case for
the rest of the section.
The Hamiltonian for the \(d=2\)-dimensional $\mathbb{Z}_p$ Higgs model
reads
\begin{eqnarray}\label{AH_N_H}
H_{\sf pAH}&=&\frac{1}{2} \sum_{\bm{r}}\ ( V_{\bm{r}}-\lambda B_{({\bm{r}},3)}\ +\ {\sf
h.c.}) \\
&+&\frac{1}{2} \sum_{\bm{r}}\sum_{\nu=1,2}\ (V_{({\bm{r}},\nu)}- \kappa\
U^\dagger_{{\bm{r}}}(U^\dagger_{({\bm{r}},\nu)})^qU_{\bm{r+e_\nu}}+\ {\sf
h.c.}),\nonumber
\end{eqnarray}
with gauge symmetries
\begin{equation}
G_{\bm{r}}=(V^\dagger_{\bm{r}})^q V_{({\bm{r}},1)} V_{({\bm{r}},2)}V^\dagger_{({\bm{r}}-{\bm{e_1}},1)}
V^\dagger_{({\bm{r}}-{\bm{e_2}},2)},\ \ \ \ \ \ [G_{\bm{r}},H_{\sf G}]=0,
\end{equation}
and \(G^\dagger_{\bm{r}}\).
As usual, $U_{{\bm{r}}}, V_{{\bm{r}}}$ denote site (vertex) operators, and
$U_{({\bm{r}}, \nu)}, V_{({\bm{r}}, \nu)}$ reside at the link $({\bm{r}},\nu)$.
Notice also that the charge \(q\) is a $\mathbb{Z}_p$ charge now,
and can only take one of the values \(q=0,\cdots,p-1\). The self-duality
mapping reads
\begin{eqnarray}
V_{\bm{r}} &\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & B_{({\bm{r}},3)}, \ \ \ \ \ \ \ \ \ \ \ \ B_{({\bm{r}},3)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ V_{{\bm{r}}+{\bm{e_1}}+{\bm{e_2}}},
\label{zphiggs_sd}\\
U^\dagger_{\bm{r}} (U^\dagger_{({\bm{r}},1)})^q U_{{\bm{r}}+{\bm{e_1}}}&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & V_{({\bm{r}}+{\bm{e_1}},2)},\ \ \ \ \ \ \ \ \
V_{({\bm{r}},1)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ U_{{\bm{r}}+{\bm{e_1}}}U^q_{({\bm{r}}+{\bm{e_1}},2)}U^\dagger_{{\bm{r}}+{\bm{e_1}}+{\bm{e_2}}}
,\nonumber\\
U^\dagger_{\bm{r}} (U^\dagger_{({\bm{r}},2)})^q U_{{\bm{r}}+{\bm{e_2}}}&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ &\ V^\dagger_{({\bm{r}}+{\bm{e_2}},1)},\ \ \ \ \ \ \ \
V_{({\bm{r}},2)}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ U^\dagger_{{\bm{r}}+{\bm{e_2}}}(U^\dagger_{({\bm{r}}+{\bm{e_2}},1)})^qU_{{\bm{r}}+{\bm{e_1}}+{\bm{e_2}}}. \nonumber
\end{eqnarray}
Strictly speaking, \(\Phi_{{\sf d}}\) is an isomorphism only on the restriction
of the bond algebra to the subspace of gauge-invariant states, so this
is another example of an {\it emergent self-duality} (we encountered
a similar situation in \(d=3\) QED, Section \ref{sec6.2}). To see this,
notice that one can expand the identity operator \(\mathbb{1}\) as a product of
bonds in many different ways, and that all these different expansions are mapped to
products of gauge symmetries. It follows that \(\Phi_{{\sf d}}\) is a {\it multivalued}
homomorphism, unless it is restricted to the subspace of gauge-invariant states.
Since \(H_{\sf pAH}\) features both {\it matter and gauge}
fields in interaction, it is interesting to compare the self-duality
of Equation \eqref{zphiggs_sd} to the self-duality properties {\it expected}
of QED in the presence of suitable sources.
In general, it is argued that by introducing
magnetic charges, the self-duality of QED
in the absence of sources, Section \ref{sec6.2},
could be extended to include sources as well. This putative
self-duality, however, would not mix matter fields with gauge fields. In
contrast, in the Higgs case, the self-duality establishes an
equivalence between matter and gauge fields.
The \(p\)-state models $H_{\sf pAH}$ can be completely gauge-reduced
by dualities that are very similar to that for \(p=2\), worked
out in Section \ref{sec5.5}. There is also a variation of the
\({\mathbb{Z}}_p\) Higgs models where the VP-like interactions are
replaced by P-like interactions \cite{kogut_potts}. This Potts-Higgs model
has the advantage that its
phase diagram can be studied analytically in a \(1/p\) expansion.
\subsection{The self-dual St\"uckelberg model}\label{sec_stuckel}
In this section we discuss the St\"uckelberg model of mass
generation, and show that it is self-dual in \(d=2\) dimensions.
The massless free boson of Section \ref{sec6.1}
\begin{equation}
S_{\sf FB}=\int d^Dx\ \frac{1}{2}\eta^{\mu \nu}
\partial_\mu\phi\partial_\nu\phi
\end{equation}
(now in any dimension \(D=d+1\)) has a global, internal continuous symmetry
of the form \(
\phi({\bm{x}}) \mapsto \phi({\bm{x}})+\alpha\),
\(\alpha\in\mathbb{R}\), but this fact is rarely of interest because
the conservation law that follows is tantamount to the equation of motion.
Things become more interesting if apply the gauge principle
to this symmetry, and gauge it to make it local, at the expense of introducing
a vector potential \(A_\mu\),
\begin{equation}\label{s_action}
S_{\sf S}=\int d^Dx\ \frac{1}{2}\eta^{\mu \nu}
(\partial_\mu\phi-mA_\mu)(\partial_\nu\phi-mA_\nu)-\frac{1}{4}F^{\mu \nu}F_{\mu \nu}.
\end{equation}
This is the {\it St\"uckelberg model of mass generation} (see \cite{das}, and
references therein), proposed by St\"uckelberg in 1958.
The gauge symmetries of \(S_{\sf S}\) are
\begin{equation}
\phi({\bm{x}})\ \mapsto\ \phi({\bm{x}})+\alpha({\bm{x}}),\ \ \ \
A_\mu({\bm{x}})\ \mapsto\ A_\mu({\bm{x}})+\frac{1}{m}\partial_\mu\alpha({\bm{x}}),
\end{equation}
and show that {\it both} the \(\phi\) and \(A_\mu\) field are {\it not
observable}. Combined, however, they describe {\it in a gauge invariant fashion}
a {\it massive} vector field (a Proca
field) of bare mass \(m\) (to see this, just impose the
gauge-fixing condition \(\phi={\sf constant}\) in Equation \eqref{s_action}).
Since the Proca field
can be completely quantized in operator form \cite{takahashi}, we expect
that the same holds for \(S_{\sf S}\), and
we work directly in the continuum. Still, let us point out that analogous
calculations in the lattice Hamiltonian approach
return results that are perfectly compatible with the continuum,
in the naive continuum limit $a \rightarrow 0$.
The St\"uckelberg
model is self-dual only in \(d=2\), as will become clear soon,
so from now on we work out just that case.
The canonical quantization of \(S_{\sf S}\) has the usual complications
coming from gauge-invariance. The simplest way to proceed is to partially
fix the gauge by imposing the axial constraint \(A_0=0\). This allows to put
\(S_{\sf S}\) in canonical form, so that we can apply the standard quantization
procedures to get
\begin{eqnarray}
&&H_{\sf S}=\frac{1}{2}\int d^2x\left ( \vec{\Pi}^2+\
\left(\partial_1 A_2-\partial_2 A_1\right)^2+\ \pi^2+\ \left(
\nabla\phi-m\vec{A}\right)^2 \right ),\\
&&[\phi({\bm{x}}),\ \pi({\bm{x}}')]=\ i\ \delta({\bm{x}}-{\bm{x}}'),\ \ \ \ \ \
[A_\mu({\bm{x}}),\ \Pi_\nu({\bm{x}}')]=\ i\ \delta_{\mu,\nu}\ \delta({\bm{x}}-{\bm{x}}'),\nonumber
\end{eqnarray}
with \(\mu,\nu=1,2\), and every other commutator set equal to zero.
The subspace of physical, gauge-invariant states is characterized by
\begin{equation}
(-\frac{1}{m}\nabla\cdot\vec{\Pi}+\pi)|\Psi\rangle=0.
\end{equation}
That \(H_{\sf S}\) is self-dual follows from the mapping
\begin{eqnarray}
\pi({\bm{x}})&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & \left(\partial_1 A_2-\partial_2 A_1\right)({\bm{x}}),\ \
\left(\partial_1 A_2-\partial_2 A_1\right)({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \pi({\bm{x}}), \nonumber
\\
\Pi_1({\bm{x}})&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & -(\partial_2\phi-mA_2)({\bm{x}}),\ \ \ \ \
(\partial_2\phi-mA_2)({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\Pi_1({\bm{x}}),
\\
\Pi_2({\bm{x}})&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & (\partial_1\phi-mA_1)({\bm{x}}),\ \ \ \ \ \ \
(\partial_1\phi-mA_1)({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \Pi_2({\bm{x}}).\nonumber
\end{eqnarray}
Since we can write \(0=-\partial_1(\partial_2\phi-mA_2)+\partial_2
(\partial_1\phi-mA_1)-m(\partial_1A_2-\partial_2A_1)\) (among other
possibilities), we have that
\begin{eqnarray}
\Phi_{{\sf d}}(0)=\nabla\cdot\vec{\Pi}-m\pi.
\end{eqnarray}
This means that just as in QED, Section \ref{noncompactQED},
\(\Phi_{{\sf d}}\) represents a self-duality isomorphism only
in the sector of gauge-invariant states, and is multi-valued
in the full, gauge-redundant state space.
This completes our discussion of the St\"uckelberg model, but let us
point out in closing that all of the previous results follows just as easily
in the lattice Hamiltonian approach.
\subsection{Field theory and dimensional reduction}
\label{6.8}
There is an intimate connection between $d$-dimensional systems
possessing $\bar{d}$-dimensional gauge-like symmetries \cite{tqo},
and the phenomenon of {\it dimensional reduction}, where the physical
system in $d$ dimensions behaves
in many ways as if it had effectively a smaller $\bar{d}<
d$ number of dimensions \cite{tqo,3detc}. Mathematically, this
connection results from establishing bounds for the correlation
functions of the $d$-dimensional theory in terms of another theory in
$\bar{d}$ dimensions. A very broad and exciting field where
dimensional reductions, $\bar{d}$-dimensional gauge-like symmetries,
and dualities come to the fore is that of topological quantum order
\cite{tqo,3detc}. In topologically ordered systems, the state of the
system cannot be characterized by local measurements but rather by
topological quantities.
In this paper, we have considered the duality properties of several
lattice models that display topological quantum order, including the XM and
POC models of Sections \ref{sec4.4} and \ref{sec5.1} \cite{tqo},
and the paradigmatic ETC model of Section \ref{sec5.5}.
Here we develop continuum (field-theoretic)
versions of those lattice models, where dimensional reduction occurs
because of the existence of $\bar{d}$-dimensional gauge-like
symmetries. Consider then the non-relativistic, $d=2$-dimensional QFT
\begin{equation}\label{bosonicCXM}
H_{\sf P}=\frac{1}{2}\int d^2x\ ( \pi^2+ \lambda
\left(\partial_1\partial_2\phi\right)^2),\ \ \ \ \ \
[\phi({\bm{x}}),\ \pi({\bm{x}}')]=i\delta({\bm{x}}-{\bm{x}}').
\end{equation}
By construction, this model is invariant under the \(\bar{d}=1\) gauge-like symmetry
\(\phi(x)\rightarrow \phi(x)+\alpha(x^1)+\beta(x^2)\), where \(\alpha,\ \beta\) are
smooth, real functions of {\it one} variable.
Also, it is self-dual, as follows from the mapping
\begin{eqnarray}\label{dual_bosonicXM}
\!\!\!\!\!\!\!\!\!\!\!\!\!\!
\phi({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\int_{-\infty}^{x^1}\int_{x^2}^{\infty} d^2x'\ \ \pi({\bm{x}}'),
\ \ \ \ \ \ \
\pi({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\partial_1\partial_2\phi({\bm{x}})\ ,
\end{eqnarray}
that defines the dual variables of the problem. Notice that formally,
\(g_i(\alpha)\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ 0\). This is the standard manifestation (seen many times
in formally infinite lattice models, see the discussion in Section \ref{sec3.7})
of the fact that to make \(\Phi_{{\sf d}}\) rigorous we need to specify
self-dual BCs.
In what follows we are going to study these results
in some detail on the lattice, to showcase the connection of
the model of Equation \eqref{bosonicCXM} to other lattice
models with gauge-like symmetries. The lattice Hamiltonian
approach applied to Equation \eqref{bosonicCXM} returns
\begin{equation}\label{bosonicXM}
H_{\sf PL}=\frac{1}{2a^2}\sum_{\bm{r}}\ (\pi^{2}_{\bm{r}} +\ \lambda (\square
\phi_{\bm{r}})^2 ), \ \ \ \ \ \ [\phi_{\bm{r}},\ \pi_{{\bm{r}}'}]=i\delta_{{\bm{r}},{\bm{r}}'},
\end{equation}
with $\square \phi_{\bm{r}} \equiv \phi_{\bm{r}}-\phi_{\bm{r+e_2}}
+\phi_{\bm{r-e_1+e_2}}-\phi_{\bm{r-e_1}}$, that is self-dual by virtue of
\begin{equation}
\pi_{\bm{r}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\square\phi_{\bm{r-e_1+e_2}}, \ \ \ \ \ \
\square {\bm{x}}_{\bm{r}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \pi_{\bm{r-e_1+e_2}}.
\end{equation}
The structure of \(H_{\sf PL}\) suggests setting up a \(p\)-state model
of the form
\begin{equation}
H_{\sf pPL}=\frac{1}{2}\sum_{\bm{r}}\ \left (V_{\bm{r}}+\ \lambda\square U_{\bm{r}}
+\ {\sf h.c.} \right),
\end{equation}
with $\square U_{\bm{r}}=U_{\bm{r}}
U^\dagger_{\bm{r+e_2}}U_{\bm{r-e_1+e_2}}U^\dagger_{\bm{r-e_1}}$
(the \(U\) and \(V\) operators where introduced in Section \ref{sec4.4}).
For \(p\geq3\), \(H_{\sf pPL}\) defines a {\it class
of self-dual models that has not been studied before} to the best of
our knowledge. For \(p=2\), \(H_{\sf pPL}\) becomes
identical to \(H_{\sf XM}\), the XM model of Section \ref{sec4.4}.
The connection between \(H_{\sf P}\), \(H_{\sf PL}\), and
\(H_{\sf pPL}\) stands on their common self-dual structure {\it and the
shared presence of \(\bar{d}=1\) gauge-like symmetries} (see Section
\ref{sec4.4} for a discussion of the gauge-like symmetries of the XM
model). Thus these models
afford an excellent scenario to study
the role of dimensional reduction and topological quantum order in
more general settings, where the structure of the elementary degrees
of freedom are varied in a controlled fashion.
Since \(H_{\sf PL}\) shares some formal similarities with the XM model,
we expect it to show a duality to a model analogous to the POC
model of Section \ref{sec5.2}. The dual model turns out to be
\begin{equation}\label{BPOC}
H_{\sf PL}^D=\frac{1}{2a^2}\sum_{{\bm{r}}}\ \left((\pi_{{\bm{r}}+{\bm{e_1}}}-\pi_{\bm{r}})^2+\lambda
(\phi_{{\bm{r}}+{\bm{e_2}}}-\phi_{\bm{r}})^2\right),
\end{equation}
as follows from the mapping
\begin{equation}
\label{bpoc2bxm}
\pi_{{\bm{r}}+{\bm{e_1}}}-\pi_{\bm{r}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ -\square \phi_{\bm{r+e_1}},
\ \ \ \ \ \ \ \ \ \
\phi_{{\bm{r}}+{\bm{e_2}}}-\phi_{\bm{r}}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \pi_{{\bm{r}}+{\bm{e_2}}}.
\end{equation}
Interestingly, it is easy to take the naive continuum limit \(a\rightarrow 0\)
of \(H_{\sf PL}^D\),
\begin{equation}
H_{\sf P}^D= \frac{1}{2}\int d^2x\ ((\partial_1 \pi)^2+\lambda
(\partial_2 \phi)^2),\ \ \ \ \ \ [\phi({\bm{x}}),\ \pi({\bm{x}}')]=i\delta({\bm{x}}-{\bm{x}}').
\end{equation}
That \(H_{\sf P}^D\) is indeed dual to the QFT defined in Equation
\eqref{bosonicCXM} follows from the isomorphism
\begin{equation}
\partial_1 \pi({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \partial_1\partial_2\phi({\bm{x}}),
\ \ \ \ \ \ \ \ \ \ \
\partial_2\phi({\bm{x}})\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \pi({\bm{x}}).
\end{equation}
\section{Bond-algebraic approach to classical dualities}
\label{sec8}
In this section we establish dualities for models of
classical statistical mechanics, exploiting the theory of
Section \ref{classical&quantum}, and introduce a new
classical gauge \(\mathbb{Z}_p\) model that is self-dual for any \(p\).
The aim is to illustrate the unification of the theory of dualities
in the framework of bond algebras, and the advantages of approaching classical
dualities from this new perspective. On the other hand, it is important to
keep in sight the related fact that quantum dualities are fundamentally related to
classical ones by path integrals or the
the STL decomposition (see Section \ref{classical&quantum}).
These widely used techniques approximates the exponential of a sum of
(non-commuting) operators by products of exponentials
\begin{eqnarray}
e^{-(H_{1} + H_{2})/N} = e^{-H_{1}/N} e^{-H_{2}/N}
+ O(1/N^2),
\label{explain_suzuki}
\end{eqnarray}
with an error assumed to be bounded \cite{schulman}, and, combined
with bond-algebraic techniques, they can afford a simple fast way
to connect quantum dualities to classical dualities \cite{con,bookEPTCP}.
In the following, however, we will bypass the use of the STL
decomposition technique, to obtain exact duality mappings
between finite or infinite models of classical statistical physics.
\subsection{The classical Ising model in the Utiyama lattice}
\label{sec8.2}
As mentioned in Section \ref{sec2.1}, one of the most celebrated
self-dualities discovered long ago by Kramers and Wannier
enabled a quantitative prediction \cite{KW} of the
critical temperature of the classical Ising model in a square lattice.
{}From a bond-algebraic perspective this self duality is a consequence of
the self-duality of the quantum Ising chain of Section \ref{sec3.7}.
Since this calculation is already
available in textbook form \cite{bookEPTCP}, we
present in this section the most general case of a duality for the
classical Ising model
in the Utiyama lattice \cite{utiyama} defined in Figure \ref{2Dutiyama}.
The partition function that describes Utiyama's anisotropic, bipartite
$D=2$ classical Ising model
\begin{eqnarray}
&&\ \ \ \mathcal{Z}_{\sf U}(K_1,K_2,K_3,K_4)=\\
&&\sum_{\{\sigma_{\bm{r}}\}}\ \exp \left[\sum_{{\bm{r}}\in {\sf even}}\ (K_4\sigma_{{\bm{r}}+{\bm{e_1}}}\sigma_{\bm{r}}+
K_1\sigma_{{\bm{r}}+{\bm{e_2}}}\sigma_{\bm{r}})+
\sum_{{\bm{r}}\in {\sf odd}}\ (K_2\sigma_{{\bm{r}}+{\bm{e_1}}}\sigma_{\bm{r}}+
K_3\sigma_{{\bm{r}}+{\bm{e_2}}}\sigma_{\bm{r}})\right] \ ,\nonumber
\end{eqnarray}
features four different nearest-neighbor couplings, $K_\mu$, arranged as
shown in Figure \ref{2Dutiyama} (a point \({\bm{r}}\) is {\sf even}
if \(r^1+r^2={\sf even}\), and {\sf odd} otherwise).
The advantage in studying \(\mathcal{Z}_{\sf U}(K_1,K_2,K_3,K_4)\) is that
it describes in a unified fashion several important renditions of the \(D=2\)
Ising model. In particular,
\begin{itemize}
\item
$K_1=K_3$ and $K_2=K_4$ corresponds to the Ising model on a square lattice,
\begin{equation}\label{isquare}
\mathcal{Z}_{\sf I}(K_1,K_2)=\mathcal{Z}_{\sf U}(K_1,K_2,K_1,K_2).
\end{equation}
\item
$K_4 \rightarrow \infty$ corresponds to the Ising model on a triangular lattice,
\begin{equation}\label{itr}
\mathcal{Z}_{\sf IT}(K_1,K_2,K_3)=\lim_{K_4 \rightarrow \infty}\mathcal{Z}_{\sf U}(K_1,K_2,K_3,K_4).
\end{equation}
\item
$K_3 \rightarrow 0$ corresponds to the Ising model on an hexagonal lattice,
\begin{equation}\label{ihex}
\mathcal{Z}_{\sf IH}(K_1,K_2,K_4)=\lim_{K_3 \rightarrow 0}\mathcal{Z}_{\sf U}(K_1,K_2,K_3,K_4).
\end{equation}
\end{itemize}
The Ising model
on a triangular lattice is dual to the Ising model on an hexagonal lattice,
and the Ising model on a square lattice is self-dual. As will be shown next, both
of these results can be determined at once from a bond-algebraic analysis
of Utiyama's partition function.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.40\columnwidth]{./2Dutiyama.eps}
\end{center}
\caption{Utiyama's version of the $D=2$-dimensional (classical)
Ising model features four different coupling constants \(K_1,\cdots,K_4\)
distributed in checkerboard fashion. Sometimes this arrangement
is referred to as the Ising model in the Utiyama lattice.}
\label{2Dutiyama}
\end{figure}
We start by recasting
\(\mathcal{Z}_{\sf U}(K_1,K_2,K_3,K_4)\) for a $2M\times 2N$ lattice in terms
of a transfer matrix, $T$, to later relate it to a quantum Hamiltonian
problem.
The transfer matrix elements, most convenient to our purposes, read
\begin{eqnarray}\label{transfer_uti}
&& \langle \sigma'|T(K_1,K_2,K_3,K_4)|\sigma\rangle=\exp\Big [K_2\sigma_{2M}^{\;}+
K_4\sigma_{2M-1}\sigma_{2M}\\
&+&\sum_{i=1}^{M-1} \left(K_4\sigma_{2i-1}\sigma_{2i}+K_2\sigma_{2i}\sigma_{2i+1} \right )
+\sum_{i=1}^{M} \left (K_1\sigma_{2i-1}'\sigma_{2i-1}+
K_3\sigma_{2i}'\sigma_{2i}\right) \Big ],\nonumber
\end{eqnarray}
where \(2M\) is the number of sites along the horizontal direction subject to
open BCs. The Ising spin variable $\sigma_j=\pm 1$
belongs to the horizontal row, while \(\sigma_{j}'\)
denotes the spin immediately above \(\sigma_{j}\) (that
is, on the next horizontal line). Notice also that we have introduced a boundary
term \(K_2\sigma_{2M}\), inconsequential in the thermodynamic limit, but
that will turn out to be essential to define classical self-dual BCs.
If we impose periodic BCs in the
vertical direction (so that any column contains \(2N\) sites), we can write
\begin{equation}\label{pf_uti}
\tilde{\mathcal{Z}}_{\sf U}(K_1,K_2,K_3,K_4)={{\sf Tr}\ } \left[
T(K_1,K_2,K_3,K_4)T(K_3,K_4,K_1,K_2)\right]^N
\end{equation}
for the Utiyama-Ising model with self-dual BCs. The
next step is to write \(T\) using techniques similar to those of Reference \cite{schultz}
\begin{eqnarray}
&& \frac{T(K_1,K_2,K_3,K_4)}{\left(4\sinh(2K_1)
\sinh(2K_3)\right)^{M/2}}=e^{-H^1[h_1,h_3]}e^{-H^0[K_4,K_2]},
\end{eqnarray}
where
\begin{eqnarray}\label{hs_uti}
H^1[h_1,h_3]&=&-\sum_{i=1}^{M}\ (h_1\sigma^x_{2i-1}+h_3\sigma^x_{2i}),
\\
H^0[K_4,K_2]&=&-K_2\sigma^z_{2M}-K_4\sigma^z_{2M-1}\sigma^z_{2M}-\sum_{i=1}^{M-1}\
(K_4\sigma^z_{2i-1}
\sigma^z_{2i}+K_2\sigma^z_{2i}\sigma^z_{2i+1}) \nonumber ,
\end{eqnarray}
with \(h_\nu=-\frac{1}{2}\ln\tanh K_\nu\), \(\ \nu=1,3\).
At this point one can apply the bond-algebraic results of
Section \ref{sec3.7} to show that \(H^1\)
is dual to \(H^0\),
\begin{equation}
H^1[h_1,h_3]=\mathcal{U}_{{\sf d}}^\dagger H^0[h_3,h_1]\mathcal{U}_{{\sf d}},\ \ \ \
H^0[K_4,K_2]=\mathcal{U}_{{\sf d}}^\dagger H^1[K_2,K_4]\mathcal{U}_{{\sf d}},
\end{equation}
and, together with the cyclic property of the trace, this implies that
\begin{eqnarray}\label{sd_utiyama_ising}
\frac{\tilde{\mathcal{Z}}_{\sf U}(K_1,K_2,K_3,K_4)}
{\left(4\sinh(2K_1)
\sinh(2K_3)\right)^{MN}}&=&{{\sf Tr}\ } \left[e^{-H^1[h_1,h_3]}e^{-H^0[K_4,K_2]}
e^{-H^1[h_3,h_1]}e^{-H^0[K_2,K_4]}\right]^N \nonumber\\
&=&{{\sf Tr}\ } \left[e^{-H^1[K_4,K_2]}e^{-H^0[h_3,h_1]}e^{-H^1[K_2,K_4]}
e^{-H^0[h_1,h_3]}\right]^N \nonumber\\
&=&\frac{\tilde{\mathcal{Z}}_{\sf U}(K_1^*,K_2^*,K_3^*,K_4^*)}
{\left(4\sinh(2K_1^*)
\sinh(2K_3^*)\right)^{MN}},
\end{eqnarray}
demonstrating an {\it exact} self-dual mapping of the Utiyama-Ising model
for {\it any}
finite lattice.
The dual couplings follow from comparing the first and second lines
of Equation \eqref{sd_utiyama_ising}
\begin{eqnarray}\label{utiyama_dual_couplings}
h_1^*&\equiv&-\frac{1}{2}\ln\tanh K_1^*=K_4,\ \ \ \
h_3^*\equiv-\frac{1}{2}\ln\tanh K_3^*=K_2,\\
K_4^*&=&h_3\equiv-\frac{1}{2}\ln\tanh K_3,\ \ \ \
K_2^*=h_1\equiv-\frac{1}{2}\ln\tanh K_1.\nonumber
\end{eqnarray}
This completes the bond-algebraic study of the self-duality properties
of the Utiyama-Ising lattice model \(\tilde{\mathcal{Z}}_{\sf U}(K_1,K_2,K_3,K_4)\).
Equations \eqref{sd_utiyama_ising} and \eqref{utiyama_dual_couplings},
together with Equations \eqref{isquare}, \eqref{itr}, and \eqref{ihex} afford a simple
derivation of the self-duality relation for the square lattice Ising model first derived by
Kramer and Wannier \cite{KW},
\begin{equation}\label{is}
\mathcal{Z}_{\sf I}(K_1,K_2)=A (K_1,K_1^*) \, \mathcal{Z}_{\sf I}(K_1^*,K_2^*) ,
\end{equation}
and of the duality relation between the hexagonal and triangular lattices referred to
by Onsager \cite{onsager} and written down by Wannier \cite{wannier}
\begin{equation}
\mathcal{Z}_{\sf IH}(K_1,K_2,K_4)=A(K_1,K_1^*,K_3^*) \, \mathcal{Z}_{\sf IT}(K_1^*,K_2^*,K_3^*)
\end{equation}
(with analytic functions $A$, see Section \ref{classical&quantum}).
The last duality follows from
the fact that if \(K_3\rightarrow 0\), then \(K_4^*\rightarrow \infty\).
It is important to stress that we have derived the self-duality of
Equation \eqref{sd_utiyama_ising}, using bond algebras, for any finite
or infinite lattice. One could also derive this self-duality by starting from an
appropriate quantum Ising chain and use the STL decomposition
(see Reference \cite{bookEPTCP}). However, in this case it is necessary to
perform the thermodynamic limit in the extra dimension such that
the conditions of the Trotter theorem are satisfied \cite{schulman}.
\subsection{The classical vector Potts model}
\label{sec8.3}
The \(D=2\) VP model introduced
in Section \ref{sec4.1.1} \cite{taroni, bookEPTCP}
has interesting but hard to uncover duality properties. Bond-algebraic methods
provide a powerful approach to
unveil those duality properties
that rely on the bond-algebraic isomorphism
of Section \ref{sec4.1.1} and the results from Appendix \ref{appF}.
The transfer matrix of the VP model in an $M\times N$ lattice, with partition function written
in Equation \eqref{classicalVP}, is given by ($\theta_i=2\pi s_i/p$, $s_i=0,1,\cdots,p-1$)
\begin{eqnarray}
&&\langle s'| {T}_{\sf VP}(K_x,K_y)|s\rangle\equiv\\
&&\exp\left[\sum_{i=1}^M\ K_y\cos(\theta_{i}'-\theta_i)+\sum_{i=1}^{M-1}\
K_x\cos(\theta_{i+1}-\theta_i)+ K_x\cos(\theta_M)\right],\nonumber
\end{eqnarray}
where \(M\) is the number of sites along the horizontal
direction. The last term represents
a classical self-dual BC that will allow us exploit the
exact bond-algebraic isomorphism discussed in Section \ref{sec4.1.1}.
Thus, in the following we will assume open BCs along the
horizontal direction and periodic BCs along the vertical direction.
If we further
assume that the states \(|s\rangle\equiv\bigotimes_{i=1}^M |s_i\rangle\), \(s_i=
0,\cdots, p-1\) represent the basis diagonalizing the Weyl group algebra matrices \(U_i\)
of Equation \eqref{VsandUs}, then we can write
\begin{eqnarray}
{T}_{\sf VP}(K_x,K_y)=e^{-H^0[K_x]}\ \prod_{i=1}^M\ e^{-H^1_i[K_y]},
\label{transfermaVPD2}
\end{eqnarray}
where
\begin{eqnarray}
H^0[K_x]=-\frac{K_x}{2}\sum_{i=1}^{M-1} (U_{i+1}^\dagger U_i+ U_{i}^\dagger U_{i+1})
-\frac{K_x}{2}\left(U_M+U_M^\dagger\right),
\label{transfer_VPH0}
\end{eqnarray}
and $H^1_i[K_y]$ is an operator whose matrix elements are given by
\begin{eqnarray}
\langle s_i'|e^{-H^1_i[K_y]}|s_i\rangle=e^{K_y\cos(\theta_{i}'-\theta_i)}.
\label{transfer_VP}
\end{eqnarray}
To determine \(H^1_i[K_y]\), we rewrite Equation \eqref{transfer_VP} in matrix form
\begin{equation}
e^{-H^1_i[K_y]}=\sum_{m=0}^{p-1}\ e^{K_y \cos \theta_m}\ V_i^m ,
\end{equation}
and apply the results of Appendix \ref{appF} to show that \(H^1_i[K_y]\) must
be of the form
\begin{equation}\label{non_diag_VP}
H^1_i[K_y]=-\sum_{m=0}^{p-1}\ h_m(K_y)\ V_i^m\ ,
\end{equation}
(recall that \((V^{m})^\dagger=V^{p-m}\) and \((U^{m})^\dagger=U^{p-m}\)) with
\begin{equation}\label{manyhs}
h_m(K_y)=\frac{1}{p}\sum_{s=0}^{p-1}\ \cos(m \theta_s) \ln\left(\sum_{l=0}^{p-1}
\ e^{K_y\cos \theta_l}\cos(l \theta_s) \right)\ .
\end{equation}
This completes the factoring of the transfer matrix \({T}_{\sf VP}(K_x,K_y)\)
of Equation \eqref{transfermaVPD2}.
The isomorphism defined in Equations \eqref{aut_finite_VP1} and
\eqref{aut_finite_VP2} determines dual forms
for \(H^0[K_x]\) and \(H^1[K_y]=\sum_i H_i^1[K_y]\),
\begin{eqnarray}
H^{0}[K_x]&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ &H^{0 D}[K_x]= -\frac{K_x}{2}\sum_{i=1}^{M}\ (V_i+V_i^\dagger),\\
H^{1}[K_y]&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ &H^{1 D}[K_y]= -\sum_{i=1}^{M-1} \sum_{m=0}^{p-1}\ h_m(K_y)\ (U^\dagger_{i+1}
U_i)^m-\sum_{m=0}^{p-1}\ h_m(K_y)U_M^m\nonumber,
\end{eqnarray}
that translates into an {\it exact} classical duality for the VP model. The
{\it exact} classical dual reads \(\widetilde{\mathcal{Z}}^D_{\sf VP}={{\sf Tr}\ }\left(e^{-H^{0 D}[K_x]}
e^{-H^{1 D}[K_y]}\right)^N\), and can be written down longhand
with the help of the results of Appendix \ref{appF}. The important point
to notice is that the Boltzmann weight of the dual model has the general
structure
\begin{equation}
e^{\sum_{m=0}^{p-1}\ K^*_{\nu m}\cos (m \theta'-m \theta)},\ \ \ \
\nu=x,y,
\end{equation}
with \(s',\ s\) representing the states of a pair of nearest neighbors, and
\(K^*_{x m}\) a function of \(K_y\) alone, while
\(K^*_{y m}\) a function of \(K_x\) alone.
Clearly, the VP model {\it is not self-dual} for arbitrary \(p\) and
arbitrary couplings. However, the model is approximately self-dual
in the extreme anisotropic limit with \(K_y \gg K_x\) and it is
exactly self-dual for \(p=2,3,4\). We study these aspects of the
VP model in the next two sections.
\subsubsection{Approximate self-duality in the extreme anisotropic limit}
As explained in Appendix \ref{appF}, in the limit in which \(K_y\)
becomes extremely large \(K_y\rightarrow\infty\), Equation \eqref{non_diag_VP}
simplifies to
\begin{equation}
H^1_i[K_y]\approx K_y+\frac{\lambda}{2}(V_i+V^\dagger_i),
\ \ \ \ \ \ \frac{\lambda}{2}=e^{K_y(\cos\frac{2\pi}{p} -1)}
\end{equation}
(see Equation \eqref{plarge_K}), so that
\begin{eqnarray}
\widetilde{\mathcal{Z}}_{\sf VP}(K_x,K_y)&\approx& e^{MNK_y}{{\sf Tr}\ }\left[e^{\frac{\lambda}{2}\sum_i
(V_i+V^\dagger_i)}e^{-H^0[K_x]}\right]^N\nonumber\\
&\approx&e^{MNK_y}{{\sf Tr}\ }\left[ e^{-H^0[\lambda]} e^{\frac{K_x}{2}\sum_i
(V_i+V^\dagger_i)}\right]^N \\
&\approx&e^{MN(K_y-K_y^*)}\widetilde{\mathcal{Z}}_{\sf VP}(K_x^*,K_y^*), \nonumber
\end{eqnarray}
with dual couplings
\begin{equation}
\lambda^*\equiv 2e^{K_y^*(\cos\frac{2\pi}{p} -1)}=K_x,\ \ \ \
K_x^*=\lambda \equiv 2 e^{K_y(\cos\frac{2\pi}{p} -1)}.
\end{equation}
We emphasize that this approximate self-duality, in the extreme anisotropic limit,
is valid for {\it any} value of $p$. We next consider {\it exact} self-dualities for the
particular cases $p=2,3,4$.
\subsubsection{The particular cases \(p=2, 3\), and $4$}\label{p234}
Let us start with the simplest $p=2$ Ising case.
If \(p=2\), $U=U^\dagger=\sigma^z$ and \(V=V^\dagger=\sigma^x\). Then, \({T}_{\sf VP}(K_x,K_y)=
e^{-H^0[K_x]}e^{-H^1[K_y]}\)
can be written in terms of
\begin{eqnarray}
H^0[K_x]&=&-K_x\sum_{i=1}^{M-1}\sigma^z_i\sigma^z_{i+1}-K_x\sigma^z_M , \\
H^1[K_y]&=&-\frac{M}{2}\ln(2\sinh (2K_y))+\frac{1}{2}\ln\tanh K_y \sum_{i=1}^M\ \sigma^x_i,
\end{eqnarray}
which is simply the anisotropic Ising model on a rectangular $M\times N$ lattice, studied
in Section \ref{sec8.2}.
If \(p=3\), then \(V^2=V^\dagger\), and $H^1[K_y]$ becomes
\begin{equation}
H^1[K_y]=-M h_{0,3}
-\frac{1}{3}\ln\left[\frac{e^{K_y}+2e^{-\frac{1}{2}{K_y}}}{e^{K_y}-e^{-\frac{1}{2}
{K_y}}}\right]\sum_{i=1}^M
\ (V_i+V_i^\dagger),
\end{equation}
with \(h_{0,3}=\frac{1}{3}\ln\left[(e^{K_y}+2e^{-\frac{1}{2}{K_y}})
(e^{K_y}-e^{-\frac{1}{2}{K_y}})^2\right]\), where $h_{m,p}=h_m$ of Equation \eqref{manyhs}
for a particular $p$.
\(H^0[K_x]\) is identical to Equation \eqref{transfer_VPH0}, with \(U\)s appropriate for
\(p=3\). It follows that \(H^1\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ H^0\) and \(H^0\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ H^1\), rendering
\(\widetilde{\mathcal{Z}}_{\sf VP}\) self-dual. The classical dual couplings
follow as usual from comparing
the transfer matrices of the original and dual models
\begin{equation}
K_x^*=\frac{2}{3}\ln\left[\frac{e^{K_y}+2e^{-\frac{1}{2}{K_y}}}{e^{K_y}-e^{-\frac{1}{2}
{K_y}}}\right],
\ \ \ \ \ \ K_x=\frac{2}{3}\ln\left[\frac{e^{K^*_y}+2e^{-\frac{1}{2}{K^*_y}}}{e^{K^*_y}
-e^{-\frac{1}{2}{K^*_y}}}\right] .
\end{equation}
Finally, consider the case \(p=4\). The general structure of \(H^1[K_y]\) is
\begin{equation}
H^1[K_y]=-M h_{0,4}-h_{1,4}\sum_{i=1}^M\ (V_i+V_i^\dagger) - h_{2,4}\sum_{i=1}^M\ V_i^2,
\end{equation}
but, as a matter of fact, \(h_{2,4}\) vanishes. A simple application of
Equation \eqref{manyhs} shows that
\begin{equation}
h_{2,4}=\frac{1}{4}\ln\left[\frac{(2+e^{K_y} +e^{-K_y})(-2+e^{K_y}
+e^{-K_y})}{(e^{K_y} -e^{-K_y})^2}\right]=0.
\end{equation}
Also, \(h_{1,4}=-(1/2)\ln\tanh(K_y/2)\).
It follows that \(\widetilde{\mathcal{Z}}_{\sf VP}\) is again self-dual, with dual couplings
\begin{equation}
\frac{K^*_x}{2}=-\frac{1}{2}\ln\tanh\frac{K_y}{2}, \ \ \ \ \ \
\frac{K_x}{2}=-\frac{1}{2}\ln\tanh\frac{K_y^*}{2} ,
\end{equation}
which are equivalent to the ones for the Ising model. This is not
surprising since the \(p=4\) VP model is equivalent to two decoupled classical Ising models
\cite{savit}.
\subsection{The classical eight-vertex model}
\label{eightvertexM}
In this section we use bond algebras to study some aspects
of the eight-vertex (8V) model on a square lattice
\cite{sutherland,fan_wu}.
The 8V has an almost self-evident self-duality,
and a duality that clarifies to some extent its
connection to the anisotropic $d=1$ quantum Heisenberg model.
The 8V model is among the most thoroughly studied exactly solvable models
in statistical mechanics \cite{baxter}, largely because it shows
non-universal critical exponents. The first breakthrough in the
study of the 8V model
came with Sutherland's observation \cite{sutherland}
that its transfer matrix commutes with the quantum Heisenberg
Hamiltonian, provided both models'
couplings are suitably adjusted. It is natural then to ask
whether there is a connection stronger than integrability.
Libero and Drugowich de Felicio \cite{defelicio}
have shown that the STL (path integral) representation
of (a model dual to) the anisotropic Heisenberg model is dual to the
8V model on suitable regions of those models coupling spaces. The problem
with this remarkable connection is that (strictly speaking) it only
holds true provided some couplings are
infinitesimally small, and others infinitely large
(typical of the STL representation of quantum models \cite{suzuki}).
In contrast, our results in this section are exact, and reproduce
those of Reference \cite{defelicio} for appropriately chosen couplings.
The transfer matrix used by Sutherland \cite{sutherland} in his original
work is not convenient for our purposes.
To associate a simple transfer matrix to the 8V model
we consider its Ashkin-Teller (AT) representation \cite{wu},
\begin{eqnarray}
\mathcal{Z}_{\sf 8V/AT}=\sum_{\{\mu_{\bm{r}},\tau_{\bm{r}}\}}\exp\Big[\sum_{\bm{r}}\sum_{\nu=1,2}
&&(K_{0,\nu}+K_{1,\nu}\ \mu_{\bm{r}}\mu_{{\bm{r}}+\bm{e_\nu}}\\
&&+K_{2,\nu}\ \tau_{\bm{r}}
\tau_{{\bm{r}}+\bm{e_\nu}}+K_{4,\nu}\ \mu_{\bm{r}}\mu_{{\bm{r}}+\bm{e_\nu}}\tau_{\bm{r}}
\tau_{{\bm{r}}+\bm{e_\nu}})
\Big],\nonumber
\end{eqnarray}
that features two independent classical Ising variables
\(\mu_{\bm{r}}, \tau_{\bm{r}}=\pm1\) at each site of a square lattice of size \(M\times N\).
The relations connecting the couplings \(K_{i,\nu}\) to the parameters
of the 8V model can be found in Reference \cite{defelicio}.
The additive constants \(K_{0,1}, K_{0,2}\) are irrelevant to what follows,
so we ignore them, but they can easily be reintroduced if needed. Then
we can write as usual \(\mathcal{Z}_{\sf 8V/AT}={{\sf Tr}\ }(T_1T_0)^N\), with
\begin{eqnarray}
T_0&=&\prod_{i=1}^{M-1}e^{K_{1,1}\ \mu^z_i\mu^z_{i+1}}e^{K_{2,1} \tau^z_i\tau^z_{i+1}}
e^{ K_{4,1}\ \mu^z_i\mu^z_{i+1}\tau^z_i\tau^z_{i+1}},\label{transferAT}\\
T_1&=&\prod_{i=1}^M(e^{K_{1,2}}+e^{-K_{1,2}}\ \mu^x_i)(e^{K_{2,2}}+e^{-K_{2,2}}\ \tau^x_i)
(e^{K_{4,2}}+e^{-K_{4,2}}\ \mu^x_i\tau^x_i).
\end{eqnarray}
Since the \(\mu\)-spins are independent of (commute with) the \(\tau\)-spins,
we see that \(\mathcal{A}_{\sf 8V/AT}=\mathcal{A}_{\sf I}\otimes\mathcal{A}_{\sf I}\),
where \(\mathcal{A}_{\sf I}\) is the bond algebra of the quantum Ising chain studied
in Section \ref{sec3.4}. It follows that we can apply the self-duality mapping
of the Ising chain to the \(\mu\)-bonds while leaving the \(\tau\)-bonds fixed
(or {\it viceversa}), or we can apply the mapping simultaneously to both types of bonds.
The first case defines a duality that we will not
study here any further. The second one defines a self-duality.
The 8V model is connected through integrability \cite{sutherland}
to the anisotropic Heisenberg model, so we would like to see if we can find
a representation of \(\mathcal{A}_{\sf 8V/AT}\) in terms of the bonds of the Heisenberg
model. A simple bond-algebraic analysis reveals one such representation:
\begin{eqnarray}
\mu^x_i&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ &\sigma^x_{2i-1}\sigma^x_{2i},\ \ \ \ \ \ \ \ \
\tau^x_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^z_{2i-1}\sigma^z_{2i}, \ \ \ \ \ \
i=1,\cdots,M, \label{atellertoh}\\
\mu^z_i\mu^z_{i+1}&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & \sigma^z_{2i}\sigma^z_{2i+1},\ \ \ \
\tau^z_{i}\tau^z_{i+1}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{2i}\sigma^x_{2i+1},\ \ \ \ \ \
i=1,\cdots,M-1.\nonumber
\end{eqnarray}
This duality mapping \(\Phi_{{\sf d}}\) is illustrated in Figure \ref{attohfig}.
\begin{figure}
\centering
\includegraphics[angle=0, width=.8\columnwidth]{./attoh_scaled.eps}
\caption{The duality mapping \(\Phi_{{\sf d}}\) of Equation \eqref{atellertoh},
for \(M=3\) sites. The top and bottom chains represent the two types of independent
bonds in the AT model, the middle chain represents the independent
bonds in the Heisenberg model (the bond \(\sigma^y_i\sigma^y_{i+1}=
-\sigma^x_i\sigma^x_{i+1}\sigma^z_i\sigma^z_{i+1}\) is not independent).
Notice that both models act on state spaces of the same
dimensionality \(2^{2M}\).}.
\label{attohfig}
\end{figure}
It follows that the transfer matrix of Equation \eqref{transferAT}
admits the dual representation
\begin{eqnarray}\label{heisenbergtransfer}
T_0^D&=&\prod_{i=1}^{M-1}e^{K_{1,1}\ \sigma^z_{2i}\sigma^z_{2i+1}}\
e^{K_{2,1}\ \sigma^x_{2i}\sigma^x_{2i+1}}\
e^{-K_{4,1}\ \sigma^y_{2i}\sigma^y_{2i+1}},\\
T_1^D&=&\prod_{i=1}^M(e^{K_{1,2}}+e^{-K_{1,2}}\ \sigma^x_{2i-1}\sigma^x_{2i})
(e^{K_{2,2}}+e^{-K_{2,2}}\ \sigma^z_{2i-1}\sigma^z_{2i})
(e^{K_{4,2}}-e^{-K_{4,2}}\ \sigma^y_{2i-1}\sigma^y_{2i}) \nonumber \\
&=&C^M\prod_{i=1}^Me^{K_{1,2}^*\ \sigma^x_{2i-1}\sigma^x_{2i}}\
e^{K_{2,2}^*\ \sigma^z_{2i-1}\sigma^z_{2i}}\
e^{-K_{4,2}^*\sigma^y_{2i-1}\sigma^y_{2i}} , \nonumber
\end{eqnarray}
where
\begin{equation}
\sinh(2K_{i,2}) \sinh(2K_{i,2}^*)=1,\ \ \ \ \ \ \ \ i=1,2,4,
\end{equation}
and \(C^2=1/(8\sinh(2K_{1,2}^*)\sinh(2K_{2,2}^*)\sinh(2K_{4,2}^*))\), see Equation
\eqref{dualt0dising}.
We can re-write Equation \eqref{heisenbergtransfer} as
\begin{equation}
T_0^D\ =\ e^{H_{\sf even}},\ \ \ \ \ \ \ \ T_1^D\
=\ C^M\ e^{H_{\sf odd}},
\end{equation}
where
\begin{equation}
H_{\sf even}=\sum_{i=1}^{M-1}\sum_{\nu=1}^3\ J_\nu^{\sf e}\ \sigma^\nu_{2i}\sigma^\nu_{2i+1},
\ \ \ \ \ \ \ \
H_{\sf odd}=\sum_{i=1}^{M}\sum_{\nu=1}^3\ J_\nu^{\sf o}\ \sigma^\nu_{2i-1}\sigma^\nu_{2i},
\end{equation}
with couplings
\(J^{\sf e}_1=K_{2,1},\ J_2^{\sf e}=-K_{4,1},\
J_3^{\sf e}=K_{1,1}\), and \(J^{\sf o}_1=K_{1,2}^*,\
J_2^{\sf o}=-K_{4,2}^*,\ J_3^{\sf o}=K_{2,2}^*\).
Next we can use the BCH formula to write
\begin{equation}\label{h8vbch}
T_1^DT_0^D\ =\ C^M\ e^{H_{\sf odd}+H_{\sf even}+[H_{\sf odd},\ H_{\sf even}]/2+\ \cdots}.
\end{equation}
The expression \(H_{\sf H}=H_{\sf odd}+ H_{\sf even}\) defines an anisotropic Heisenberg
model in the region of coupling space where \(J_\nu^{\sf e}=J_\nu^{\sf o}\).
Also, the result of Reference \cite{defelicio} follows by choosing the couplings so that
we can neglect \([H_{\sf odd},\ H_{\sf even}]\) and higher-order terms in Equation
\eqref{h8vbch}. This completes our discussion of the classical 8V model in its
AT representation.
In closing, we would like to comment on the bond algebra mapping of Equation
\eqref{atellertoh} from the point of view of {\it quantum} dualities,
specially in connection to the problem of generating STL (path integral)
representations of the Heisenberg model. The STL representation of
quantum lattice models is the starting point for quantum
Monte Carlo techniques \cite{booksuzuki}, and the fact that
the Heisenberg model does not have a simple STL representation was
noticed already in the early Reference \cite{suzuki} (by simple we mean a local or
quasi-local representation with real and positive Boltzmann weights). In contrast,
the mapping of Equation \eqref{atellertoh} (or more precisely, its
inverse \(\Phi_{{\sf d}}^{-1}\) produces a dual representation
of the Heisenberg model \(H_{\sf H}=
\sum_{i=1}^{2M-1}\sum_{\nu=1}^3\ J_\nu\ \sigma^\nu_{i}\sigma^\nu_{i+1}\),
\begin{eqnarray}
H_{\sf H}^D&=&(J_1\mu^x_M-J_2\mu^x_M\tau^x_M+J_3\tau^x_M)+\\
&&\sum_{j=1}^{M-1}(J_1(\mu^x_j+\tau^z_j\tau^z_{j+1})-J_2(
\mu^x_j\tau^x_j+\mu^z_j\mu^z_{j+1}\tau^z_j\tau^z_{j+1})
+J_3(\tau^x_j+\mu^z_j\mu^z_{j+1})),\nonumber
\end{eqnarray}
that, as shown in Reference \cite{defelicio},
does have a simple STL representation (e.g., the classical AT
model just studied. For this reason the Hamiltonian \(H_{\sf H}^D\) above is
sometimes known in the literature as the quantum AT model \cite{kadanoffat}).
This fact illustrates one of the important applications of our bond-algebraic
approach stressed in Reference \cite{con}, that is, that the method allows
a systematic search for simple STL (or path integral) representations of
quantum problems.
Let us present in closing the dual variables that follow from
Equation \eqref{atellertoh}. After extending the duality
mapping as
\begin{equation}
\Phi_{{\sf d}}(\tau^z_1)=\sigma^x_1,\ \ \ \ \ \ \ \ \ \ \ \
\Phi_{{\sf d}}(\mu^z_M)=\sigma^z_{2M},
\end{equation}
we can compute (\(j=1,\cdots,M\)) the dual variables
\begin{eqnarray}
\hat{\mu}^x_j&=&\sigma^x_{2j-1}\sigma^x_{2j},\ \ \ \ \ \ \ \ \ \ \ \
\hat{\mu}^z_j=\prod_{m=2j}^{2M}\sigma^z_m,\\
\hat{\tau}^x_j&=&\sigma^z_{2j-1}\sigma^z_{2j},\ \ \ \ \ \ \ \ \ \ \ \
\hat{\tau}^z_j=\prod_{m=1}^{2j-1}\sigma^x_{m}.
\end{eqnarray}
where we write \(\hat{\mathcal{O}}\) for the image of \(\mathcal{O}\) under
\(\Phi_{{\sf d}}\), \(\mathcal{O}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \hat{\mathcal{O}}\).
The formal counterparts of these dual variables
for the infinite-size quantum AT model were presented in Reference \cite{shankar}.
\subsection{Classical dualities in \(D=3\) and \(D=4\) dimensions}
\label{sec8.4}
\subsubsection{\(D=3\) Ising/\(\mathbb{Z}_2\) gauge models}
The $D$-dimensional Ising gauge (or \(\mathbb{Z}_2\) gauge) model features Ising
spin variables $\sigma_{({\bm{r}},\mu)}=\pm 1$ residing on the links of a
$N=N_1\times N_2\times \cdots \times N_D$ hyper-cubic lattice.
The model was introduced by Wegner \cite{wegner_ising} as an example
of a system displaying a phase transition without the existence of
a (local) Landau order parameter. Its partition function is given by
\begin{equation}\label{wegner_model}
\mathcal{Z}_{\sf IG}(K)=\sum_{\{\sigma_{({\bm{r}},\mu)}\}}\ \exp\left[K\sum_{\bm{r}}\sum_{\mu>\nu}\
\sigma_{({\bm{r}},\mu)}\sigma_{({\bm{r}}+\bm{e_\mu},\nu)}\sigma_{({\bm{r}}+\bm{e_\nu},\mu)}
\sigma_{({\bm{r}},\nu)}\right] .
\end{equation}
In \(D=3\) dimensions \(\mu,\nu=1,2,3\), and the vertex operator $A_{\bm{r}}$ that flips
the sign, \(\sigma \rightarrow -\sigma\), of all six Ising spins
sharing the vertex ${\bm{r}}$ is a gauge symmetry of the model.
To set up a transfer matrix for \(\mathcal{Z}_{\sf IG}\)(K), one introduces
a condition that partially fixes the gauge invariance. That condition
amounts to consider only those spin configurations that satisfy the
constraint \(\sigma_{({\bm{r}},3)}=1\) at every link \(({\bm{r}},3)\). Since any spin configuration
that does not satisfy this condition can be obtained from a compliant
configuration, it follows that
\begin{equation}
\mathcal{Z}_{\sf IG}(K)=\tilde{N}_{\sf G} \, \mathcal{Z}_{\sf IG}'(K),
\end{equation}
where \(\mathcal{Z}_{\sf IG}'(K)\) is the partition function of Equation \eqref{wegner_model},
with the sum over spin configurations being replaced by the sum over those
configurations that satisfy the constraint, \(\sum_{\{\sigma_{({\bm{r}},\mu)}\}}\rightarrow
\sum_{\{\sigma_{({\bm{r}},\mu)}\}}'\), and $\tilde{N}_{\sf G}$ is a factor that takes into account
the remaining configurations not included in \(\mathcal{Z}_{\sf IG}'(K)\).
As long as we compute only expectation values of
gauge-invariant quantities, we can replace \(\sum_{\{\sigma_{({\bm{r}},\mu)}\}}\rightarrow
\tilde{N}_{\sf G} \, \sum_{\{\sigma_{({\bm{r}},\mu)}\}}'\) everywhere, and the factor
\(\tilde{N}_{\sf G}\) drops out,
which is a manifestation of the fact that the gauge symmetries of the model provide
a redundant description. On the other hand, the partially gauge fixed partition function
\(\mathcal{Z}_{\sf IG}'(K)\) can be written in terms of the transfer matrix
\begin{equation}
\frac{T_{\sf IG}(K)}{(2\sinh(2K))^{N_1N_2/2}}=e^{-H^1[K]}\ e^{-H^0[K]},
\end{equation}
with operators
\begin{equation}
H^1[K]\equiv\frac{1}{2}\ln\tanh K \sum_{\bm{r}}\sum_{\mu=1,2}\ \sigma^x_{({\bm{r}},\mu)},
\ \ \ \ \ H^0[K]=-K\sum_{\bm{r}} B_{({\bm{r}},3)},
\end{equation}
where now \({\bm{r}}=m^1{\bm{e_1}}+m^2{\bm{e_2}}\) denotes the {\it vertices of a $N_1\times N_2$ square lattice},
representing the planes of constant
\(r^3\) in the original cubic lattice, and \(B_{({\bm{r}},3)}\) is the plaquette operator defined
in Equation \eqref{gauge_plaquette}. For ease of presentation, we assume that \(N_1N_2\)
is sufficiently large and thus avoid introducing dual boundary corrections, that could
anyway be computed
by the techniques already developed in previous sections.
The residual gauge invariance
present in \(\mathcal{Z}_{\sf IG}'(K)\) translates into gauge symmetries of
\(H^1[K]\) and \(H^0[K]\) that were discussed at length in Section \ref{sec3.12},
together with their duality properties.
It finally follows that the Ising gauge partition function can be written as
\begin{equation}
\mathcal{Z}_{\sf IG}(K)=N_{\sf G} (2\sinh(2K))^{N/2}{{\sf Tr}\ }\left[(e^{-H^1[K]}
e^{-H^0[K]})^{N_3}P_{\sf GI}\right]
\end{equation}
where \(N_{\sf G}\) counts the total gauge redundancy, and \(P_{\sf GI}\) is
the orthogonal projector onto the space of gauge invariant states. The next
step is to use the projective, gauge reducing duality introduced in Section
\ref{sec3.12} to write (recall that \(P_{\sf GI}=U_{{\sf d}}^\dagger U_{{\sf d}}\) is a projector and
commutes with $H^0[K]$ and $H^1[K]$)
\begin{eqnarray}\label{classical_wegner}
\frac{\mathcal{Z}_{\sf IG}(K)}{N_{\sf G} (2\sinh(2K))^{N/2}}&=&
{{\sf Tr}\ }\left[(U_{{\sf d}} e^{-H^1[K]}U_{{\sf d}}^\dagger\
U_{{\sf d}} e^{-H^0[K]}U_{{\sf d}}^\dagger)^{N_3}\right] \nonumber \\
&=&{{\sf Tr}\ }\left[ (e^{-\frac{1}{2}\ln\tanh K \, \sum_{\bm{r}}\sum_{\mu=1,2}\sigma^z_{\bm{r}}
\sigma^z_{{\bm{r}}+\bm{e_\mu}}}\ e^{K\sum_{\bm{r}} \sigma^x_{\bm{r}}})^{N_3}\right] \nonumber \\
&=&\frac{\mathcal{Z}_{\sf I}(K^*)}
{(2\sinh(2K^*))^{N/2}},
\end{eqnarray}
with \(K^*=-\frac{1}{2}\ln\tanh K\), and \(\mathcal{Z}_{\sf I}\) the partition function
of the \(D=3\) Ising model. This completes the bond-algebraic proof of the
duality between the \(D=3\) Ising and Ising gauge models \cite{wegner_ising}.
The quantum duality underlying this classical duality was extended
to cover $p$-state VP/\(\mathbb{Z}_p\) gauge models in
Section \ref{sec5.3}. We can thus generalize the classical duality of this
section to
$p>2$. If \(p=2,3,\) or \(4\) (\(p=2\) being the case we
just covered in Equation \eqref{classical_wegner}), then the quantum duality
implies that the corresponding \(D=3\) classical VP/\(\mathbb{Z}_p\)
gauge models are dual (the partition function for the \(\mathbb{Z}_p\)
gauge models is presented in the next section, Equation \eqref{class_zpgauge}).
On the other hand, if \(p\geq 5\), then the quantum duality
translates into an {\it exact} classical duality between models that are
modified versions of the \(D=3\) classical VP/\(\mathbb{Z}_p\)
gauge models. These modified classical models can be computed with the
aid of Appendix \ref{appF}, along the lines sketched in Section \ref{sec8.3}.
We next concentrate on the specifics of this generalization but in $D=4$ dimensions.
\subsubsection{A new family of \(D=4\) self-dual \(\mathbb{Z}_p\) gauge theories}
\label{classical_zpgauge}
In this section we study classical \(D=4\) \(\mathbb{Z}_p\) gauge theories
of the Wilson type \cite{yoneya}, and of a new type that has the advantage
of being exactly self-dual for any \(p\). The two theories coincide for \(p=2,3,4\),
corresponding to the cases where the \(\mathbb{Z}_p\) partition function/Euclidean
path integral of the Wilson type is exactly self-dual. The following discussion closely
parallels that of the VP model of Section \ref{sec8.3}, and clarifies
the strong connection between the self-dual structure of both models \cite{con}.
The Wilson-type action of $D$ dimensional \(\mathbb{Z}_p\) gauge theories
is a VP-like generalization of Wegner's Ising ($p=2$)
gauge model of Equation \eqref{wegner_model},
\begin{equation}\label{class_zpgauge}
\mathcal{Z}_{\sf WG}(K)=\sum_{\{\theta_{({\bm{r}},\mu)}\}}\ \exp\left[K\sum_{\bm{r}}\sum_{\mu>\nu}\
\cos \Theta_{({\bm{r}},\mu\nu)} \right],
\end{equation}
where
\begin{eqnarray}
\Theta_{({\bm{r}},\mu\nu)}=
\theta_{({\bm{r}},\mu)}+\theta_{({\bm{r}}+\bm{e_\mu},\nu)}-\theta_{({\bm{r}}+\bm{e_\nu},\mu)}-
\theta_{({\bm{r}},\nu)}, \ \ \ \ \ \mu,\nu=1,\cdots,D ,
\end{eqnarray}
with discrete angles \(\theta_{({\bm{r}},\nu)}=2\pi
s_{({\bm{r}},\nu)}/p\), \(s_{({\bm{r}},\nu)}=0,\cdots,p-1\), placed on the links of
a hyper-cubic lattice with vertices \({\bm{r}}\). {}From now on, we focus on the
\(D=4\) case. The initial interest in \(\mathcal{Z}_{\sf WG}\) \cite{yoneya}
was stimulated by work of 't Hooft on confinement in QCD \cite{thooft}, that stresses
the importance to confinement of the fact that the center of
the gauge group \(SU(p)\) is \(\mathbb{Z}_p\) (\(p\) here stands for the number
of colors). {}From this viewpoint, \(p=3\) is especially important, and we will
show that \(\mathcal{Z}_{\sf WG}\) is self-dual \cite{yoneya}. On the other hand,
for large \(p\), one can think of \(\mathcal{Z}_{\sf WG}\) as an approximation to compact
QED \cite{horn,frohlich_gauge}. In \(D=4\), compact QED is known to show a
phase transition between a confinement
and Coulomb phases \cite{vettorazzo_forcrand, panero}, and it was shown in
Reference \cite{frohlich_gauge} that this feature is shared by \(\mathcal{Z}_{\sf WG}\)
for \(p\) sufficiently large.
To determine the transfer matrix of \(\mathcal{Z}_{\sf WG}\), $T_{\sf WG}$, we
need to partially fix the gauge of the model by considering only configurations
that satisfy the constraint \(\theta_{({\bm{r}},4)}=0\) (we take \(\mu=4\) to
be the Euclidean time direction). Since any other configuration can obtained
from one satisfying this constraint by a gauge transformation, the restriction
has no physical consequence as long as we compute averages of gauge-invariant
observables only. Under these conditions, we can write
\begin{equation}
\mathcal{Z}_{\sf WG}(K)=N_{\sf G}{{\sf Tr}\ } [{T_{\sf WG}(K)}]^{N_4},\ \ \ \ \ \ \ \ \mbox{}
\ \ \ \ T_{\sf WG}(K)= e^{-H^1[K]}e^{-H^0[K]},
\end{equation}
with \(N_{\sf G}\) a counting factor introduced to compensate for the gauge-fixing
condition, and
\begin{equation}
H^1[K]=-\sum_{\bm{r}}\sum_{\nu=1}^3\sum_{m=0}^{p-1}\ h_m(K)V^m_{({\bm{r}},\nu)},
\ \ \ \ \ \ \ H^0[K]= -K\sum_{\bm{r}}\sum_{\nu=1}^3\ B_{({\bm{r}},\nu)}.
\end{equation}
The bonds \(B_{({\bm{r}},\nu)}\) were defined in Equation \eqref{3d_Uplaquettes}, and
the couplings \(h_m(K)\) are {\it identical} to those computed for the
VP model, Equation \eqref{manyhs} (see also Appendix \ref{appF}).
This exact form of the transfer matrix \(T_{\sf WG}\) is,
to the best of our knowledge,
computed here explicitly for the first time for arbitrary \(p\) (In Reference
\cite{yoneya}, Yoneya performs this computation for \(p=2,3,4\), but
is unable to extend his approach to larger \(p\)).
{}From this point on, the analysis of the duality properties of \(\mathcal{Z}_{\sf WG}\) proceeds
just as that for the VP model, except that the appropriate
bond-algebra mapping is the one of Section \ref{sec6.2}, Equation \eqref{map_gauge_zp}.
It follows, just as for the VP model, that \(\mathcal{Z}_{\sf WG}\) is self-dual for
\(p=2,3,4\) \cite{yoneya} (see Section \ref{p234}), and for $p>4$ dual to a
different \({\mathbb{Z}}_p\) gauge model with Boltzmann weights of the form
\begin{equation}
\exp\left[K^*_{\mu m}\cos(m\Theta_{({\bm{r}},\mu\nu)})\right],
\end{equation}
with \(\mu<\nu=1,2,3,4,\) and \(m=0,\cdots,p-1\).
The dual couplings \(K^*_{\mu m}\) can be computed in closed form with
the help of the results of Appendix \ref{appF}. This completes our discussion of the
duality properties of Wilson-type \({\mathbb{Z}}_p\) gauge theories.
Next we would like to follow a different approach that emphasizes the possibility
of using {\it quantum} models with known duality properties to construct {\it classical}
models with known duality properties as well. Notice that this methodology
(quantum to classical)
reverses the
direction of the path that we have been following in Section \ref{sec8} (classical
to quantum).
In the light of previous discussion,
we see that the quantum $p$-state gauge model studied in Section \ref{sec6.2},
for $p\geq 5$,
is not
exactly related to the Wilson-type action of Equation \eqref{class_zpgauge}.
On the other hand, that quantum model is exactly self-dual for any \(p\), so we would like
to determine its related classical \(p\)-state gauge theory.
To this end, we need to identify what classical gauge model \(\mathcal{Z}_{\sf sdG}\),
in an appropriate gauge, has
\begin{equation}
T_{\sf sdG}=e^{\frac{\kappa}{2}\sum_{\bm{r}}\sum_{\mu=1}^{3}\ (V_{({\bm{r}},\mu)}+
V^\dagger_{({\bm{r}},\mu)})}\ e^{\frac{K}{2}\sum_{\bm{r}}\sum_{\mu=1}^{3}\ (B_{({\bm{r}},\mu)}+
B^\dagger_{({\bm{r}},\mu)})}
\end{equation}
for transfer matrix. Then it will follow from the self-duality of Equation
\eqref{map_gauge_zp} that \(\mathcal{Z}_{\sf sdG}\) is exactly self-dual {\it for any p}.
It is not difficult to compute
the matrix elements of \(T_{\sf sdG}\) (see Appendix \ref{appF}). Then we can reconstruct the
partition function
\begin{eqnarray}\label{pstateEM}
\mathcal{Z}_{\sf sdG}&=&\sum_{\{\theta_{({\bm{r}},\mu)}\}}\ \exp \left [ \sum_{\bm{r}} K (
\cos \Theta_{({\bm{r}},12)}+ \cos \Theta_{({\bm{r}},23)}+\cos \Theta_{({\bm{r}},31)}) \right ] \times\\
&&\ \ \ \exp \left [\sum_{\bm{r}}
\sum_{m=0}^{p-1}\ \epsilon_m(\kappa) (\cos(m\Theta_{({\bm{r}},41)})+\cos(m\Theta_{({\bm{r}},42)})
+\cos(m\Theta_{({\bm{r}},43)}) ) \right] .\nonumber
\end{eqnarray}
This is the new, exactly self-dual classical gauge theory we were after. Notice that this
self-duality exchanges the Boltzmann weight of the second line of Equation
\eqref{pstateEM}, that can be associated with the Boltzmann weight of the
\({\mathbb{Z}}_p\) electric field, with the one of the first line, that can be associated
with the Boltzmann weight of the \({\mathbb{Z}}_p\) magnetic field.
Finally, let us compute the coefficients \(\epsilon_m(\kappa)\),
which are an essential ingredient in the definition of \(\mathcal{Z}_{\sf sdG}\).
The starting point is to rewrite
\(
e^{\frac{\kappa}{2}(V_{({\bm{r}},\mu)}+
V^\dagger_{({\bm{r}},\mu)})}=\sum_{m=0}^{p-1}e^{u_\kappa(m)}V^m_{({\bm{r}},\mu)}\ ,
\)
with (\(\theta_s=2\pi s/p\))
\begin{equation}
u_\kappa(m)=\ln\left[\frac{1}{p}\sum_{s=0}^{p-1}\cos(m\theta_s)e^{\kappa
\cos\theta_s}\right] .
\end{equation}
Since \(u_\kappa(l)\) is even,
it can be represented as a discrete cosine series, \(u_\kappa(l)=\sum_{m=0}^{p-1}\
\epsilon_m(\kappa) \cos(l\theta_m)\).
This completes the specification of the \(\epsilon_m(\kappa)\).
\subsection{Dualities for continuum models of classical statistical physics}
\label{sec8.5}
We have so far considered lattice models of classical statistical mechanics. In this
section we want to show by example that our bond-algebraic approach can indeed
be used to establish duality transformations in many-body problems defined
in continuum space-time.
Consider the Hamiltonian of a chain of coupled quantum harmonic oscillators
\begin{equation}\label{sd_phonons}
H_{\sf Ph}[1/m, k]=
\sum_i\ ( \frac{p_i^2}{2m}+\frac{1}{2}k(x_{i+1}-x_i)^2)
=\ \mathcal{U}_{{\sf d}}^\dagger\, H_{\sf Ph}[k,1/m]\, \mathcal{U}_{{\sf d}}\ ,
\end{equation}
whose normal modes represent acoustic phonons. Last equality indicates the
quantum self-duality relation derived in Section \ref{sec4.3} with
$k=m\omega^2$.
On one hand, standard manipulations which involve the
STL decomposition used before \cite{bookEPTCP},
but now applied to continuous degrees of freedom \cite{schulman},
lead to
\begin{eqnarray}
{{\sf Tr}\ } e^{-\beta H_{\sf Ph}[1/m,k]}&=&\mathcal{N}\lim_{N\rightarrow\infty}
\int\prod_{j=1}^N\prod_i dx_{i,j} \\
&\times& \exp\left[-\sum_{j=1}^N\sum_i\ \left(
\frac{mN}{2\beta}(x_{i,j+1}-x_{i,j})^2+\frac{k\beta}{2N}
(x_{i+1,j}-x_{i,j})^2\right)\right] \nonumber ,
\end{eqnarray}
with $\cal N$ a normalization factor. This mapping
relates a chain of coupled
quantum harmonic oscillators to a classical $D=2$ array of springs.
On the other hand, it is standard in the path integral
context to interpret the limit
\(N\rightarrow \infty\) as a sum over classical path
configurations (in Euclidean time),
that is, as an Euclidean path integral \cite{schulman}
\begin{eqnarray}
{{\sf Tr}\ } e^{-\beta H_{\sf Ph}[1/m,k]}&=&\\
&&\hspace*{-3.0cm}=\int \prod_i \mathcal{D}x_i\ \exp\left[\int_0^\beta d\tau\
\sum_i\ \left(\frac{1}{2}m\left(\frac{dx_i}{d\tau}\right)^2+
\frac{1}{2}k(x_{i+1}-x_i)^2\right)\right]\nonumber\\
&&\hspace*{-3.0cm}=\int \prod_i \mathcal{D}x_i\ \exp\left[\int_0^\beta d\tau\
\sum_i\ \left(\frac{1}{2}m^*\left(\frac{dx_i}{d\tau}\right)^2+
\frac{1}{2}k^*(x_{i+1}-x_i)^2\right)\right ], \nonumber
\end{eqnarray}
where \(\mathcal{D}x_i\) denotes the Wiener measure on the space of paths.
The last equality results from applying the quantum self-duality of Equation
\eqref{sd_phonons}, and thus constitutes an elementary example of a duality for Feynman
path integrals. The dual couplings \(m^*,\ k^*\) satisfy the relations
\begin{equation}
km^*=1,\ \ \ \ \ \ \ \ \ \ mk^*=1.
\end{equation}
\subsection{Classical disorder variables from quantum ones}
\label{appB}
This section exploits the combined application of
the transfer matrix and bond-algebraic
techniques to compute classical disorder variables for arbitrary
models of classical statistical mechanics.
The notion of {\it disorder variable}
was introduced in the context of the $D=2$ classical Ising model
\cite{kadanoff}, and further exploited
in Reference \cite{Kadanoff_Ceva}, as part of
a scheme, based on the operator
product expansion technique, to compute its critical exponents.
It seems
reasonable to expect that classical disorder variables (defined in terms
of the classical Kramers-Wannier duality
\cite{Kadanoff_Ceva}) should be related to quantum dual variables,
but the explicit connection has not been published
before, most likely because the relation only becomes self-evident
when quantum dualities are recognized as unitary transformations.
This connection is, however, of great importance because
there is no simple way to generalize the construction
of classical disorder variables for the Ising model to different models,
other than the {\it quantum route} that we are going to take next.
This route depends critically on recognizing quantum dualities
as unitary transformations.
Consider a $D=2$ Ising model on an $M\times N$ lattice with
cylindrical topology as in Section \ref{sec8.2}
(\(N\) is the number of sites along the periodic BC).
We can write its partition function in terms of a transfer matrix $T$ as
\begin{equation}
\mathcal{Z}_{\sf I}(K_1,K_2)= {{\sf Tr}\ } T^N,
\end{equation}
and more importantly, we can write two-point correlation functions as
\begin{equation}\label{two_point}
\langle \sigma_{m',n'}\ \sigma_{m,n}\rangle(K_1,K_2)=
\frac{{{\sf Tr}\ } ( T^{(N-n')}\ \sigma^z_{m'}\ T^{(n'-n)}\ \sigma^z_m\ T^n)}{{{\sf Tr}\ } T^N} .
\end{equation}
On the other hand, the transfer matrix \(T\) can be related to a quantum
problem
\begin{eqnarray}
&& \frac{T(K_1,K_2)}{\left(2\sinh(2K_1) \right)^{M/2}}=e^{-H^1[h_1]}e^{-H^0[K_2]},
\end{eqnarray}
where
\begin{eqnarray}
H^1[h_1]=-h_1 \sum_{i=1}^{M} \sigma^x_{i} , \ \ \ \
H^0[K_2]=-K_2\sigma^z_{M}-K_2\sum_{i=1}^{M-1}
\sigma^z_{i}\sigma^z_{i+1} ,
\end{eqnarray}
with \(h_1=-\frac{1}{2}\ln\tanh K_1\).
Applying the bond-algebraic results of
Section \ref{sec3.7} one can show that \(H^1\)
is dual to \(H^0\),
\begin{equation}
H^1[h_1]=\mathcal{U}_{{\sf d}}^\dagger H^0[h_1]\mathcal{U}_{{\sf d}},\ \ \ \
H^0[K_2]=\mathcal{U}_{{\sf d}}^\dagger H^1[K_2]\mathcal{U}_{{\sf d}},
\end{equation}
which implies that the two-point correlation
defined in Equation \eqref{two_point} can be written as
\begin{eqnarray}
\langle \sigma_{m',n'}\ \sigma_{m,n}\rangle(K_1,K_2)&=& \nonumber \\
&=&
\frac{{{\sf Tr}\ } ( \mathcal{U}_{{\sf d}} T^{(N-n')}\mathcal{U}_{{\sf d}}^\dagger\
\mathcal{U}_{{\sf d}} \sigma^z_{m'}\mathcal{U}_{{\sf d}}^\dagger\ \mathcal{U}_{{\sf d}}
T^{(n'-n)}\mathcal{U}_{{\sf d}}^\dagger\ \mathcal{U}_{{\sf d}}\sigma^z_m\mathcal{U}_{{\sf d}}^\dagger\
\mathcal{U}_{{\sf d}} T^n\mathcal{U}_{{\sf d}}^\dagger)}{{{\sf Tr}\ }
(\mathcal{U}_{{\sf d}} T^N\mathcal{U}_{{\sf d}}^\dagger)} \nonumber \\
&=&\frac{{{\sf Tr}\ } (\widehat{T}^{(N-n')}\ \mu^z_{m'}\ \widehat{T}^{(n'-n)}\ \mu^z_m\
\widehat{T}^{n})}{{{\sf Tr}\ } \widehat{T}^{N}} \nonumber \\
&=& \langle \mu_{m',n'}\ \mu_{m,n}\rangle(K_1^*,K_2^*) ,
\label{kada_ceva_dual}
\end{eqnarray}
with quantum dual variables $\mu^z_m$ defined in Section \ref{sec3.10},
and $K^*_1=-\frac{1}{2}\ln\tanh K_2$, $K^*_2=h_1$.
Since the classical disorder variables of Reference \cite{Kadanoff_Ceva}
are essentially defined through the relation \eqref{kada_ceva_dual},
we see that the quantum dual variables \(\mu^z_m\) are indeed
quantum disorder variables themselves. Notice that in contrast
to the classical approach of Reference \cite{Kadanoff_Ceva}, the
quantum approach allow us to compute
\begin{equation}
\langle\mu_{m,n}\rangle(K_1^*,K_2^*)=
\frac{{{\sf Tr}\ } (\widehat{T}^{(N-n)}\ \mu^z_m\
\widehat{T}^{n})}{{{\sf Tr}\ } \widehat{T}^{N}},
\end{equation}
a quantity that could not even be defined at the classical level
(the approach of Reference \cite{Kadanoff_Ceva} can only make sense
of {\it correlators} of disorder variables).
While it would be impossible to extend the techniques
of Reference \cite{Kadanoff_Ceva} to, say, the VP model,
we see now that our bond-algebraic technique afford an straightforward
solution to the problem of defining {\it classical} disorder
variables in general: They can be derived from their quantum
counterparts and the transfer matrix formalism.
\section{Applications of dualities}
\label{sec9}
Thus far, we illustrated, how to derive (hitherto known and also unknown)
dualities within our bond algebraic approach.
As we have shown, our approach applies to both quantum and classical
dualities. In the current section, we will discuss applications of dualities.
We will present some spectral consequences of dualities, general
techniques such as fermionization in arbitrary spatial dimensions, integrability conditions,
and dimensional reductions.
\subsection{Self-dualities and phase transitions}
\label{sec9.1}
As is well appreciated, one of the most powerful consequences of
dualities are constraints that may be imposed on the phase diagrams
of dual systems. These become particularly
potent in the case of self-dual systems.
Thus, we will now turn to the practical consequences
of self-dualities. Specifically, we will:
(i) Analyze the relation between self-dualities and the existence/non-existence of phase
transitions, and the resulting constraints for the spectrum of
self-dual systems. We discuss extensions to situations wherein more
than one coupling constant is present and in which
a self-dual point is replaced by a self-dual line
or surface.
(ii) Detail some consequences of the constraint
for derivatives of general quantities at and away
from the self-dual point.
(iii) Briefly discuss the general solution to the constraint discussed above.
\subsubsection{Self-duality and the existence/non-existence of
phase transitions}
A finite temperature phase transition is characterized by a non-analyticity in the
free energy of the system at the transition point (or transition line, etc.). Similarly, a zero temperature
(quantum) phase transition relates a level crossing, or an avoided level crossing,
and non-analyticities in the ground state energy
of the quantum Hamiltonian at hand.
When present, a self-duality relates the energy levels (and thus the free energy and all related
thermodynamic quantities) at one coupling (or temperature) to those at another
coupling (or temperature). Thus, whenever a phase transition
occurs at one value of the coupling $\lambda$ it must also occur
at the dual coupling $\lambda^*$. A corollary of the above is that
(1) if the system exhibits an odd
number of transitions as a function of $\lambda$ then
one transition must occur at the self-dual point $\lambda_{\sf sd}$.
Similarly, (2) if an even number of transitions are present, then
no phase transition can occur at $\lambda_{\sf sd}$.
A further consequence of self-dualities is that (3) if a system is devoid of
transitions in a particular region $\lambda_a < \lambda <\lambda_b$ then it must
also be devoid of transitions in the related dual region whose
endpoints are given by $\lambda_a^*$ and $\lambda_b^*$
(where $\lambda_a^*$ and $\lambda_b^*$ are
the dual counterparts to $\lambda_a$ and $\lambda_b$). Similar remarks can
be made when more than one coupling constant (and/or temperature)
are involved. In such cases self-dual points translate into
self-dual lines or surfaces and regions devoid of singularities
appear in a higher dimensional parameter space. These statements
are simple yet proved to be extremely potent
over many decades.
We comment on examples in which cases (1)-(3) above are, respectively, realized:
(1) As discussed in detail in earlier sections,
the classical $D=2$ Ising model is self-dual \cite{KW}
and displays only a single transition separating the high temperature
disordered phase to the low temperature ordered phase. Thus,
the $D=2$ Ising model orders at the self-dual inverse
temperature given by $\beta_{c} = (\ln(1+ \sqrt{2})/2)$.
This value found by Kramers and Wannier matches, as it
must, the critical temperature found by Onsager
in his exact solution of the same model.
(2) The $p$-state VP model of Section \ref{sec4.1.1} exhibits, for $p>4$, three phases
and thus its self-dual point does not correspond to a point of non-analyticity \cite{ortiz}.
(As noted earlier, for $p=2$, the VP model becomes the Ising model
of case (1) above; the same also holds true for $p=4$.)
(3) By the use of the self-duality of the $D=3$ Ising matter coupled gauge theory,
it can be shown \cite{nprd} that the confining
phase of this system (weak couplings) is smoothly connected to
its Higgs phase (when all couplings
are large). When the union of the region of
phase space that is free of transitions
(as proved by the Lee-Yang theorem)
is taken with its dual counterpart, there
is a region that is free of non-analyticities
connecting the above two phases \cite{fradkin_shenker}.
The above three consequences, which can be appended by additional
constraints that we will elaborate below, can be applied, mutatis mutandis,
not only to questions concerning thermodynamic phase transitions
but also to general non-equilibrium phenomena. For instance, we may consider the dynamics
derived from a self-dual Hamiltonian (whether classical
or quantum). The equations of motion governing the system
dynamics are identical under
the interchange of a coupling constant $\lambda$ with its dual $\lambda^*$.
Consequently, both in the quantum and classical arenas, any transitions associated
with the character of the system dynamics as parameters are changed
must satisfy relations (1)-(3) when the Hamiltonian is self-dual.
We now comment on the spectral properties of self-dual Hamiltonians.
Consider a Hamiltonian of the form
\begin{eqnarray}
H[\lambda] = H_{0} + \lambda H_{1}.
\label{hlh1}
\end{eqnarray}
for which a duality transformation
$\mathcal{U}_{{{\sf d}}} H_{0} \, \mathcal{U}_{{{\sf d}}}^{\dagger} = H_{1}$, $
\mathcal{U}_{{{\sf d}}} H_{1} \, \mathcal{U}_{{{\sf d}}}^{\dagger} = H_{0}$,
interchanges the types of bonds present in the two Hamiltonians $H_{0}$ and $H_{1}$.
Thus, $\mathcal{U}_{d}$ is a unitary operator that
implements a self-duality of the Hamiltonian of Equation \eqref{hlh1}, i.e.,
$\mathcal{U}_{{{\sf d}}} H[\lambda] \, \mathcal{U}_{{{\sf d}}}^{\dagger}=\lambda H[1/\lambda]$
with a self-dual point $\lambda_{\sf sd}=1$. This implies that the eigenvalues satisfy
\begin{equation}
E_{n}(\lambda) = \lambda E_{n}(1/\lambda).
\label{evaldual}
\end{equation}
Equation (\ref{evaldual})
constitutes the most general constraint of self-duality
on a Hamiltonian of the type of Equation (\ref{hlh1}) \cite{kogut}.
(We allude here to the ``most general'' constraint as
dualities encompass unitary transformations
(thus preserving the spectrum) and Equation \eqref{evaldual}
is the sole constraint on the energy eigenvalues $E_{n} (\lambda)$ that arises from the duality.)
It follows that the free energy of the
quantum system
at an inverse temperature $\beta$
similarly satisfies $F_q(\beta, \lambda) = \lambda F_q(\beta, 1/\lambda)$.
By taking derivatives of the free energy, it is seen that
the (average) internal energy and other general thermodynamic
quantities satisfy identical relations. Relating
values of $\lambda >1$ to those with $\lambda <1$ suggests
a ``halving" of the degrees of freedom. We will, later on, explicitly see
various manifestations of this.
With the identification of the self-duality relation $\lambda \leftrightarrow \lambda^* = 1/\lambda$,
we may now invoke the likes of corollaries (1)-(3) above.
If a level crossing occurs at a point $\lambda$
then a level crossing must occur at the point $\lambda^* = 1/ \lambda$.
That is, if there exist two levels (denoted by $n$ and $m$)
that cross at a particular coupling $\lambda$: $E_{n}(\lambda) = E_{m}(\lambda)$
then it follows from Equation \eqref{evaldual} that $E_{n}(1/\lambda) = E_{m}(1/\lambda)$.
Also, if there exists two quantum phase transitions (two non-analyticities in
$E_0(\lambda)$, as in the VP model of Section \ref{sec4.1.1}), and one happens at
the point $\lambda_{c1}$, the second must happen at $\lambda_{c2}=1/\lambda_{c1}$, such that
$\lambda_{c1} E_0(\lambda_{c2})=E_0(\lambda_{c1})$.
When examining the classical analogue of a zero temperature quantum system
defined by a Hamiltonian of the form of Equation (\ref{hlh1}), the $\lambda \leftrightarrow 1/\lambda$
duality transformation translates, similar to discussions in earlier sections,
into a (generally non-trivial) duality transformation
relating the inverse temperatures $\beta \leftrightarrow \beta^*$.
Such a case occurs, as we saw earlier, in the quantum Ising
chain (whose Hamiltonian is of the form of Equation \eqref{hlh1} and
whose classical counterpart is given by the $D=2$ Ising model).
A relation similar to that of Equation \eqref{evaldual} trivially appears for
a Hamiltonian that is symmetric under the permutations of
more than one type of a pair of bonds. That is,
we may consider $H[\lambda_{1}, \lambda_{2}, \cdots, \lambda_{p}]=
H_{0} + \lambda_{1} H_{1} + \lambda_{2} H_{2} + \cdots
+ \lambda_{p} H_{p}$ with different unitary operators $\mathcal{U}_{{{\sf d}}}$ that may
exchange, for instance, $H_{0}$ with $H_{i >0}$ (i.e.,
$\mathcal{U}_{{{\sf d}}} H_{i} \, \mathcal{U}_{{{\sf d}}}^{\dagger} = H_{0}$
and its inverse). In such a case, the simple extension
of Equation \eqref{evaldual} in a higher dimensional coupling space
is
\begin{eqnarray}
H[\lambda_{1}, \cdots, \lambda_{p}] = \lambda_{i}
H[\lambda_{1},\cdots, \lambda_{i-1}, 1/\lambda_{i}, \lambda_{i+1}, \cdots, \lambda_{p}].
\end{eqnarray}
We next briefly and explicitly discuss constraints appearing for classical self-dualities
where the free energies of two dual classical systems are the same
only up to an additive non-singular contribution (see Equation \eqref{first_time_statistics}).
In such
cases the free energy of the classical system $F$ satisfies
\begin{eqnarray}
\label{simple_f}
F(K) = F(K^*) + f (K,K^*)
\end{eqnarray}
where $f$ is a regular function
of $K$ and $K^*$.
Differentiation of $F$ relative to temperature yields the average energy.
At the self-dual point $K= K^*=K_{c}$, the energy is given by
$E= \frac{\partial f}{\partial K}|_{K=K_{c}}/
( 1- \frac{\partial K^*}{\partial K}|_{K= K_{c}})$.
For the classical $D=2$ ($N$ sites) Ising model wherein $\exp(-2 K^*)= \tanh K$,
we may determine the exact energy at the self-dual point, which is equal to
$E/N= -\sqrt{2}$ \cite{bookEPTCP}.
Similarly, by differentiating Equation \eqref{simple_f} twice relative to $K$,
we find that
\begin{eqnarray}
T^{2} C_{V}(K) - (T^{*})^{2} C_{V}( K^*)
\left (\frac{\partial K^{*}}{\partial K}\right )^{2} =
- \left (\frac{\partial^{2} f}{\partial K^{2}} + E(K^*)
\frac{\partial^{2} K^*}{\partial K^{2}} \right ),
\end{eqnarray}
where $C_{V}$ is the specific heat at constant volume.
As is well appreciated \cite{bookEPTCP}, the existence of a self-dual point
implies that whenever it is a critical point
the critical indices
must be the same on both sides of the transition
(as they always are in any system, self dual or
not) and that the {\em amplitudes} associated
with the self-dual point must be the same
on both sides of the transition. (In
general critical systems, the amplitudes
need not be the same on both sides of
the critical point.) Any singular contributions in the vicinity
of the critical point must be given by
the dependence of the free energy $F$
on both sides of the transition point.
If $K^*$
is linear in $K$ near the critical point,
then as the derivatives of the same
function $F$ on both sides of the
transition point will determine the
behavior of any critical quantity,
by virtue of Equation \eqref{simple_f},
the critical behavior must
be the same on both
sides of the transition.
\subsubsection{Constraints in the absence of phase transitions}
Equation \eqref{evaldual} leads to constraints on the derivatives of the energy levels and
all thermodynamic quantities in the absence of phase transitions.
Specifically, by differentiating both sides of Equation \eqref{evaldual}, we find that
($E^{(j)}_n(x)=\partial^{j} E_n(x)/\partial^j x$)
\begin{eqnarray}
\label{longnd}
E_n^{(1)} (\lambda) &=& \frac{\lambda E_n(1/\lambda) - E_n^{(1)}(1/\lambda)}{\lambda}, \nonumber
\\ E_n^{(2)} (\lambda) &=& \frac{E_n^{(2)}(1 /\lambda)}{\lambda^{3}}, \nonumber
\\ E_n^{(3)}(\lambda)& =& - \frac{3 \lambda E_n^{(2)}(1/ \lambda)
+ E_n^{(3)}(1/\lambda)}{\lambda^{5}}, \nonumber
\\ E_n^{(4)}(\lambda) &=& \frac{12 \lambda^{2} E_n^{(2)}(1/\lambda)
+ 8 \lambda E_n^{(3)}(1/\lambda)
+ E_n^{(4)}(1/\lambda)}{\lambda^{7}}, \cdots .
\end{eqnarray}
If $E_{n}(\lambda$) is analytic in a domain that includes the
self-dual point $\lambda_{\sf sd} =1$, then
\begin{eqnarray}
\label{sedl}
E_{n}^{(1)} (1) &=& \frac{1}{2} E_{n}(1), \nonumber
\\ E^{(3)}(1) &=& - \frac{3}{2} E^{(2)}(1), \nonumber
\\ E^{(5)}(1) &=& 15 \left (E^{(2)}(1) - \frac{1}{2} E^{(4)}(1) \right ), \cdots .
\end{eqnarray}
The spectrum of the self-dual system captured by Equation \eqref{evaldual} gives
rise to equivalent (``dual'') pairs of equations from even and odd orders in
$(\lambda-\lambda_{\sf sd})$
about the self-dual point. This manifests the aforementioned ``halving'' of the parameters
characterizing the function. The large degeneracy
manifest in these equations enables a large number of possible
solutions to Equation \eqref{evaldual} as we discuss next.
\subsubsection{General self-dual spectra}
The spectra that are analytic at the self-dual
point form only a small subset of all possible solutions of Equation \eqref{evaldual}.
The self-dual point may mark a transition point between two different phases
(wherein $\{E_{n}(\lambda)\}$ are no longer differentiable to arbitrary order).
Most of the examples that we considered in this article fall into this category.
The self-dual point of the quantum Ising chain ($h=J$ in
Equation \eqref{infinite_ising_transverse}, i.e., $\lambda = J/h =1$ at the self-dual point)
constitutes a point where a quantum phase transition occurs;
in particular, the gap between the ground state and the first excited
state of the quantum Ising chain scales as $\Delta E = 2 h |1- \lambda|$ \cite{kogut}.
We now discuss the most general possible form of the self-dual
energies. Identical forms to those presented appear for
all thermodynamic quantities in self-dual systems.
Equation \eqref{evaldual} is the sole
condition imposed on the spectrum
from self-duality.
It is easy to see, by direct substitution, that if $w_{n}(\lambda)$
is a solution to Equation \eqref{evaldual} then so is $w_{n}(\lambda) +
\lambda w_{n} (1/\lambda)$.
Conversely, for {\em any} function $w_{n}(\lambda)$, the combination
$w_{n}(\lambda) + \lambda w_{n} (1/\lambda)$ satisfies Equation \eqref{evaldual}.
Thus, Equation \eqref{evaldual} is
satisfied if and only if
\begin{eqnarray}
E_{n}(\lambda) = w_{n}(\lambda) + \lambda w_{n}(1/\lambda),
\label{ensd}
\end{eqnarray}
with $w_{n}$ representing arbitrary functions.
This general solution suggests a halving
of the degrees of freedom allowed by the function
(the function formed by the sum in Equation \eqref{ensd} must be ``even'' under the interchange of
$\lambda$ with $1/\lambda$) and trivially allows for a rich variety of forms.
Depending on the form of
the functions $w_{n}$ in Equation \eqref{ensd},
the spectra $E_{n}(\lambda)$ can be either analytic or non-analytic at
$\lambda = \lambda_{\sf sd} =1$. Similarly, these functions
enable transition points $\lambda^*$
where level crossing occurs (in particular, those where the ground state changes character
as the gap between the ground state and the lowest excited state vanishes) to
be such that the energy variations to a given order are continuous or discontinuous.
An equivalent alternate solution to Equation (\ref{evaldual}) in the
non-analytic case is simple.
Consider any set of functions $\{E_{n}(\lambda)\}$
defined on the interval $[\lambda_{\sf sd}, \infty)$,
and from this set define the functions on the remaining segment $\lambda \in (0,\lambda_{\sf sd}]$.
This is similar to a standard complex inversion transformation
(applied now only on the half real line
$\lambda \ge 0$) with the additional scale factor of $\lambda$ on the
right-hand side of Equation \eqref{evaldual}.
We can make these functions continuous and {\em non-differentiable} at $\lambda=\lambda_{\sf sd}=1$
by defining functions on $[\lambda_{\sf sd}, \infty)$ such that the function has an
ill-defined derivative on reflection (Equation (\ref{evaldual})) as the left
hand and right hand derivatives (if they exist)
do not match.
For instance, if for a particular level (say, $n=0$) the eigenvalue behaves as
$E_{0}(\lambda) = \sqrt{\lambda-1}$
for $ \lambda \ge 1$ then we will need to set $E_{0}(\lambda) = \sqrt{\lambda(1-\lambda)}$ for
all $0 < \lambda \le 1$ in order to satisfy Equation (\ref{evaldual}).
Similar constructs can be implemented for all levels $n$.
The same may also be done for higher order derivatives
(for, e.g., a divergent second order derivative at $\lambda=1^{+}$).
\subsection{Correlation functions}
\label{sec9.2}
The unitary character of the duality transformation emphasized in this article
(Equation \eqref{d_as_u}) further allows for a related, very
simple but powerful, relation concerning the correlation lengths
(and times). To establish this relation, we can employ the
transfer matrix formalism. We consider a classical system in which along
one spatial (or temporal) direction, the system has length $N_{1}$
and a corresponding transfer matrix $T$. As detailed in Section \ref{classical&quantum} and Section
\ref{sec8},
the transfer matrix in the classical problem may be related to a quantum Hamiltonian
(and viceversa). If two quantum Hamiltonians $\tilde{H}(\tilde{\phi})$ and $H(\phi)$
are dual to each other (and thus share the same spectrum), the time evolution of
the dual fields $\tilde{\phi}$ and $\phi$ are identical (the corresponding eigenstates
of the two Hamiltonians evolve with identical frequencies). The same holds true
for the imaginary time evolution of two sets of fields. In particular,
the gap $\Delta E$ between
the ground state and the first excited state (which determines the asymptotic long time
correlations of the system) is the same in both systems. The same, of course, holds true
not only for the long time limit but for any time separation.
Within the quantum to classical correspondence \cite{bookEPTCP},
an imaginary time axis in the quantum problem is replaced by
an additional spatial axis in a static equilibrium thermodynamic
classical problem. As detailed in earlier sections,
the eigenvalues $\{ E_{n} \}$ of the quantum Hamiltonian $H$
are replaced
by the eigenvalues $\{\lambda_{n} \}$ of the classical transfer matrix $T$.
Commonly, the transfer matrix eigenvalues can be numbered in list of descending absolute
values $|\lambda_{0}| \ge |\lambda_{1}| \ge |\lambda_{2}| \ge \cdots$.
For finite size transfer matrices (as these commonly arise in $D=1$ classical systems,
or 0+1 space-time dimensional quantum systems), the Perron-Frobenius theorem
guarantees that the largest eigenvalue ($\lambda_{0}$) is non-degenerate.
In higher dimensional systems (i.e., classical $D>1$ dimensional
systems), the eigenvalues can become degenerate (and indeed
do become degenerate) at critical transitions.
In terms of the transfer matrix
eigenvalues $\{\lambda_{n} \}_{n=0}^{\dim{T}-1}$
(with $\dim{T}$ being the dimension of the transfer
matrix $T$),
the partition function reads
\begin{eqnarray}
\label{transfer_Z}
{\cal Z} = \sum_{n=0}^{\dim{T}-1} \lambda_{n}^{N}.
\end{eqnarray}
For $D=1$ systems, the transfer matrix can have a finite number
of eigenvalues while in higher dimensions the number of eigenvalues
(the size of the transfer matrix) is infinite, in the thermodynamic limit
in which the system is of infinite extent along all spatial directions.
The correlation function is completely determined by the eigenvalues
and eigenvectors of the transfer matrix. In particular, the inverse correlation length
determining the correlations at large distances
is given by
\begin{eqnarray}
\xi^{-1} = - \ln (\lambda_{1}/\lambda_{0}).
\label{transfer-xi}
\end{eqnarray}
(Within the corresponding quantum problem in imaginary time,
an analogous relation relates the correlation time with the
gap between the first excited state and the ground state; the transfer
matrix eigenvalues of the classical problem are related to exponentials of
the energies in the quantum problem.)
As emphasized in this article, a duality between two systems implies the
equivalence of the spectrum of the two theories (up to an overall constant
factor). The partition functions of dual systems are equal to one
another. In what follows, we will denote the $n$th transfer matrix eigenvalues of
systems (systems (1) and (2)) by $\lambda_{(1);n}$ and $\lambda_{(2);n}$.
{}From Equation \eqref{classd_from_q} we have
\begin{eqnarray}
{\cal Z}_{1} = \sum_{n=1}^{\dim{T}-1} \lambda_{(1);n}^{N} = A \ {\cal Z}_{2} =
A \sum_{n=0}^{\dim{T}-1} \lambda_{(2);n}^{N} .
\end{eqnarray}
Amongst other things, this implies that the correlation lengths
of the two systems are the same, as by Equation (\ref{transfer-xi})
\begin{eqnarray}
\xi_{(1)}^{-1} = - \ln(\lambda_{(1);1}/\lambda_{(1);0})
= - \ln(\lambda_{(2);1}/\lambda_{(2);0}) = \xi_{(2)}^{-1}.
\end{eqnarray}
Analogous relations follow not only for the asymptotic large distance correlation length
but rather for all correlation lengths set by $[- \ln (\lambda_{n}/\lambda_{0})]$.
\subsection{Fermionization as a duality}
\label{sec9.3}
As is well known, the quantum {\it exchange} statistics of elementary
degrees of freedom (whether fermionic, bosonic, or spin) can, quite generally,
be readily transformed. For instance, spin $S=1/2$ $SU(2)$ operators can be mapped onto
spinless fermions by a transformation known as the Jordan-Wigner transformation
\cite{JW}, and generalizations to higher spin, or even arbitrary algebras
exist \cite{GJW}. The generalized Jordan-Wigner mappings represent ingenious constructs
based on non-local isomorphic mappings between the degrees of freedom in question \cite{GJW}. Since these
dictionaries are independent of any Hamiltonian (or action) they generically fail
to preserve locality by introducing strings in the interactions, in particular in spatial dimensions $d>1$.
Amongst many other benefits, these mappings readily allows for
exact solutions of many \(d=1\) dimensional models, including the XY
and transverse field quantum Ising models. This is so as the transformations map
spin quadratic forms onto non-interacting fermionic terms (i.e., quadratic forms),
that may be exactly solved by algebraic means.
By contrast, dualities are {\it model specific transformations}
that always preserve locality (in the Hamiltonians), and are designed to take
the most advantage of each model's peculiarities. To see how transmutation of
statistics is possible, notice for example that bond algebras of fermionic systems
feature only (sums of)
{\it even products} of the elementary fermionic degrees of freedom.
Since such even products behave very differently from the fermions
themselves, they could in principle be mimicked by bonds of models
featuring other types of elementary degrees of freedom. It follows from the formalism of
Section \ref{sec3.5}, in the absence of gauge symmetries,
that the {\it dual variables} that arise from a {\it
duality connecting models with different statistics} will afford a
dictionary
mediating the two statistics. Dualities can also help improve the efficiency of
standard fermionization techniques in dimension \(d\geq 2\),
because different dual representations
of one and the same model can behave very differently under generalized
Jordan-Wigner transformations.
The outline of this section is as follows:
{(i)} First, we will analyze
the Jordan-Wigner transformation
through the prism of bond algebras
and illustrate that it is not at
all necessary to consider non-local
transformations \cite{anfossi_montorsi}. Rather, as alluded to
above, we may focus on the bond algebras
of local bonds (interactions) to illustrate
a mapping from one local Hamiltonian to another. The
Jordan-Wigner dictionaries arise as dual variable mappings.
{(ii)} We will then show that the generalization of
a higher dimensional Jordan-Wigner transformation
that maps nearest-neighbor bilinear interactions
of different underlying statistics
to one another is strictly impossible
on lattices that have closed loops. In other words,
there is no local fermionization mapping, that preserves
the number of degrees of freedom, in dimensions
$d \geq 2$. There are, of course, four-body (plaquette) interactions
where a local mapping is still possible.
{(iii)} Finally, we will show that while
mappings involving local nearest-neighbor
bilinears, and preserving the number
of degrees of freedom, are impossible, mappings
involving gauge-reducing dualities
are indeed possible. We illustrate
this point by fermionizing the \(d=2\) Ising model via its duality to the
\(\mathbb{Z}_2\) Ising gauge model. Nonetheless,
dualities that connect models with different statistics
are constrained by quantum-mechanical considerations
{\it that depend on the space-time dimension \(D\)}, and they become
increasingly challenging as \(D\) increases. We will show
an example of this in a family of models that
can be fermionized in \(d=1\) dimensions, but
cannot be fermionized in higher dimensions.
We would like to mention that $\it bosonization$, a process which not only
effects exchange statistics transmutation to canonical bosons but also modifies
the {\it exclusion} statistics \cite{GJW}, could also be in principle
interpreted as a duality mapping but it is a less precisely defined
mathematical transformation which requires some vacuum regularization process.
\subsubsection{Transmutation of statistics
via bond algebras: Dual variables and the Jordan-Wigner transformation}
\label{stba}
In this section we want to derive the standard Jordan-Wigner mapping as
a duality transformation. As
emphasized throughout this article, practically what matters is
not the algebra of the elementary degrees of freedom but rather the algebra
of the bonds that appear in the Hamiltonian. Consider the simple model of
non-interacting (spinless) fermions in $d=1$, characterized by the
$N$ sites tight-binding Hamiltonian
\begin{eqnarray}
\label{Hferm}
H_{\sf fermion} = \lambda \sum_{i=1}^{N-1}
(c^\dagger_{i}c^{\;}_{i+1}+c_{i+1}^{\dagger}
c^{\;}_{i}),
\end{eqnarray}
where $c^{\;}_{i}$ and $c_{i}^{\dagger}$ annihilate or create a
spinless fermion at site $i$, and
\begin{equation}\label{car}
\{c^{\;}_i,\ c^{\;}_j\}=0,\ \ \ \ \ \ \{c^{\;}_i,\ c^\dagger_j\}=\delta_{i,j}
\end{equation}
(\(\{A,B\}=AB+BA\)). The bonds $c_{i+1}^{\dagger}c^{\;}_{i}$
(and their Hermitian conjugates) are
nilpotent, \((c_{i+1}^{\dagger}c^{\;}_{i})^2=0\).
However, unlike the elementary fermionic degrees of freedom,
bonds of the type
\(c_{i+1}^{\dagger}c^{\;}_{i}\) (with $1<i<N-1$) each fail
to commute with only three bonds;
\begin{equation}\label{bonds_fermions}
c_{i+1}^{\dagger}c^{\;}_{i}\ \ \ \ \mbox{fails to commute with}\ \ \ \
c^\dagger_ic^{\;}_{i-1},\ \ c^\dagger_{i}c^{\;}_{i+1},\ \ c^\dagger_{i+2}c^{\;}_{i+1}.
\end{equation}
Boundary bonds fail to commute with only two bonds.
(That is, $c_{2}^{\dagger} c^{\;}_{1}$ does not commute with $c^{\dagger}_{1} c^{\;}_{2}$ or
$c_{3}^{\dagger} c^{\;}_{2}$ and, similarly, $c_{n}^{\dagger} c^{\;}_{n-1}$ does not
commute with $c^{\;}_{n} c_{n-1}^{\dagger}$ nor $c^{\dagger}_{n-1} c^{\;}_{n-2}$.)
A model of spin \(S=1/2\) degrees of freedom that, we show next, is
closely related to $H_{\sf fermion}$ is the isotropic XY model
\begin{eqnarray}
\label{Hspin}
H_{\sf XY} = \lambda
\sum_{i=1}^{N-1} (\sigma_{i}^{+}\sigma_{i+1}^{-}+ \sigma_{i+1}^{+}\sigma_{i}^{-}),
\end{eqnarray}
with $\sigma^{\pm} = (\sigma^{x} \pm i \sigma^{y})/2$ which satisfy the Pauli algebra
\begin{eqnarray}
[\sigma_{i}^{+},\ \sigma_{j}^{-}] = \sigma_{i}^{z}\ \delta_{i,j},
~~~~~[\sigma_{j}^{z},\ \sigma_{i}^{\pm}] = \pm 2\sigma_{i}^{\pm}\ \delta_{i,j}.
\end{eqnarray}
As for the spinless fermions, each bond is nilpotent,
\((\sigma_{i+1}^{+}\sigma_{i}^{-})^2=0\). Similarly, bonds that share no common site commute, while
(for $1<i<N-1$)
\begin{equation}\label{bonds_spinons}
\sigma_{i+1}^{+}\sigma_{i}^{-}\ \ \ \ \mbox{fails to commute with}\ \ \ \
\sigma^+_{i}\sigma^-_{i-1},\ \sigma^+_{i}\sigma^-_{i+1},\ \
\sigma^+_{i+2}\sigma^-_{i+1} .
\end{equation}
Similar to the fermionic case, the boundary bonds (those with $i=1$ and $i=(N-1)$ above)
fail to commute with only two bonds.
A direct comparison of Equations \eqref{bonds_fermions} and \eqref{bonds_spinons},
together with the fact that both types of bonds are nilpotent, suggests that
the mapping
\begin{equation}\label{fermasd}
c_{i+1}^{\dagger}c^{\;}_{i}\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma_{i+1}^{+} \sigma_{i}^{-},
\end{equation}
(and the corresponding Hermitian-conjugate relation)
may define a duality isomorphism. The easiest way to check the
assertion is to
assume that this is indeed the case, and use \(\Phi_{{\sf d}}\) as
explained in Section \ref{sec3.5} to compute dual
variables. Then, when written in terms of dual variables, it is easy
to check that \(\Phi_{{\sf d}}\) defines a bond algebra isomorphism.
{\it These dual variables turn out to define the Jordan-Wigner
transformation.}
The starting point consists in writing \(c^\dagger_i\) in terms of bonds.
To this end, we need to {\it extend the bond algebra} by adding two more
generators, \(c^{\;}_1\) and \(c_1^\dagger\). We can then compute
\(c^{\;}_i,\ i=2,3,\cdots,N\), recursively from the relation
\begin{equation}\label{toiterate}
[c^{\;}_{i-1},\ c_{i-1}^\dagger c^{\;}_i]=c^{\;}_i,
\end{equation}
so that \(c_i\) can be written as a nested multiple commutator of bonds.
The next step is to extend the action of \(\Phi_{{\sf d}}\) to the new generators.
We propose
\begin{equation}\label{mapping_at_boundary}
c^\dagger_1\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^+_1, \ \ \ \ c^{\;}_1\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^-_1,
\end{equation}
to be consistent with Equation \eqref{fermasd}, and it follows that
we have to extend the bond algebra of the XY model as well
by adding the two generators \(\sigma^+_1,\ \sigma^-_1\).
Now we can proceed to compute the dual variables.
Equations \eqref{fermasd},
and \eqref{mapping_at_boundary} turn the recursion relation Equation \eqref{toiterate}
into
\begin{eqnarray}
c^{\;}_1&\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ & \hat{c}^{\;}_1=\sigma^-_1 \\
~[c^{\;}_{i-1},\ c_{i-1}^\dagger c^{\;}_i]=c^{\;}_i &\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ &
[\hat{c}^{\;}_{i-1},\ \sigma^+_{i-1}\sigma^-_i]=\hat{c}^{\;}_i,\nonumber
\end{eqnarray}
that can be solved to yield
\begin{equation}
c^{\;}_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \hat{c}^{\;}_i=\prod_{j=1}^{i-1}(-\sigma^z_j) \, \sigma^-_i , \ \
c^{\dagger}_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \hat{c}^{\dagger}_i=\prod_{j=1}^{i-1}(-\sigma^z_j) \, \sigma^+_i
\label{JWisod}
\end{equation}
(for instance, \(\hat{c}^{\;}_2=[\hat{c}^{\;}_{1},\
\sigma^+_{1}\sigma^-_2]=[\sigma^-_1,\
\sigma^+_{1}\sigma^-_2]=-\sigma^z_1\sigma^-_2\)).
{\it This is nothing else than the Jordan-Wigner transformation}.
In particular, the fermion number operator $n_i$ transforms as
\begin{equation}
n_i=c^\dagger_i c^{\;}_i\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \hat{c}^{\dagger}_i\hat{c}^{\;}_i=
\frac{\mathbb{1}+\sigma^z_i}{2} .
\end{equation}
The fact that the dual variables \({\hat{c}}^\dagger_i,\ {\hat{c}}^{\;}_i\) satisfy the
fermionic algebra of Equation \eqref{car} confirms that the mapping
\(\Phi_{{\sf d}}\) of Equations \eqref{fermasd} and \eqref{mapping_at_boundary}
defines an isomorphism of (extended) bond algebras.
We see that {\it fermionization is a process local in the bonds},
and that the non-local structure of the Jordan-Wigner transformation
has the same origin as that of other dual variables. Notice that one can define the
inverse duality isomorphism
\begin{equation}\label{fermasd}
\sigma^+_1\ \stackrel{\Phi^{-1}_{{\sf d}}}{\longrightarrow}\ c^\dagger_1,\ \ \ \
\sigma_{i+1}^{+} \sigma_{i}^{-} \ \stackrel{\Phi^{-1}_{{\sf d}}}{\longrightarrow}\ c_{i+1}^{\dagger}c^{\;}_{i},
\end{equation}
by inspection of Equation \eqref{JWisod}, and express spin variables in terms of
fermions
\begin{equation}\label{invJWisod}
\sigma^-_i\ \stackrel{\Phi^{-1}_{{\sf d}}}{\longrightarrow}\ \hat{\sigma}^-_i=\prod_{j=1}^{i-1}(1-2n_j) \, c^{\;}_i=
e^{i\pi\sum_{j=1}^{i-1}n_j} \ c^{\;}_i ,
\end{equation}
with the corresponding Hermitian-conjugate relation.
There is some flexibility in the definition of the Jordan-Wigner transformation
that reflects two independent facts. First, we could
have extended the bond algebra by adding the boundary bonds
\(c^{\;}_N,\ c^\dagger_N\) (and \(\sigma^-_N,\ \sigma^+_N\)), with the resulting
string operator running from site \(i+1\) to site \(N\) in Equations \eqref{JWisod} and
\eqref{invJWisod}. Second, the boundary term mapping
is only defined up to a phase term $\eta$. We could have set
\(c^\dagger_1\ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \eta \, \sigma^+_1\), where \(\eta\eta^*=1\), with the corresponding dual
variables displaying this overall phase.
Moreover, one can extend these ideas to establish a duality-based
derivation of generalized Jordan-Wigner transformations \cite{GJW}.
\subsubsection{Non-locality and fermionization in $d \geq 2$ dimensions}
We next ask whether nearest-neighbor spin and fermion models
can be related on arbitrary lattices. On general lattices on which
{\it closed loops} may be drawn, the bond algebras become richer and, in general, prohibit
such a mapping between the simplest choice of elementary bonds involving two
nearest-neighbor spins and two nearest-neighbor fermions.
The natural extension of the mapping
of Equation \eqref{fermasd} reads
\begin{equation}\label{fermasd_highd}
c^\dagger_{\bm{r}} c^{\;}_{{\bm{r}}'} \ \stackrel{\Phi_{{\sf d}} (?)}{\longrightarrow}\ \sigma^+_{\bm{r}} \sigma^-_{{\bm{r}}'},
\end{equation}
where the vectors \({\bm{r}}, {\bm{r}}'\) now label the endpoints of links on
{\em an arbitrary $d$-dimensional lattice}.
As we will show below, Equation \eqref{fermasd_highd} generally
fails to define a bond algebra isomorphism whenever the lattice contains
at least two closed loops \(C,\ C'\) that share a link.
To see this, let \({\bm{r}}_1,{\bm{r}}_2,\cdots, {\bm{r}}_N\equiv{\bm{r}}_1\)
label the sites in \(C\), and \({\bm{r}}_1',{\bm{r}}_2',\cdots, {\bm{r}}_M'\equiv{\bm{r}}_1'\) the
sites in \(C'\), listed in order of appearance
(for some orientation of the loops), and let
\begin{equation}
{\bm{r}}_1={\bm{r}}_1',\ \ \ \ \ \ \ \ \ {\bm{r}}_{N-1}={\bm{r}}_{M-1}' ,
\end{equation}
denote the endpoints of the shared link. Then, on one hand, we can compute
the following nested anti-commutator along \(C\),
\begin{eqnarray}\label{xy_loop_indep}
S_{{\bm{r}}_1,{\bm{r}}_{N-1}} &\equiv &
\{\{\{\cdots\{\{\sigma^+_{{\bm{r}}_1}\sigma^-_{{\bm{r}}_2},\ \sigma^+_{{\bm{r}}_2} \sigma^-_{{\bm{r}}_3}\},\
\sigma^+_{{\bm{r}}_3} \sigma^-_{{\bm{r}}_4}\}\cdots\},\ \sigma^+_{{\bm{r}}_{N-2}} \sigma^-_{{\bm{r}}_{N-1}}\},\
\sigma^+_{{\bm{r}}_{N-1}} \sigma^-_{{\bm{r}}_1}\} \nonumber \\
&=& \frac{1}{2}(1-\sigma^z_{{\bm{r}}_1}\sigma^z_{{\bm{r}}_{N-1}}),
\end{eqnarray}
and the outcome is in fact {\it independent of the loop} (the same computation along
\(C'\) would have returned the same result).
In contrast, the corresponding nested anti-commutator in terms of fermionic bonds
does depend on the loop,
\begin{eqnarray}
F_C &\equiv& \{\{\{\cdots\{\{c_{{\bm{r}}_1}^\dagger c^{\;}_{{\bm{r}}_2},\ c_{{\bm{r}}_2}^\dagger c^{\;}_{{\bm{r}}_3}\},\
c_{{\bm{r}}_3}^\dagger c^{\;}_{{\bm{r}}_4}\}\cdots\},\ c^\dagger_{{\bm{r}}_{N-2}} c^{\;}_{{\bm{r}}_{N-1}}\},\
c^\dagger_{{\bm{r}}_{N-1}} c^{\;}_{{\bm{r}}_1}\} \nonumber\\
&=& (n_{{\bm{r}}_{N-1}}-n_{{\bm{r}}_1})^2\prod_{m=2}^{N-2}(2n_{{\bm{r}}_m}-1)
\end{eqnarray}
with \(n_{\bm{r}}=c^\dagger_{\bm{r}} c^{\;}_{\bm{r}}\).
If we compute the same quantity along \(C'\), the result reads
\begin{equation}
F_{C'}=(n_{{\bm{r}}_{M-1}'}-n_{{\bm{r}}_1'})^2\prod_{m=2}^{M-2}(2n_{{\bm{r}}_m'}-1)=
(n_{{\bm{r}}_{N-1}}-n_{{\bm{r}}_1})^2\prod_{m=2}^{M-2}(2n_{{\bm{r}}_m'}-1) ,
\end{equation}
so that \(F_C\neq F_{C'}\). On the other hand, it follows from Equations
\eqref{fermasd_highd} and \eqref{xy_loop_indep} that
\begin{equation}
F_C, \ F_{C'}\ \stackrel{\Phi_{{\sf d}} (?)}{\longrightarrow}\ S_{{\bm{r}}_1,{\bm{r}}_{N-1}}.
\end{equation}
This shows that the mapping of Equation \eqref{fermasd_highd} cannot
define a bond algebra isomorphism (a fermionization of the
XY model in $d\geq 2$), since it defines a many-to-one mapping. It is a good illustration
of the type of problems that arise in attempts at fermionizing spin models
in more than one dimension, while preserving locality \cite{highdJW} and the
dimension of the state space \cite{verstraete_cirac}.
Thus, our circle of ideas has closed on itself. We see how a common (and slightly imprecise) lore, relating
the existence of closed loops with the impossibility of transmutation of statistics
in a direct physical way, is mathematically realized within the bond
algebraic approach in the above case.
One cannot exclude the existence of
other mappings between systems having longer range interactions
and/or being of a simpler form in other representations. For example,
examples in which more complicated spin interactions
(i.e., not those involving $S=1/2$ spins on nearest-neighbor sites)
may admit a transmutation of statistics.
Each proposed isomorphism that may
replace Equation \eqref{fermasd_highd} can be verified
(and more generally excluded) by examining the bond algebras of the
candidate dual systems. This is the subject of the next section. In particular,
we will describe a different kind of fermionization scheme that exploits
gauge-reducing dualities, thus relating two systems with {\it different number of
degrees of freedom}.
\subsubsection{Dualities and fermionization in
\(d\geq2\) dimensions}
The result of the last section is a rigorous manifestation of a
general obstruction to mapping nearest-neighbor
spin Hamiltonians to local fermionic models, and {\it viceversa},
in more than one spatial dimension \cite{highdJW},
while preserving the dimension of the (Hilbert) state space.
There are some interesting exceptions, however,
where the obstructions to higher dimensional fermionization
have been overcome.
These include the fermionization of (i) the POC model \cite{chen} of Section \ref{sec5.1},
(ii) the $d=2$ dimensional Kitaev's honeycomb spin $S=1/2$ model \cite{Kitaev2006,pachosannals}
(and extensions thereof) of Section \ref{sec3.6},
and (iii) some general related models discussed in Reference \cite{galitski}.
A mapping between
{\em local} fermionic models and local spin models is important in practice.
Some treatments \cite{verstraete_cirac,galitski}
invoke the need to add auxiliary fermions (spins) to map any local fermionic (spin)
model to a {\it sector of a local spin (fermionic)} model, that acts in general on a larger state
space.
Bond-algebraic dualities, that are always local transformations,
afford practical alternative (and more general)
approaches to the problem of fermionization in higher dimensions, that may
or may not involve a change in the number of degrees of freedom. First there
is the basic approach of looking directly for a bond-algebraic duality
that connects the fermion/spin model of interest to a dual spin/fermion model,
as we did at the beginning of this section. At times it is more
convenient to look for a duality of the model of interest to a dual model
that features the same type of degrees of freedom, but displays interaction terms
that are amenable to standard fermion/spin or spin/fermion transformations.
In either case, a change in the number of degrees of freedom may occur naturally
as a consequence of a gauge-reducing duality. This affords a natural picture
where local fermionization becomes possible, perhaps at the cost of introducing gauge
symmetries. We next illustrate this idea in \(d=2\) dimensions.
The quantum Ising model in \(d=2\) (see Equation \eqref{anyd_ising}) has a large, flexible
bond algebra generated by
\begin{equation}
\sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+{\bm{e_1}}},\ \ \ \ \sigma^z_{\bm{r}}\sigma^z_{{\bm{r}}+{\bm{e_2}}},
\ \ \ \ \sigma^x_{\bm{r}}.
\end{equation}
This bond algebra contains the bonds of most other \(d=2\) dimensional
spin models of interest (we already
exploited this fact in Section \ref{sec3.6} to find new dualities
for the Heisenberg model). It follows that if we can fermionize the Ising model,
we can translate that fermionization to many other spin models, including the Heisenberg model,
but it is well
known that any attempt to rewrite it in terms of fermionic operators (while
preserving the dimension of the state space) returns a {\it non-local} fermionic
Hamiltonian (this is to be expected in the light of the discussion of the
previous section). As explained in Section \ref{sec3.12},
the \(d=2\) dimensional Ising model is dual to the \(d=2\) dimensional
\({\mathbb{Z}}_2\) gauge theory defined in Equation \eqref{ising_gauge}. Seen in reverse,
the mapping of Equation \eqref{gauge_to_ising} that establishes this duality,
\begin{eqnarray}
\sigma^z_{{\bm{r}}-{\bm{e_2}}}\sigma^z_{\bm{r}} \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{({\bm{r}},1)},\ \ \ \ \ \
\sigma^z_{{\bm{r}}-{\bm{e_1}}}\sigma^z_{\bm{r}} \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ \sigma^x_{({\bm{r}},2)},\ \ \ \ \ \
\sigma^x_{\bm{r}} \ \stackrel{\Phi_{{\sf d}}}{\longrightarrow}\ B_{({\bm{r}},3)},
\label{gauge_to_ising_in_reverse}
\end{eqnarray}
can be understood as a prescription to turn the Ising model (and any other
model whose bonds can be written in terms of Ising bonds) into a model
with gauge symmetries, that is identical to the Ising model
{\it if projected onto the sector of gauge invariant states}.
{}From a different perspective, the duality of Equation
\eqref{gauge_to_ising_in_reverse}
introduces {\it in a natural way} one extra auxiliary spin (in the language
of References \cite{galitski,verstraete_cirac}) for each spin in
the original model.
The advantage of this approach is that {\it the dual \({\mathbb{Z}}_2\) gauge theory can be
fermionized} straightforwardly
by a Jordan-Wigner transformation, to read
\begin{eqnarray}\label{fermionic_gauge}
H_{\sf G}&=&\sum_{\bm{r}}\ \left[\bar{n}_{({\bm{r}},1)}+\bar{n}_{({\bm{r}},2)} \right.\\
&+&\lambda(c^\dagger_{({\bm{r}},1)}-c^{\;}_{({\bm{r}},1)})
\left.(c^\dagger_{({\bm{r}}+{\bm{e_1}},2)}+c^{\;}_{({\bm{r}}+{\bm{e_1}},2)}) (c^\dagger_{({\bm{r}}+{\bm{e_2}},1)}+c^{\;}_{({\bm{r}}+{\bm{e_2}},1)})
(c^\dagger_{({\bm{r}},2)}-c^{\;}_{({\bm{r}},2)})\right],\nonumber
\end{eqnarray}
where we have introduced \(\bar{n}_{({\bm{r}},\nu)}=1-2c^\dagger_{({\bm{r}},\nu)}c^{\;}_{({\bm{r}},\nu)}\),
that has the simple properties
\begin{equation}
\bar{n}^2_{({\bm{r}},\nu)}=1,\ \ \ \ \ \ \bar{n}_{({\bm{r}},\nu)}
c_{({\bm{r}}',\mu)}\bar{n}_{({\bm{r}},\nu)}=(1-2\delta_{{\bm{r}},{\bm{r}}'}\delta_{\nu,\mu})
c_{({\bm{r}}',\mu)}.
\end{equation}
In terms of fermions, the gauge symmetries (constraints) read
\begin{equation}
G_{\bm{r}}=\bar{n}_{({\bm{r}},1)}\bar{n}_{({\bm{r}},2)}\bar{n}_{({\bm{r}}-{\bm{e_1}},1)}\bar{n}_{({\bm{r}}-{\bm{e_2}},2)}.
\end{equation}
We can combine this result with the results of Section \ref{sec3.6} to
obtain a fermionization of the \(d=2\) quantum Heisenberg model.
Let us close this section with two additional examples of fermionization.
The fermionization
of Kitaev's honeycomb model \cite{Kitaev2006,pachosannals}, discussed
briefly near the end of Section \ref{sec3.6}, can be
achieved by (i) bond-algebraic techniques \cite{3detc}, while taking note of local
symmetries or (ii) via a special projective method from an extended Hilbert space
(Kitaev's original solution \cite{Kitaev2006}) or, alternatively, via a brute force high dimensional
Jordan-Wigner transformation that leads to a local Hamiltonian due to the presence of
the local symmetries of the model \cite{nussinov_chen}. Another new example is afforded
by the XM model of Section \ref{sec4.4}, that can be fermionized to read
\begin{eqnarray}
H_{\sf XM}&=&-\sum_{\bm{r}}\ \left [ h\ \bar{n}_{\bm{r}} \right .\\
&+&\left . J (c^{\;}_{{\bm{r}}-{\bm{e_1}}}+c^\dagger_{{\bm{r}}-{\bm{e_1}}})
(c^{\;}_{{\bm{r}}-{\bm{e_1}}+{\bm{e_2}}}-c^\dagger_{{\bm{r}}-{\bm{e_1}}+{\bm{e_2}}})
(c^{\;}_{{\bm{r}}}+c^\dagger_{{\bm{r}}})(c^{\;}_{{\bm{r}}+{\bm{e_2}}}-c^\dagger_{{\bm{r}}+{\bm{e_2}}}) \right ].\nonumber
\end{eqnarray}
\subsection{Self-dualities and quantum integrability}
\label{sec9.4}
Two self-dual models known to any physicist,
electromagnetism without sources and the quantum Ising chain,
happen to be integrable as well. Since the presence of an
exact self-duality
is a rather unusual property in itself, one is naturally tempted to
conjecture a connection between quantum integrability and self-duality.
We have seen by now more than enough examples of
self-dual models that are (most likely) {\it not} integrable
to grant that the previous argument is without force. That is not,
however, quite the end of the story. As it turns out, one of the
few known criteria for quantum integrability, the Dolan-Grady relations
\cite{dolan_grady},
is somewhat connected to self-duality, in that those relations
may in principle be more easily satisfied by self-dual models.
In fact, of the very few (all one-dimensional) models known to satisfy
these relations, the most relevant one is the quantum Ising chain.
Closer scrutiny, however, seems to make very clear that the Dolan-Grady
relations are essentially unrelated to self-duality. Let us explain
this more clearly.
Assume that a Hamiltonian can be partitioned
into two pieces $A$ and $B$,
\begin{eqnarray}
H=A+\lambda B,
\end{eqnarray}
such that \(A\) and
\(B\) satisfy the {\it Dolan-Grady relations}
\begin{eqnarray}\label{dg}
[A,[A,[A,B]]]=c[A,B],\ \ \ \ \ \ [B,[B,[B,A]]]=c[B,A],
\end{eqnarray}
with $c$ some c-number.
It was shown in \cite{dolan_grady}, in a remarkable {\it
tour-de-force}, that the constraints \eqref{dg} alone
suffice to guarantee that \(H\) is a member of a
family with an infinite number of conserved charges (which can furthermore be
explicitly written down in terms of \(A\) and \(B\)). Clearly,
if \(H\) is self-dual under an exchange of \(A\) and \(B\),
so that \(A=\mathcal{U}_{{\sf d}} B \, \mathcal{U}_{{\sf d}}^\dagger\), then
either Dolan-Grady relation implies the other.
In this sense, relations \eqref{dg}
may be in principle more easily satisfied by self-dual models.
But other than that, there is no reason to believe that
self-duality is related to quantum integrability in any deeper way.
First and foremost, the infinite set of conserved charges exist
whenever \eqref{dg} holds, independently of the presence
or absence of self-duality \cite{baseilhac}.
Moreover, a chain of coupled harmonic oscillators (\(d=1\)
phonons) afford an example of a model which is both exactly
solvable and self-dual, and yet it does not satisfy neither
relation in \eqref{dg}, which shows that the latter do not constitute
a necessary condition for integrability either. It is important to
notice also that the Ising chain has {\it two}
finite, self-dual renditions,
\begin{eqnarray}
A&=&\sum_{i=1}^{N-1}\sigma^z_i\sigma^z_{i+1}+\sigma^z_N,\ \ \ \ \ \
B=\sum_{i=1}^{N}\sigma^x_i;\nonumber\\
A&=&\sum_{i=1}^{N-1}\sigma^z_i\sigma^z_{i+1},\ \ \ \ \ \ \ \ \ \ \ \ \ \,
B=\sum_{i=1}^{N-1}\sigma^x_i,
\end{eqnarray}
{\it but only the latter satisfies the Dolan-Grady relations}.
It seems safe to conclude that self-duality and
quantum integrability are quite independent properties, a fact
that highlights once again
the importance of self-dualities as non-perturbative
probes of strongly-coupled models.
\subsection{Duality, topological quantum order, and dimensional reduction}
\label{sec_tqo_dim}
In recent years there has been much interest in topological quantum
order \cite{wenbook, toric_code,tqo}.
In topologically ordered systems, the state of the system cannot be characterized by
local measurements but rather by topological quantities
\cite{toric_code,wenbook}. One of the main hopes further driving interest
in those systems has been that systems characterized
by non-local order will be immune to local perturbations and thus quantum
information may be protected for sufficiently long times.
One of the most important properties of low temperature topological quantum
order is indeed its robustness to local perturbations \cite{toric_code,nsf}.
As demonstrated in \cite{tqo}, several models harboring that order
reduce, {\em via a bond algebraic duality}, to one-dimensional Ising chains with short-range interaction.
This suggested the possibility of short autocorrelation
times in these particular models,
and several new ones that have been devised and investigated since.
One of the consequences of this dimensional reduction borne by bond-algebraic dualities
is that these systems possess memory times that are finite (i.e., system size
independent) at all temperatures. This is so as the memory times $\tau$ gleaned from
(time $t$) autocorrelation functions such
as $\langle Z(0) Z(t) \rangle \sim \exp(-|t|/\tau)$ with $Z(t)$ a quantum bit (qubit) operator are reduced,
via a bond algebraic duality, to the autocorrelation function of a
local quantity in a lower dimensional system. As the dual lower dimensional system
remains ergodic at all non-zero temperatures (and thus harbors a finite autocorrelation time),
the memory time in the higher dimensional topologically ordered system
always remains finite as well. This phenomenon is known as {\it thermal fragility} \cite{3detc}.
There have been further suggestions that
measurement of non-local entanglement may detect (and, possibly, even quantify)
topological order
\cite{tee1,tee2}. These ideas seem exciting and some results
have recently appeared in one dimensional systems \cite{1dtqo}. We would like to point out, however,
that in general, there is no single measure of entanglement that uniquely characterizes
the quantum state of a system as exemplified by discussions of {\it generalized entanglement} \cite{barnum1}.
In systems that exhibit conventional ``Landau orders'', global symmetries link degenerate states to
one another (and, in particular, link the degenerate ground states to one another).
Let us label any such orthonormal ground states by
$\{ |g_{\alpha} \rangle \}$.
It was suggested in a series of works \cite{tqo,tqo1} that what
differentiates such topologically
ordered states from conventional states is the existence of ``$d$-dimensional gauge-like
symmetries'' \cite{tqo,batista_nussinov}. We briefly comment on particular aspects of these \cite{tqo,tqo1}.
Symmetry operators ($T_{\alpha \beta}$) may
connect topologically ordered ground states to one another,
\begin{eqnarray}
T_{\alpha \beta} | g_{\alpha} \rangle = | g_{\beta} \rangle.
\end{eqnarray}
The operators $\{T_{\alpha \beta} \}$ realize a group
that characterizes
classes of topologically ordered states. (Moreover, these symmetries ensure that
topological order exists at non-zero temperatures \cite{tqo,tqo1}.) In the jargon
of generalized entanglement \cite{barnum1},
the topologically ordered states are entangled relative to local observables.
Measures of entanglement such as topological entanglement entropy
\cite{tee1,tee2} are, on their own, insufficient for the
determination of topological quantum order \cite{tqo,tqo1}. The group theoretical
classification of $d$-dimensional gauge-like symmetries, however, does enable a natural framework
for the analysis of systems with topological order.
One would like, of course, to have a generalized order parameter that may ascertain the existence
of phase transitions in topologically ordered systems.
Dualities are of paramount importance in such systems as they enable the construction
of precisely such order parameters when they exist. In those cases its determination
is based
on the following maxim:
``Given a Hamiltonian (or corresponding classical action) that displays a phase transition,
there exists a dual Hamiltonian (or action) for which the phase transition is made
evident by the existence of an order parameter. That is, there exists a duality
that maps systems with topological quantum orders into systems with standard
(Landau type) order parameters that signal the breaking of a global symmetry when transitions
are present. In the original language, the dual variable (the order parameter
of the dual theory) may be non-local. As illustrated in \cite{tqo}, the correlations between
local quantities in one such dual basis can become, in the original basis,
non-local correlation functions (that lead to ``string'' or ``brane'' orders).
In prominent examples of topologically ordered systems such as Kitaev's toric
code model \cite{toric_code} and three-dimensional extensions \cite{3detc}
such a duality can be constructed as discussed
earlier (see also \cite{3detc,tqo,tqo1}). The system
is no longer entangled in the dual basis.
To make the discussion concrete, we will first consider the XXYYZZ
model \cite{chamonmodel,sergey}, then briefly comment on another
spin $S=1/2$ model on a honeycomb lattice \cite{ly}, and finally
discuss the three-dimensional Kitaev's TC model. Details concerning
the bond algebraic mappings leading to the results given below will
be presented in a forthcoming publication \cite{noc}.
The XXYYZZ model is a spin $S=1/2$ system on a face centered cubic (FCC) lattice.
As is well known, an FCC lattice can be viewed as comprised of all of the odd (or even) sites
of a cubic lattice. By even sites, we allude here to sites for which the
sum of the $x$, $y$, and $z$ coordinates (in units of
the lattice constant which we set to unity) is even. The basis
vectors along the Cartesian directions are $\hat{e}_{x,y,z}$.
The Hamiltonian \cite{chamonmodel, sergey} is of the form
\begin{eqnarray}
H = -J \sum_{u \in {\sf even}} h_{u},
\label{cham}
\end{eqnarray}
with interaction terms
given by
\begin{eqnarray}
\label{huxxyyzz}
h_{u} = S^{x}_{u-\hat{e}_{x}} S^{x}_{u+\hat{e}_{x}} S^{y}_{u-\hat{e}_{y}} S^{y}_{u+\hat{e}_{y}}
S^{z}_{u-\hat{e}_{z}} S^{z}_{u+\hat{e}_{z}}.
\end{eqnarray}
That is, the interaction terms are given by the product of spins over
all sites surrounding an even sublattice
site with the product being of the form ``XXYYZZ'', wherein the component of the spin appearing
in the product is determined by its relative location relative to the center of the octahedron formed by the
six sites.
By a bond algebraic mapping \cite{noc} {\em the $d=3$ dimensional XXYYZZ model can be
mapped into four decoupled ($D=1$) Ising spin
chains}. That is, the bond algebra satisfied by the interaction terms
$h_{u}$ of Equation \eqref{huxxyyzz} is identical to that of the bonds in four decoupled
classical Ising chains. Thus, a duality transformation implements
an exact dimensional reduction. A consequence of this mapping is that
the system remains ergodic at all temperatures $T>0$.
We next briefly comment on the honeycomb lattice model of Reference \cite{ly}.
In our language, that Hamiltonian is composed of a sum of two types of mutually commuting
bond operators for each minimal hexagonal plaquette. The square of each of the bond operators,
as well as the product of all bond operators (of each of the two types) over the entire lattice is
constrained to be one.
Replicating the analysis of \cite{con,3detc,tqo,tqo1}, this system
can be exactly mapped onto two decoupled classical Ising chains. As in the system
of \cite{chamonmodel}, this model exhibits finite, system size independent,
autocorrelation times at all temperatures.
Finally, we focus on the $d=3$ Kitaev's TC model defined in Reference
\cite{3detc}, which corresponds to Equation \eqref{KTCMd3} with $J_x=J_z$ and $h_x=h_z=0$.
As discussed in \cite{3detc,tqo}, the finite temperature partition function of this
system is given by the product of the partition functions
of a $D=1$ Ising chain and a $D=3$ Ising gauge model:
${\cal Z}_{{\sf Kitaev}, d=3}= {\cal Z}_{{\sf I}, D=1} \times {\cal Z}_{{\sf IG}, D=3}$.
Specifically, $[A_{{\bm{r}}}, B_{({\bm{r}},\nu\mu)}]=0$ and the bond algebras formed by the two sets
of operators are decoupled. Moreover, the bond algebra formed by the vertex operators $A_{{\bm{r}}}$ alone
is isomorphic to that of bonds in a classical nearest-neighbor Ising chain. The bond algebra formed by
the operators $B_{({\bm{r}},\nu\mu)}$ alone is identical to that of bonds (plaquette terms) in the classical $D=3$
Ising gauge theory (which is dual to the $D=3$ dimensional Ising model).
We now discuss how order in this topologically ordered system can be ascertained by non-local measures.
Below the critical temperature of the Ising gauge theory (which is dual to that of the Ising model),
the system exhibits non-trivial order as ascertained by the non-local asymptotic character of the Wilson
loop $W_{R} = \prod_{({\bm{r}},\nu\mu) \in R} B_{({\bm{r}},\nu\mu)}$ where the region $R$ is taken to be
arbitrarily large.
In the limit of large $R$, in the ordered low temperature phase, $W_{R}$ scales as
$W_{R} \sim \exp(-c_{1} |\partial R|)$ where $|\partial R|$ is the perimeter of the region
$R$ with $c_{1}$ a positive constant. (This asymptotic behavior of $W_{R}$
is known as a ``perimeter law''.)
By contrast, at high temperatures, $W_{R}$ satisfies for large $R$ an ``area law"
and scales as $\exp(-c_{2} |R|)$,
where $|R|$ denotes the number of plaquettes in $R$ (the area of $R$) and $c_{2}$
is also a positive constant.
These non-local measures delineating the high and low temperature phases
of the Ising gauge theory formed by the bonds $B_{({\bm{r}},\nu\mu)}$
(fleshed out by the Wilson loops, or an equivalent
kink operator formed in the quantum version of the model)
can be mapped onto local measures in the dual theory (a local on-site magnetization
in the $D=3$ classical Ising model or the $d=2$ quantum version, respectively).
|
\section{\@startsection {section}{1}{\z@}%
{-3.5ex \@plus -1ex \@minus -.2ex}%
{2.3ex \@plus.2ex}%
{\setcounter{equation}{0}\noindent\normalsize\bfseries\uppercase}}
\makeatother
\makeatletter
\renewcommand\subsection{\@startsection{subsection}{2}{\z@}%
{-3.25ex\@plus -1ex \@minus -.2ex}%
{1.5ex \@plus .2ex}%
{\noindent\normalsize\bfseries}}
\makeatother
\makeatletter
\renewcommand\subsubsection{\@startsection{subsubsection}{2}{\z@}%
{-3.25ex\@plus -1ex \@minus -.2ex}%
{1.5ex \@plus .2ex}%
{\noindent\normalsize\itshape}}
\makeatother
\let\oldfigure=\figure
\let\endoldfigure=\endfigure
\renewenvironment{figure}{
\begin{oldfigure}
\begin{center}
}{
\end{center}
\end{oldfigure}}
\newenvironment{figure2}{
\begin{figure*}
\begin{center}
}{
\end{center}
\end{figure*}}
\let\oldtable=\table
\let\endoldtable=\endtable
\renewenvironment{table}{
\begin{oldtable}
\begin{center}
\begin{footnotesize}
}{
\end{footnotesize}
\end{center}
\end{oldtable}}
\newenvironment{table2}{
\begin{table*}
\begin{center}
\begin{footnotesize}
}{
\end{footnotesize}
\end{center}
\end{table*}}
\citestyle{aa}
\let\oldthebibliography=\thebibliography
\let\endoldthebibliography=\endthebibliography
\renewenvironment{thebibliography}[1]{
\renewcommand\bibsection{\section*{References}}
\begin{oldthebibliography}{#1}
\setlength{\itemsep}{0pt}
\begin{small}
}{
\end{small}
\end{oldthebibliography}}
\newcommand{Paper~I\xspace}{Paper~I\xspace}
\newcommand{\ensuremath{\mathrm{d}}}{\ensuremath{\mathrm{d}}}
\newcommand{\U}[1]{\ensuremath{\mathrm{~#1}}}
\newcommand{\U{erg}}{\U{erg}}
\newcommand{\U{erg~s}}{\U{erg~s}}
\newcommand{\U{yr}}{\U{yr}}
\newcommand{\U{Myr}}{\U{Myr}}
\newcommand{\U{Gyr}}{\U{Gyr}}
\newcommand{\U{pc}}{\U{pc}}
\newcommand{\U{kpc}}{\U{kpc}}
\newcommand{\U{Mpc}}{\U{Mpc}}
\newcommand{\U{M}_{\odot}}{\U{M}_{\odot}}
\newcommand{\U{km\ s^{-1}}}{\U{km\ s^{-1}}}
\newcommand{H{\sc i} }{H{\sc i} }
\newcommand{H{\sc ii} }{H{\sc ii} }
\newcommand{\reft}[1]{Table~\ref{tab:#1}}
\newcommand{\reff}[1]{Figure~\ref{fig:#1}}
\newcommand{\tn}[1]{[\emph{T.N.}: #1]}
\newcommand{\todo}[1]{\par\noindent$\bullet$ \fbox{\parbox[t]{0.96\columnwidth}{To do: \texttt{#1}}}\par}
\newcommand{\fb}[1]{{\bf{#1}}}
\newcommand{\mc}[1]{\mathcal{#1}}
\newcommand{\e}[1]{\mathrm{e}^{#1}}
\begin{document}
\thispagestyle{empty}
\onecolumn
\noindent{\LARGE\bf On the dynamical evolution of globular clusters}
\vspace{0.7cm}
\noindent{\Large Michel~H\'enon$^{\star}$, translated by Florent~Renaud$^{1,2}$}
\vspace{0.2cm}
{\footnotesize \noindent \it
$^{\star}$ Institut d'Astrophysique, Paris (now at the Observatoire de Nice)\\
$^1$ Observatoire Astronomique and CNRS UMR 7550, Universit\'e de Strasbourg, 11 rue de l'Universit\'e, F-67000 Strasbourg, France\\
$^2$ Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge, CB3 0HA, UK\\
\phantom{$^2$} \emph{<EMAIL>}}\\
\noindent{\footnotesize Originaly published in October 1961; translated in October 2010}
\vspace{0.5cm}
\begin{flushright}
\begin{minipage}{12cm}
\setlength{\parindent}{15pt}
This paper is an English translation of Michel H\'enon's thesis, \emph{Sur l'\'evolution dynamique des amas globulaires} originally published in French in the Annales d'Astrophysique, Vol. 24, p.369 (1961).
Conventions and notations are as in the original version, for consistency. The English version is written so that it is as faithful to the French text as possible. The translator added some notes [\emph{T.N.}] for the sake of clarity, when required. Original French, English and Russian abstracts written by M.~H\'enon are available in the original version of the paper. The English part is reproduced below.
FR thanks Michel H\'enon for the enthusiasm and kindness he expressed when he was asked for permission to translate his work, as well as Douglas Heggie and Mark Gieles for their careful proofreading.
As the author explains himself on the cover page of the original version: \\
This work is the main thesis that M.M.~H\'enon presented to obtain the degree of \emph{Docteur \`es Sciences Physiques}, at the \emph{Falcult\'e des Sciences de l'Universit\'e de Paris}. The thesis has been defended in Paris, on 1961 December 11, before a jury composed of Messieurs Danjon (president of the jury), Schatzman and Delcroix. This work is not related to a series of papers previously published and entitled \emph{L'amas isochrone}.
\end{minipage}
\end{flushright}
\vspace{0.5cm}
\begin{multicols}{2}
\section*{Original abstract}
Chapter~I: The structure and evolution of globular clusters are closely linked; we propose here to study them simultaneously, with a purely theoretical approach. The essential hypotheses are: (1) spherical symmetry; (2) quasi-permanent regime; (3) isotropy of the velocities at all points; (4) mass equality.
Chapter~II: We establish the system of fundamentals equations (2.25). The cluster model is reduced to a canonic form by a homology transformation with time-dependent parameters. We will look for a model of invariable canonic form, that is a model which remains similar to itself while evolving.
Chapter~III: We show that the galactic field imposes the relation (3.8): the cluster's radius is proportional to the cube root of its mass.
Chapter~IV: Preliminary calculations show that the density must be infinite at the center of the cluster. We obtain the asymptotic expression (4.14) for the distribution function near the center. The conservation of mass imposes condition (4.32). The existence of an energy flux toward the cluster center is predicted.
Chapter~V: The ensemble of equations and conditions obtained in the preceding chapters forms a system which is resolved numerically by means of an electronic calculator. We find a unique solution: the ``homology model'' \tn{homologous model} (\reft{1}, Figures~\ref{fig:4}, \ref{fig:5}, \ref{fig:6}, \ref{fig:15}, \ref{fig:16}). Its mass and radius are finite. The outer radius is approximately 10 times the mean radius. The mass decreases linearly as a function of time. About one-third of the negative energy of the cluster is carried off by the stars which escape; the other two-thirds accumulate in the center, in multiple stars.
Chapter~VI: We introduce a small number of stars of different mass in the homology model and we calculate their distribution (Figures~\ref{fig:8} and \ref{fig:9}) and their escape rate (\reff{10}). We find that as their mass increases, they are more concentrated, and escape less rapidly. In particular, if their mass is greater than 3/2 the mass of the normal stars, they are nearly all collected near the center of the cluster, and their escape rate is zero.
Chapter~VII: We show that should the initial central density of a cluster be finite, it will grow and become infinite within a finite time (\reff{12}). We study next the evolution of a cluster differing slightly from the homology model: we find that the differences decreases, whatever their form. Thus the homology model is very probably the final state toward which the clusters tend.
\end{multicols}\twocolumn\noindent
Chapter~VIII: The theoretical results are compared with the observational data on globular clusters. First we construct an ``artificial cluster'' (Figures~\ref{fig:15} and \ref{fig:16}) which permits a general comparison. Then we compare the projected densities in detail, Sandage's star counts for the cluster M3, and the brightness measurements for 47~Tuc by Gascoigne \& Burr. The agreement is very satisfying (Figures~\ref{fig:18}, \ref{fig:19}, \ref{fig:21}). For the stars of M3, we find a definite mass-luminosity relation, which however is in disagreement with the relation indicated by stellar evolution theory (\reff{20}). The observed masses and radii are in good agreement with the theoretical relation (\reff{22}). The outer radii of the clusters are about twice as large as the observed radii. The absolute escape rate is constant and equals: $2.3\times 10^{-6} \U{M}_{\odot}/\U{yr}$. This constancy is confirmed by the observed distribution of globular cluster masses (\reff{23}). We obtain the initial mass distribution. Observation shows that $\omega$~Cen is the only globular cluster whose central density is not yet infinite. From this we deduce that the age of the globular clusters must lie between $24 \times 10^{9} \U{yr}$ and $44 \times 10^{9} \U{yr}$, in good agreemen [\emph{sic}] with the age deduced from stellar evolution.
Chapter~IX: We list a number of directions in which development of research is to be hoped for.
\section{Introduction}
The problem of the dynamical evolution of a globular cluster can be stated very simply. The only important force is the mutual attraction between the stars of the cluster; the other forces (like radiative pressure, electromagnetic forces, relativistic effects, etc.) are negligible. Therefore, the topic is the classical $n$-body problem: finding the motion of $n$ points of given masses, mutually attracting themselves as the inverse square of their distance.
This exposition, whereas simple, relates to an extremely arduous mathematical problem. Despite a large number of studies, it has not been possible to find an explicit solution, which very likely does not exist. Hence, one can think of the numerical integration. This way, \citet{vonHoerner1960} computed the evolution of artificial clusters comprising up to 16 stars, thanks to an electronic device. But high values of $n$ are out of reach with such a method, as the computational time becomes rapidly extreme, even for a machine; the case of $n=16$ already corresponds to a system of 96 simultaneous differential equations.
In the globular clusters, $n$ is of the order of magnitude of $10^{5-6}$. Such a high value naturally suggests to give up following the individual motions, and to use a statistical method. The structure of the cluster is then defined at all times thanks to a distribution function in a 7-dimensional space: 3 position coordinates, 3 velocity coordinates, and the mass. One can write a system of equations that allows, in principle, to calculate the evolution of this function, given its initial form. Unfortunately, the numerical resolution seems absolutely out of reach with this general form: indeed, one faces an integro-differential system with 8 independent variables!
Thus, it is necessary to make simplifications, or additional hypotheses. Two of them are classical and well valid:\\
(H1) we assume that the cluster is spherically symmetric;\\
(H2) we assume that the cluster has reached the steady state;\\
The number of independent variables is then reduced from 8 to 4 (see Equation 2.2 below); but the equations are still too involved. Therefore, one must make new simplifications, much more arbitrary, that are often only justified by their practical utility. The main one consists in considering the structure and the evolution problems separately. This is how a series of studies (e.g. \citealt{Plummer1911, Plummer1915, Eddington1916, Jeans1916, Chandrasekhar1942} Section 5.8, \citealt{Camm1952, Woolley1956, Henon1959}) has focussed on the structure of the clusters, without considering the evolution. In this case, the distribution function can take any form; it is set by an additional, arbitrary hypothesis, changing from one author to the other. The models obtained this way represent well the observed clusters; however this agreement does not mean much, because one of the fundamental equation of the problem as been suppressed, and replaced by an \emph{ad hoc} hypothesis.
Other works have, on the contrary, studied the evolution by assuming a known structure for the cluster (among others, see \citealt{Spitzer1940, Chandrasekhar1943b, Chandrasekhar1943c, King1958a, King1958b, King1958c, Spitzer1958b, vonHoerner1958, Agekian1959, Henon1960, King1960, Michie1961}). In most of the cases, it is assumed that the cluster is homogeneous and that the gravitational potential is uniform; the advantage of this is to get rid completely of the space variables. However, the results obtained (the major one being the escape rate of the stars from the cluster) often disagree from one author to the next (see \reff{10}), as a consequence of the arbitrary hypothesis, that has been introduced here again.
In fact, the two aspects, structure and evolution, cannot be told apart; they are intimately related. Equations show this clearly (see Equation 2.22 below). First solving the problem of the structure of the cluster, and then finding out its evolution, or the other way round, is not possible. All the equations have to be treated and solved as a whole.
We hope that the present work constitutes a first step toward this. The homologous model presented in Chapter~V is obtained by solving simultaneously the equations of structure and evolution. However, this model is far from being a satisfactory solution to the problem. Indeed, it has been necessary to keep two arbitrary hypotheses, to avoid a too high complexity in the calculations:\\
(H3) we assume that the distribution function only depends on the total energy (this is equivalent to assuming that the velocity distribution is everywhere isotropic);\\
(H4) we assume that all the stars have the same mass.
It is difficult to known how these simplifications affect the correctness of the results. Therefore, the conclusions should be taken carefully, and considered as illustrations instead of definitive statements. In fact, the main goal of this work is not to immediately obtain useful informations, but rather to set a more rigorous theory by getting rid of one of the arbitrary postulates it used to have, and by pointing at those that remain to be narrowed (see Chapter~IX).
One can ask whether it would be possible to make progress thanks to direct observations, and thus to circumvent or at least diminish, the difficulties of the theoretical work. Unfortunately, it seems that such a method is not very successful. Observing globular clusters is difficult, because of a number of reasons: the stellar density varies a lot from the center to the outskirts, so that it is impossible to get a satisfactory image of all the parts of the cluster on a single plate; one only observes the projected density and not the spatial density; measurements of individual proper motions are impossible to make; measurements of radial velocities are very imprecise, because of large distances. Worst, one observes only the brightest stars of the cluster; these stars only contribute to a small fraction of the total mass. In other words, the essential structure of the cluster remains invisible.
Observations of globular clusters bring too little information if one is to derive the theoretical model, characterized by a multi-variable function. This point can be illustrated thanks to the following quotes about the comparison between theoretical models and observed clusters: \tn{in english in the original version} ``the law ... represents the structure of a globular cluster with close approach to accuracy everywhere, with the exception of the region immediately surrounding the center'' \citep{Plummer1911}; ``this model appears to fit the variations of brightness of the observed disk'' \citep{Camm1952}; ``calculations with this model give projected densities which agree much better with observations than do those of the simple isothermal case'' \citep{Woolley1956}; ``with the first models we can obtain a rather fair fit'' \citep{vonHoerner1957}; ``l'amas isochrone ... ressemble d'une mani\`ere surprenante aux amas r\'eels'' \tn{the isochrone cluster ... looks surprisingly similar to the real clusters} \citep{Henon1959}. But each time, the model is different...
Thus, we can say that the agreement with observations is a necessary, but not sufficient condition for the validity of a theory on the dynamics of the clusters. Observations can be a final test for the theory, but not a starting point.
However, the theory can precisely progress in a purely deductive way, only owing to its own strengths. As we have seen, the physical problem can be translated into a system of equations describing it, if not perfectly rigorously, at least with a very good precision. This system encloses all the information required to solve the problem. That is, we feel that efforts should be made in this direction. Obviously difficulties exist; but they are only technical, mathematical difficulties, not conceptual ones. Thanks to the powerful computing machines nowadays, it does not seem impossible to solve the problem anymore.
\section{Equations}
\subsection{Definitions and hypotheses}
Let
\begin{equation}
\varphi(x,y,z,v_x,v_y,v_z,m,t)\,\ensuremath{\mathrm{d}} x\,\ensuremath{\mathrm{d}} y\,\ensuremath{\mathrm{d}} z\,\ensuremath{\mathrm{d}} v_x\,\ensuremath{\mathrm{d}} v_y\,\ensuremath{\mathrm{d}} v_z\,\ensuremath{\mathrm{d}} m
\end{equation}
be, at time $t$, the number of stars of mass between $m$ and $m+\ensuremath{\mathrm{d}} m$ and whose the six spacial and velocity coordinates lie between $x$ and $x+\ensuremath{\mathrm{d}} x$ ... $v_z$ and $v_z+\ensuremath{\mathrm{d}} v_z$. As we will see, this distribution function $\varphi$ fully describes the structure of the cluster.
We assume that the cluster is spherically symmetric and that it has reached a steady state or, strictly speaking, a quasi-steady state because evolution makes it change slowly \citep[see][]{Spitzer1940, Kuzmin1957}. One can show \citep{Jeans1915, Kurth1955} that the distribution function depends on the position and velocity coordinates only through two invariants (or strictly speaking, quasi-invariant):\\
$A$, the angular momentum of the star;\\
$E$, the total energy of the star, per unit mass (we will refer to this quantity as total energy, for conciseness).\\
Therefore, we have:
\begin{equation}
\varphi(x,y,z,v_x,v_y,v_z,m,t) = f(E, A, m, t).
\end{equation}
$E$ is given by the fundamental relation:
\begin{equation}
E= U + \frac{v^2}{2},
\end{equation}
where $U$ is the gravitational potential, and $v$ is the velocity of the star.
Here comes our first arbitrary hypothesis. We assume, for the rest of the paper, that $f$ does not depend on the angular momentum $A$. This is equivalent to assuming that the velocity distribution is isotropic (i.e. spherically symmetric) everywhere in the cluster, as shown in (2.3). Thus, the distribution function shrinks to:
\begin{equation}
f(E, m, t).
\end{equation}
Thanks to this hypothesis, one can have the following reasoning. The mutual perturbations of the stars tend to establish a Maxwellian, therefore isotropic, velocity distribution everywhere. The escape phenomenon goes, as we know, against such a distribution: every star rising above a certain critical velocity leaves the cluster (see Chapter~III). But this limitation only concern the modulus of the velocity and not its orientation; in other words, it is itself spherically symmetric. We can therefore assume that the Maxwellian equilibrium, which is impossible to establish in term of the distribution of the velocity moduli, does occur in term of the distribution of the orientations, i.e. the velocity distribution becomes isotropic everywhere, after a sufficient time.
But this reasoning is ambiguous: the stars travel across the entire cluster, and thus, it is forbidden to consider the evolution of a small region in isolation. Therefore, the validity of our hypothesis is not ensured\footnote{\citet{Michie1961} highlights an increase of the anisotropy in the external regions, for a particular model.}. We will examine the perspectives about this point in Chapter~IX.
\subsection{Equations of structure}
These equations are well-known. The density at a given point comes from the distribution function through:
\begin{equation}
\rho = \int\int\int\int m \varphi\, \ensuremath{\mathrm{d}} v_x\, \ensuremath{\mathrm{d}} v_y\, \ensuremath{\mathrm{d}} v_z\, \ensuremath{\mathrm{d}} m,
\end{equation}
or, using the isotropy of the velocities and (2.3):
\begin{equation}
\rho = 4\pi \int_0^{\infty} m\, \ensuremath{\mathrm{d}} m \int_U^{\infty} \left(2E-2U\right)^{1/2} f\, \ensuremath{\mathrm{d}} E.
\end{equation}
Furthermore, $\rho$, $U$ and the distance to the center $r$ are linked through Poisson's equation:
\begin{equation}
\frac{\partial^2 U}{\partial r^2} + \frac{2}{r} \frac{\partial U}{\partial r} = 4\pi G \rho.
\end{equation}
We note that $U$ depends on both $r$ and the time $t$ because of the evolution of the cluster: $U = U(r,t)$. In the same way: $\rho = \rho(r,t)$. The equations (2.6) and (2.7) give $\rho$ and $U$, i.e. the structure of the cluster, once the distribution function $f$ is known. It remains to write an essential equation: the one ruling the variation of $f$ with time.
\subsection{Local evolution equations}
If the potential of the cluster was perfectly regular, every star would keep its total energy forever and no evolution would occur. In reality however, the potential is created by a discrete distribution of masses and yields irregular features, that are random and change with time. As a consequence, the total energy of the star experiences perturbations. This question has often been addressed \citep[see e.g.][Section 2.1]{Chandrasekhar1942}; one showed that the star mostly experiences many small perturbations, whose cumulative effect leads to a slow, quasi-continuous, variation of the total energy of the star.
This total effect is proportional to $\ln{(l/l_1)}$, where $l$ is the maximum distance to the perturber stars, and $l_1$ is the ``impact parameter'', i.e. the distance by which two stars have to be separated to be deviated by an angle of 90$^\circ$. That is, even the very distant stars have a non-negligible effect. However, $l_1$ is very small: one can show that:
\begin{equation}
l_1 \simeq \frac{r_e}{n},
\end{equation}
where $r_e$ is the radius of the cluster. (Recall that $n$ is the total number of stars.) Consequently, the majority of the perturbations are caused by relatively close stars: e.g., if the cluster counts $10^5$ stars, $l_1$ equals $r_e \times 10^{-5}$, and 80\% of the perturbations would come from stars closer than $r_e \times 10^{-1}$. Therefore, in order to calculate the perturbations at a given point in the cluster, we will admit that the distribution of the velocities is the same in this point as everywhere else. This approximation probably leads to an error of about 10\%, which is acceptable given the current state of the theory.
The distribution function depends on $r$ and $v$, through $E$; to emphasize this, we set:
\begin{equation}
f(E, m, t) = f\left(U+\frac{v^2}{2}, m, t\right) = a(r, v, m, t).
\end{equation}
In this entire Section, we will study what happens at a given point in the cluster, at the distance $r$ from the center. The function $a$ defined in (2.9) describes the velocity distribution at this point.
\citet{Rosenbluth1957} gave the equation of evolution of such a distribution, with the effects of perturbations, in the very general case where there is no symmetry for the velocities, as well as in the case of axisymmetry. In our case of spherical symmetry for the velocities, their equation becomes (after some calculations)\footnote{This equation has recently been cited and used by \citet{King1960}.}:
\begin{eqnarray}
\left(\frac{\partial a}{\partial t}\right)_p & = & 16 \pi^2 G^2\ \ln{(n)} \int_0^{\infty} m_1\, \ensuremath{\mathrm{d}} m_1\\
& & \frac{1}{v^2}\frac{\partial}{\partial v}\bigg[m a \int_0^v a_1 v_1^2\, \ensuremath{\mathrm{d}} v_1 \nonumber\\
& & + \frac{m_1}{3} \frac{\partial a}{\partial v} \bigg( \frac{1}{v} \int_0^v a_1 v_1^4\, \ensuremath{\mathrm{d}} v_1 \nonumber \\
& & + v^2 \int_v^{\infty} a_1 v_1\, \ensuremath{\mathrm{d}} v_1 \bigg)\bigg]\nonumber
\end{eqnarray}
where, for simplicity, we set $a(r,v_1,m_1,t) = a_1$. The maximum distance $l$ to the perturber stars has been taken equal to the radius of the cluster. The subscript $p$ reminds us that the variation of $a$ is due to perturbations.
\subsection{Global evolution equations}
It is now time to focus on the entire cluster. Let's consider the subset of stars of given mass $m$ and energy $E$, at time $t$. These stars are situated at different locations in the cluster; but, according to our hypotheses, they all corresponds to the same value of $f$, thus of $a$. And yet, after a very short time, the functional form of $a$ has changed. This modification, given by (2.10), is different from one position in the cluster to another, in general. As a consequence, $a$ does not take the same value over the entire subset of stars anymore. In other words, the perturbations instantaneously destroy the steady-state regime.
However, the steady-state re-appears almost immediately, because of the rapid circulation of the stars within the cluster; the perturbations become equal for all the stars of the subset, so that $a$ takes the same value for all of them, once again.
Thus, the actual variation of $a$, after the equalization, is obtained when computing the derivative of (2.9), i.e.:
\begin{equation}
\frac{\partial a}{\partial t} = \frac{\partial f}{\partial E} \frac{\partial U}{\partial t} + \frac{\partial f}{\partial t},
\end{equation}
and we write that $\partial a / \partial t$, given in (2.11), is equal in average to $\left(\partial a / \partial t\right)_p$, given in (2.10).
Let $g(E,m,t)\, \ensuremath{\mathrm{d}} E\, \ensuremath{\mathrm{d}} m$ be the total number of stars in the subset. One has:
\begin{equation}
g(E, m, t)\, \ensuremath{\mathrm{d}} E\, \ensuremath{\mathrm{d}} m = \int_0^{r_m} 4\pi r^2\, \ensuremath{\mathrm{d}} r \cdot 4\pi a v^2\, \ensuremath{\mathrm{d}} v\, \ensuremath{\mathrm{d}} m,
\end{equation}
where the integral is with respect $\ensuremath{\mathrm{d}} r$. The maximum value $r_m$ taken by $r$ for a given energy $E$, is set by:
\begin{equation}
U(r_m,t) = E,
\end{equation}
and the velocity $v$ is given in (2.3), i.e.:
\begin{equation}
v = \left(2E-2U\right)^{1/2}.
\end{equation}
Equation (2.12) becomes:
\begin{equation}
g = 16\pi^2 \int_0^{r_m} r^2\, \ensuremath{\mathrm{d}} r \cdot av,
\end{equation}
and the number of stars in the subset changes as:
\begin{equation}
\frac{\partial g}{\partial t} = 16\pi^2 \int_0^{r_m} r^2\, \ensuremath{\mathrm{d}} r \left[v\frac{\partial a}{\partial t} + \frac{\partial (av)}{\partial v}\left(\frac{\partial v}{\partial t}\right)_E\right].
\end{equation}
We write that this integral remains the same whether using (2.10) or (2.11). The second term in the integral is the same in both cases; therefore we get:
\begin{equation}
\int_0^{r_m} r^2\, \ensuremath{\mathrm{d}} r\, v \left[\left(\frac{\partial a}{\partial t}\right)_p - \frac{\partial f}{\partial E}\frac{\partial U}{\partial t}-\frac{\partial f}{\partial t}\right] = 0,
\end{equation}
which is the equation of evolution we were looking for\footnote{This equation has already been presented, in a slightly different form however, by \citet[Equation 2.5]{Kuzmin1957}.}.
By changing $a$ and $a_1$ into $f$ and $f_1$ and using (2.14), it becomes:
\begin{eqnarray}
0 = \int_0^{r_m} r^2\, \ensuremath{\mathrm{d}} r \Bigg\{ 16\pi^2 G^2\ \ln{(n)} \int_0^{\infty} m_1\, \ensuremath{\mathrm{d}} m_1 \\
\frac{\partial}{\partial E} \bigg[ m f \int_U^E f_1 \left(2E_1-2U\right)^{1/2} \ensuremath{\mathrm{d}} E_1 \nonumber\\
+ \frac{m_1}{3}\frac{\partial f}{\partial E} \int_U^E f_1 \left(2E_1-2U\right)^{3/2} \ensuremath{\mathrm{d}} E_1 \nonumber\\
+ \frac{m_1}{3} \frac{\partial f}{\partial E} \left(2E-2U\right)^{3/2} \int_E^{\infty} f_1\, \ensuremath{\mathrm{d}} E_1 \bigg] \nonumber\\
- \left(2E-2U\right)^{1/2} \left(\frac{\partial f}{\partial E}\frac{\partial U}{\partial t}+\frac{\partial f}{\partial t}\right)\Bigg\}.\nonumber
\end{eqnarray}
We can switch the operations $\int_0^{r_m} \ensuremath{\mathrm{d}} r$ and $\frac{\partial}{\partial E}$, because the quantity between square-brackets becomes zero for $r=r_m$. Then, we switch the order of the integrations over $r$, $m_1$ and $E_1$, by recalling that $f$ and $f_1$ do not depend on $r$. Furthermore, we set:
\begin{equation}
\frac{1}{3}\int_0^{r_m} \left(2E-2U\right)^{3/2} r^2\, \ensuremath{\mathrm{d}} r = q(E,t),
\end{equation}
(\tn{The upper limit of the integral is missing in the original version}) and thus,
\begin{eqnarray}
\frac{\partial q}{\partial E} &=& \int_0^{r_m} \left(2E-2U\right)^{1/2} r^2\, \ensuremath{\mathrm{d}} r,\\
\frac{\partial q}{\partial t} &=& -\int_0^{r_m} \left(2E-2U\right)^{1/2} \frac{\partial U}{\partial t} r^2\, \ensuremath{\mathrm{d}} r.
\end{eqnarray}
After these transformations, (2.18) takes its final form:
\begin{eqnarray}
0 &=& \Bigg\{ 16\pi^2 G^2\ \ln{(n)} \int_0^{\infty} m_1\, \ensuremath{\mathrm{d}} m_1 \, \frac{\partial}{\partial E}\\
&& \bigg[ m f \int_{-\infty}^E f_1 q_1'\, \ensuremath{\mathrm{d}} E_1 + m_1 f' \Big( \int_{-\infty}^E f_1 q_1\, \ensuremath{\mathrm{d}} E_1 \nonumber\\
&& + q \int_E^{\infty} f_1\, \ensuremath{\mathrm{d}} E_1 \Big) \bigg] + f' \frac{\partial q}{\partial t} - q' \frac{\partial f}{\partial t} \Bigg\}\nonumber
\end{eqnarray}
(Hereafter, the symbol $'$ indicates the partial derivative with respect to $E$: $\frac{\partial f}{\partial E} = f'$, and so on.).
We note that the structure of the cluster is present in this equation only via the function $q$, defined in (2.19). This function will play a very important role in the following.
We now have the complete system of fundamental equations that rule the structure and the evolution of the cluster: these are the four equations~(2.6), (2.7), (2.19) and (2.22), linking together the four quantities $f$, $\rho$, $U$, $q$. But before actually solving this system, we first have to modify it.
\subsection{Equal masses}
In the rest of this paper (except in Chapter~VI), we limit ourselves to the study of the case where all the stars have the same mass. This is our second arbitrary hypothesis, less critical than the first one however. Indeed, whereas it is obvious that the masses of the stars are different in real clusters, one can imagine a fictitious cluster where all masses would be equal. In other words, the system we study corresponds to a physically plausible situation, while being simpler than the real systems.
Therefore, the distribution function reads:
\begin{equation}
f(E,m,t) = \delta(m-m_1)\cdot F(E,t),
\end{equation}
where $m_1$ is the mass of each star, and $\delta$ is the Dirac's distribution \tn{Dirac delta function}. Performing the integrations over $m$ in (2.6) and (2.22) is immediate.
\subsection{Normalized variables}
In order to get rid of the numerical constants, we define ``normalized variables'':
\begin{eqnarray}
\rho &=& 4\pi m_1\ D,\nonumber\\
r & = & (16\pi^2\ G\ m_1)^{-1/2}\ R,\nonumber\\
q & = & (16\pi^2\ G\ m_1)^{-3/2}\ Q,\nonumber\\
\ensuremath{\mathrm{d}} t & = & [16\pi^2\ G^2\ m_1^2\ \ln{(n)}]^{-1}\ \ensuremath{\mathrm{d}} T.
\end{eqnarray}
The four fundamental equations become:
\begin{eqnarray}
D &=& \int_U^{\infty} (2E - 2U)^{1/2} F \, \ensuremath{\mathrm{d}} E,\\
&& \frac{\partial^2 U}{\partial R^2} + \frac{2}{R}\frac{\partial U}{\partial R} = D,\nonumber\\
Q &=& \frac{1}{3}\int_0^{R_m} (2E - 2U)^{3/2} R^2\, \ensuremath{\mathrm{d}} R,\nonumber\\
0 &=& \frac{\partial}{\partial E} \bigg[ F \int^E_{-\infty} F_1 Q_1' \, \ensuremath{\mathrm{d}} E_1 + F' \Big( \int^E_{-\infty} F_1 Q_1 \, \ensuremath{\mathrm{d}} E_1 \nonumber\\
&& + Q \int_E^{+\infty}F_1 \, \ensuremath{\mathrm{d}} E_1 \Big) \bigg] + F'\frac{\partial Q}{\partial T} - Q' \frac{\partial F}{\partial T}. \nonumber
\end{eqnarray}
In addition, we set
\begin{equation}
F^{(-1)} = -\int_E^{\infty} F_1 \, \ensuremath{\mathrm{d}} E_1.
\end{equation}
\begin{equation}
S= F^{(-1)} \int_{-\infty}^E F_1 Q_1'\, \ensuremath{\mathrm{d}} E_1 - F \int_{-\infty}^E F_1^{(-1)} Q_1'\, \ensuremath{\mathrm{d}} E_1.
\end{equation}
We have
\begin{equation}
\frac{\partial S}{\partial E} = S' = F \int_{-\infty}^E F_1 Q_1'\, \ensuremath{\mathrm{d}} E_1 - F' \int_{-\infty}^E F_1^{(-1)} Q_1'\, \ensuremath{\mathrm{d}} E_1,
\end{equation}
i.e., thanks to a partial integration and by taking into account that $Q(-\infty) =0$:
\begin{eqnarray}
S' &=& F \int_{-\infty}^E F_1 Q_1'\, \ensuremath{\mathrm{d}} E_1 \\
&& + F' \left( \int_{-\infty}^E F_1 Q_1\, \ensuremath{\mathrm{d}} E_1 + Q \int_E^{\infty} F_1\, \ensuremath{\mathrm{d}} E_1 \right), \nonumber
\end{eqnarray}
so that (2.25d) can be re-written:
\begin{equation}
0 = S'' + F' \frac{\partial Q}{\partial T} - Q' \frac{\partial F}{\partial T}.
\end{equation}
\subsection{Canonization}
We know \citep[see e.g.][]{Kurth1955} that every theoretical cluster model includes two dimensional parameters; or, in other words: one can apply to the model an homology based on two arbitrary parameters. This can be easily checked with the equations: at a given time $T_0$, the cluster is completely defined by the distribution function $F(E, T_0)$. The most general homologous transformation would be to multiply $F$ on the one hand, and $E$ on the other, with two arbitrary constants. (We assume that the masses of the stars are set; thus, it is not possible to change them through a homology.)
It is interesting to use this possibility of homology to set any cluster model to a \emph{canonical form} defined by two chosen conditions. This allows one to highlight the very structure of the model, in a way that is independent of its size. The two conditions could be, for example: masses and radius equal to defined values. However, we will choose others, more convenient (see Chapter~IV, Equation 4.6).
In the case of a model in evolution, which we study here, this model will always be set to the canonical form thanks to a proper homology. Therefore, the two parameters of the homology will be functions of time. This has the great advantage to allow us to split two aspects of the evolution of the model: (1) the evolution of the size, represented by the variation of the parameters of the homology; (2) the evolution of the structure itself, dimensionless, represented by the variation of the canonical form.
Therefore, we set:
\begin{equation}
\begin{array}{l}
E = \beta\ \fb{E}\\
F = \gamma\ \fb{F},
\end{array}
\end{equation}
where $\beta$ and $\gamma$ are the two fundamental parameters of the homology, chosen so that the new model, defined by $\fb{F}(\fb{E})$, is canonical. By replacing this in the equations, we find out that the transformation formulae of the other quantities are:
\begin{eqnarray}
U &=& \beta\ \fb{U},\\
D &=& \beta^{3/2}\ \gamma\ \fb{D},\nonumber\\
R &=& \beta^{-1/4}\ \gamma^{-1/2}\ \fb{R},\nonumber\\
Q &=& \beta^{3/4}\ \gamma^{-3/2}\ \fb{Q},\nonumber\\
\ensuremath{\mathrm{d}} T &=& \gamma^{-1}\ \ensuremath{\mathrm{d}}\fb{T},\nonumber\\
S &=& \beta^{7/4}\ \gamma^{1/2}\ \fb{S}.\nonumber
\end{eqnarray}
One can easily check that the first three equations of the system (2.25) remain the same with the new variables. The fourth one changes however, because of derivatives with respect to time. From (2.31a) and (2.32e), we get:
\begin{equation}
\frac{\partial (\fb{E},\fb{T})}{\partial (E,T)} = \left| \begin{array}{cc}\displaystyle\frac{1}{\beta} & \displaystyle-\frac{1}{\beta^2}\frac{\ensuremath{\mathrm{d}}\beta}{\ensuremath{\mathrm{d}} T}E\\ \displaystyle 0 & \displaystyle\gamma \end{array}\right|,
\end{equation}
which allows us to compute:
\begin{equation}
\frac{\partial F}{\partial E}\frac{\partial Q}{\partial T}-\frac{\partial Q}{\partial E}\frac{\partial F}{\partial T} = \frac{\gamma}{\beta}\left(\frac{\partial F}{\partial \fb{E}}\frac{\partial Q}{\partial \fb{T}}-\frac{\partial Q}{\partial \fb{E}}\frac{\partial F}{\partial \fb{T}}\right).
\end{equation}
By use of (2.31b) and (2.32d), this becomes:
\begin{eqnarray}
\frac{\partial F}{\partial E}\frac{\partial Q}{\partial T}-\frac{\partial Q}{\partial E}\frac{\partial F}{\partial T} = \\
\beta^{-1/4}\gamma^{1/2} \Bigg[\frac{\partial \fb{F}}{\partial \fb{E}}\frac{\partial \fb{Q}}{\partial \fb{T}}-\frac{\partial \fb{Q}}{\partial \fb{E}}\frac{\partial \fb{F}}{\partial \fb{T}}+\fb{Q}\frac{\partial \fb{F}}{\partial \fb{E}}\big(\frac{3}{4}\frac{1}{\beta}\frac{\ensuremath{\mathrm{d}}\beta}{\ensuremath{\mathrm{d}}\fb{T}}\nonumber\\
- \frac{3}{2}\frac{1}{\gamma}\frac{\ensuremath{\mathrm{d}}\gamma}{\ensuremath{\mathrm{d}}\fb{T}} \big) - \fb{F}\frac{\partial \fb{Q}}{\partial \fb{E}}\frac{1}{\gamma}\frac{\ensuremath{\mathrm{d}}\gamma}{\ensuremath{\mathrm{d}}\fb{T}}\Bigg]. \nonumber
\end{eqnarray}
Let:
\begin{eqnarray}
\frac{1}{\beta}\frac{\ensuremath{\mathrm{d}}\beta}{\ensuremath{\mathrm{d}}\fb{T}} &=& b,\\
\frac{1}{\gamma}\frac{\ensuremath{\mathrm{d}}\gamma}{\ensuremath{\mathrm{d}}\fb{T}} &=& c.\nonumber
\end{eqnarray}
With the new variables, the system (2.25) becomes:
\begin{eqnarray}
\fb{D} &=& \int_{\fb{U}}^{\infty} (2\fb{E} - 2\fb{U})^{1/2} \fb{F} \, \ensuremath{\mathrm{d}}\fb{E},\\
&&\frac{\partial^2 \fb{U}}{\partial \fb{R}^2} + \frac{2}{\fb{R}}\frac{\partial \fb{U}}{\partial \fb{R}} = \fb{D},\nonumber\\
\fb{Q} &=& \frac{1}{3}\int_0^{\fb{R}_m} (2\fb{E} - 2\fb{U})^{3/2} \fb{R}^2\, \ensuremath{\mathrm{d}}\fb{R},\nonumber\\
\fb{S}' &=& \fb{F} \int_{-\infty}^{\fb{E}} \fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 + \fb{F}' \bigg( \int_{-\infty}^{\fb{E}} \fb{F}_1 \fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1\nonumber\\
&&+ \fb{Q} \int_{\fb{E}}^{\infty}\fb{F}_1 \, \ensuremath{\mathrm{d}}\fb{E}_1\bigg)\nonumber\\
0 &=& \fb{S}'' + \fb{F}' \frac{\partial \fb{Q}}{\partial \fb{T}} - \fb{Q'} \frac{\partial \fb{F}}{\partial \fb{T}} + \left(\frac{3}{4}b-\frac{3}{2}c\right)\fb{Q}\fb{F}-c\fb{F}\fb{Q}'.\nonumber
\end{eqnarray}
(The symbol $'$ indicates here the derivative with respect to $\fb{E}$, and not $E$ any more.)
\subsection{Secondary equations}
We write here the equations that allow us to compute several interesting quantities.
\subsubsection{Partial mass and partial kinetic energy}
Let $\mc{P}_E$ be the subset of stars whose total energy is less than $E$. The total mass of this subset $\mc{P}_E$ is, when using (2.14):
\begin{equation}
\mc{M} = \int_0^{\infty}m \, \ensuremath{\mathrm{d}} m \int_0^{r_m} 4\pi r^2\, \ensuremath{\mathrm{d}} r \int_U^E 4\pi (2E-2U)^{1/2} f_1\, \ensuremath{\mathrm{d}} E_1,
\end{equation}
and its total kinetic energy is
\begin{equation}
\mc{L} = \int_0^{\infty}m\, \ensuremath{\mathrm{d}} m \int_0^{r_m} 4\pi r^2\, \ensuremath{\mathrm{d}} r \int_U^E 2\pi (2E-2U)^{3/2} f_1\, \ensuremath{\mathrm{d}} E_1.
\end{equation}
Switching the order of the integrations and using (2.19) and (2.20), one obtains:
\begin{eqnarray}
\mc{M} &=& 16\pi^2 \int_0^{\infty}m\, \ensuremath{\mathrm{d}} m \int_{-\infty}^E f_1 q_1'\, \ensuremath{\mathrm{d}} E_1,\\
\mc{L} &=& 24\pi^2 \int_0^{\infty}m\, \ensuremath{\mathrm{d}} m \int_{-\infty}^E f_1 q_1\, \ensuremath{\mathrm{d}} E_1.\nonumber
\end{eqnarray}
We assume all the masses to be equal and we introduce the normalized variables defined as:
\begin{eqnarray}
\mc{M} &=& \left(16\pi^2\ m_1\right)^{-1/2} G^{-3/2}\ M,\\
\mc{L} &=& \left(16\pi^2\ m_1\right)^{-1/2} G^{-3/2}\ L,\nonumber
\end{eqnarray}
which results in:
\begin{eqnarray}
M &=& \int_{-\infty}^E F_1 Q_1'\, \ensuremath{\mathrm{d}} E_1,\\
L &=& \frac{3}{2}\int_{-\infty}^E F_1 Q_1\, \ensuremath{\mathrm{d}} E_1.\nonumber
\end{eqnarray}
These relations allow us to give a physical meaning to the integrals found in the evolution equation (2.25d). We also note that if we set $E = +\infty$, the subset $\mc{P}_E$ matches the entire cluster and the equations (2.40) and (2.42) give the total mass and the total kinetic energy of the cluster.
We switch to the canonical variables by setting:
\begin{eqnarray}
M &=& \beta^{3/4}\ \gamma^{-1/2}\ \fb{M},\\
L &=& \beta^{7/4}\ \gamma^{-1/2}\ \fb{L}.\nonumber
\end{eqnarray}
\subsubsection{Another partial mass}
The mass within the sphere of radius $r$ is
\begin{equation}
\mc{M}_r = \int_0^r 4\pi r_1^2\ \rho_1 \, \ensuremath{\mathrm{d}} r_1 = \frac{1}{G}r^2\ \frac{\partial U}{\partial r},
\end{equation}
which becomes, after transformations as in (2.41a) and (2.43a):
\begin{equation}
\fb{M}_\fb{R} = \int_0^\fb{R} \fb{D}_1 \fb{R}_1^2\, \ensuremath{\mathrm{d}}\fb{R}_1 = \fb{R}^2 \frac{\partial \fb{U}}{\partial \fb{R}}.
\end{equation}
One must be careful not to mix up $\fb{M}_\fb{R}$, the mass within the radius $\fb{R}$, with $\fb{M}$, the mass of the stars whose energy is less than $\fb{E}$.
\subsubsection{Projected density}
The projected density of the cluster, i.e. the mass observed per unit surface, is
\begin{equation}
\rho_P(r) = \int_{-\infty}^{+\infty} \rho\left(\sqrt{r^2+z^2}\right) \, \ensuremath{\mathrm{d}} z.
\end{equation}
We switch to the normalized, and then canonical, variables with
\begin{eqnarray}
\rho_P(r) &=& m_1^{1/2}\ G^{-1/2}\ D_P,\\
D_P(r) &=& \beta^{5/4}\ \gamma^{1/2}\ \fb{D}_P,
\end{eqnarray}
and we obtain
\begin{equation}
\fb{D}_P = \int_{-\infty}^{+\infty} \fb{D}\left(\sqrt{\fb{R}^2+\fb{Z}^2}\right) \, \ensuremath{\mathrm{d}}\fb{Z}.
\end{equation}
\subsubsection{Projected mass}
The mass enclosed within a circle of radius $r$ in the projection of the cluster is:
\begin{equation}
\mc{M}_P = \int_0^r 2\pi r_1\ \rho_P\, \ensuremath{\mathrm{d}} r_1,
\end{equation}
i.e., through canonical variables and transformation as in (2.41a) and (2.43a):
\begin{equation}
\fb{M}_P = \frac{1}{2} \int_0^\fb{R} \fb{D}_P \fb{R}\, \ensuremath{\mathrm{d}}\fb{R}.
\end{equation}
\subsection{Homologous evolution}
The system of equations (2.37) allows us to compute the evolution of the cluster from a given initial state. However, a few tests revealed that even with a powerful machine (IBM~704), this calculation is extremely long. Furthermore, the circumstances of the birth of a globular cluster (and even more, the initial energy distribution) remain unknown at present time. The choice of the function describing the initial state of the cluster is, therefore, quite arbitrary.
Consequently, it seems more rational to face the problem from the other side, and to first look for what the final evolution of the cluster will be. An analogy with gases suggests indeed that the stellar systems should tend toward a defined final state, independent of their initial state; and the similarity of the globular clusters as we observe them today \citep{vonHoerner1957} seems to confirm this assumption. This final state should be deduced from the system of equations (2.37) alone; it is the equivalent, in a way, to the Maxwellian distribution of velocities for a gas in equilibrium.
However, it is well-known that the final state of a cluster cannot be equilibrium. The stars which gain enough energy through perturbations escape forever from the cluster, whose total mass therefore decreases. Hence, a stationary state \emph{stricto sensu}, does not exist.
How then, can we imagine the final evolution of the cluster? It is time to go back to the distinction made above between the evolution of the size and the evolution of the structure. The escape of stars only implies the evolution of the size; it does not forbid the existence of a model for which only the dimensions would vary, while its structure would remain the same; in other words, a model where the evolution would only consist in ``expansions'' or ``contractions'' of the physical quantities. We call such a process \emph{homologous evolution}.
Therefore, it seems natural to seek whether the final evolution of the cluster would be of this kind. This is what we will do in the next Chapters. The answer will be yes: we will see first (Chapters~III to V) that the system (2.37) allows one and only one solution in homologous evolution, which we will name \emph{homologous model}; second we will show (Chapter~VII) that this model represents indeed the final state toward which all the clusters tend.
Writing of the above in mathematical terms is very simple. A model in homologous evolution allows an invariant canonical form: $\fb{F}$, $\fb{Q}$, and so on, are independent of the time. Thus, in particular:
\begin{equation}
\frac{\partial \fb{F}}{\partial \fb{T}} = 0, \qquad \frac{\partial \fb{Q}}{\partial \fb{T}} = 0,
\end{equation}
and the last equation of the system (2.37) shrinks to
\begin{equation}
0 = \fb{S}'' + \left(\frac{3}{4}b-\frac{3}{2}c\right)\fb{Q}\fb{F}'-c\fb{F}\fb{Q}'.
\end{equation}
This way, the system becomes independent of time.
Before focussing on the numerical solution, we still have to examine the boundary conditions. This will be the topic of the next two Chapters.
\section{Boundary conditions}
\subsection{Effect of the galactic field}
Globular clusters are not isolated systems: they are subject to the gravitational field of the galaxy, an effect that cannot be neglected.
Let $U_G$ be the galactic potential. The dimensions of a cluster being much smaller than those of the galaxy, this potential can be described with sufficient precision by a Taylor series truncated to the second order; in a reference frame centered on the cluster, moving with it, and with the right orientation, the series reads
\begin{equation}
U_G = U_G(0) + \frac{1}{2}\left(\frac{\partial^2 U_G}{\partial x^2}x^2 + \frac{\partial^2 U_G}{\partial y^2}y^2 + \frac{\partial^2 U_G}{\partial z^2}z^2\right).
\end{equation}
Furthermore, we almost have:
\begin{equation}
\frac{\partial^2 U_G}{\partial x^2} + \frac{\partial^2 U_G}{\partial y^2} + \frac{\partial^2 U_G}{\partial z^2} =0,
\end{equation}
because the density of the galaxy is very small in the regions where the globular clusters are found. As a consequence, at least one of these three main curvatures is negative. We assume that the one along $x$ is the most negative.
The potential created by the cluster is, in the external regions:
\begin{equation}
U_A = -\frac{G\mc{M}_e}{r},
\end{equation}
\tn{the subscript $A$ stands for \emph{amas}: cluster.}, where $\mc{M}_e$ is the total mass of the cluster. Thus, the total potential along the $x$-axis is (\reff{1}):
\begin{equation}
U = U_G + U_A = U_G(0) + \frac{1}{2}\frac{\partial^2 U_G}{\partial x^2}x^2 -\frac{G\mc{M}_e}{x}.
\end{equation}
It reaches a maximum for
\begin{eqnarray}
x_e = (G\ \mc{M}_e)^{1/3} \left(-\frac{\partial^2 U_G}{\partial x^2}\right)^{-1/3},\\
U_e = U_G(0) - \frac{3}{2}(G\ \mc{M}_e)^{2/3} \left(-\frac{\partial^2 U_G}{\partial x^2}\right)^{1/3}.\nonumber
\end{eqnarray}
\begin{figure}
\includegraphics[width=\columnwidth]{f01}
\caption{Galactic potential $U_G$, cluster potential $U_A$ and total potential $U_G+U_A$.}
\label{fig:1}
\end{figure}
It follows that every star that goes beyond this point escapes from the cluster forever. This implies that the stars that have a radial, oscillatory, trajectory along the $x$-axis cannot have a total energy greater than $U_e$; therefore, for these stars:
\begin{equation}
f = 0 \qquad\textrm{for } E \ge U_e.
\end{equation}
But, as we supposed that the distribution function depends only on $E$, (3.6) is, in fact, a general property: no star can have an energy higher $U_e$. In $U_e$, the distribution function has what \citet{Chandrasekhar1943d} calls an ``absorbing barrier'', and that we call here, in a more illustrative way, a ``leak''. We know that, in this case, the function $f$ is zero for $U_e$, but its derivative is not; the value of the latter is proportional to the escape rate of the stars \citep[Equations 25 and 33]{Chandrasekhar1943d}.
The galactic field has also the effect of modifying the potential inside the cluster. However, the equations (3.1) and (3.2) show that $U_G-U_G(0)$ is zero in average on a sphere centered on the cluster; as a consequence, in our approximation of spherical symmetry, the galactic field has no effect inside the cluster. Its effect reduces to the creation of a ``leak'' at $U_e$.
The boundary of the cluster is defined by the equation $U=U_e$; it is a closed surface, with a shape of a lemon, elongated along the $x$-axis (on which it has two conic points for $x=\pm x_e$). Therefore, the ``radius'' of the cluster is a function of the angle; but in the case of the spherical symmetry approximation, we will limit ourselves to considering its mean value. It is obtained from (3.3) and (3.4), by replacing the galactic potential $U_G$ with its mean value $U_G(0)$, which gives out
\begin{eqnarray}
r_e &=& \frac{G\ \mc{M}_e}{U_G(0)-U_e} \\
&=& \frac{2}{3} (G\ \mc{M}_e)^{1/3}\left(-\frac{\partial^2 U_G}{\partial x^2}\right)^{-1/3} = \frac{2}{3}x_e.\nonumber
\end{eqnarray}
This is this mean value that we will refer to as the radius of the cluster, from now on.
The relation (3.7) links the radius to the mass of the cluster. The curvature of the galactic field can be considered as constant: we know that the dynamical evolution of the galaxy is much slower than that of the clusters. It is true that the cluster travels along an orbit in the galaxy; but this motion is much faster than the evolution of the cluster, and we can use the mean value of the curvature of the field along the orbit in (3.7). Thus, the relation becomes
\begin{equation}
r_e \propto \mc{M}_e^{1/3}.
\end{equation}
\emph{Isolated cluster}
For future applications, we will also consider the case of an isolated cluster, which would experience no external gravitational field. Indeed, it seems that several systems (galaxies, galaxy clusters) are almost in this situation. In this case, at a sufficiently large distance from the center, the total potential reduces to $U_A$, given in (3.3). It does not yield a maximum but rather increases with distance. Therefore, escapes can occur only if the distribution function ranges up to $E=0$, corresponding to an infinite radius for the cluster. Furthermore, we find out from (2.20) and (3.3):
\begin{equation}
q' = \frac{\pi\ G^3\ \mc{M}_e^3}{8\sqrt{2}\ (-E)^{5/2}}.
\end{equation}
In order to have a finite total mass (which is given by Equation 2.40), the distribution function must decrease steeper than $(-E)^{3/2}$ for $E\to 0$. It follows that its derivative is zero at the leak point $E=0$, leading to an escape rate of zero. This result has already be shown by means of another method \citep{Henon1960}; it is confirmed by the behavior of the artificial clusters of \citet{vonHoerner1960} from which no stars escape, even after a time much longer than the relaxation time. Therefore, for an isolated cluster, one has:
\begin{equation}
\mc{M}_e = \textrm{cst}.
\end{equation}
\subsection{Mass-radius relation}
The two relations (3.8) and (3.10) can be represented by the unique form:
\begin{equation}
r_e \propto \mc{M}_e^\lambda,
\end{equation}
with $\lambda = 1/3$ for globular clusters (and for any system in a non-uniform external field in general) and $\lambda = \infty$ for isolated clusters. We will use this general form (3.11), so that the formulae obtained could be eventually applied to the case of isolated clusters; but all the numerical calculations, in the present paper, will be done for $\lambda = 1/3$.
Using (2.32c), (2.43a) and (2.36), we switch to the canonical variables, and the relation (3.11) becomes:
\begin{equation}
-\frac{b}{4}-\frac{c}{2}+\frac{\ensuremath{\mathrm{d}}\ln{(\fb{R}_e)}}{\ensuremath{\mathrm{d}}\fb{T}} = \lambda \left(\frac{3b}{4}-\frac{c}{2}+\frac{\ensuremath{\mathrm{d}}\ln{(\fb{M}_e)}}{\ensuremath{\mathrm{d}}\fb{T}}\right).
\end{equation}
If the model is in homologous evolution, $\fb{R}_e$ and $\fb{M}_e$ are independent of time, and we get:
\begin{equation}
0 = (3\lambda +1) b + (2-2\lambda)c.
\end{equation}
We see that the mass-radius relation translates into a definite relation between the two parameters of the homology.
\subsection{Choice of a reference level for the potential}
The potential $U$ is defined up to an additive constant. We set this constant by writing:
\begin{equation}
U_e = 0.
\end{equation}
Therefore, the potential is zero at the boundary of the cluster, negative inside, positive outside. The total energy of a star is always negative, $E=0$ corresponding to the escape threshold.
\section{Central conditions}
\subsection{Infinite central density}
\emph{A priori}, one could expect the density profile of the model we seek to have the classical shape of \reff{2} with, in particular, a well-defined central density; and our first attempts have been set to find this. However, the calculations of the evolution from such an initial state revealed an unexpected phenomenon: no matter how big the initial value of the central density is, it always increases, without limits. As an example, \reff{2} shows the spacial density computed at two successive times. The horizontal axis is normalized to the external radius; this way, we see that the increase only concerns the central region of the cluster\footnote{\citet{Michie1961} has also observed a flux of stars toward the center of his model.}. This increase cannot be diminished by means of an homologous transformation; indeed, one can measure it by computing the ratio between the central density and the mean density of the cluster, and this ratio does not change in an homology.
\begin{figure}
\includegraphics[width=\columnwidth]{f02}
\caption{Increase of the density in the central region of a cluster}
\label{fig:2}
\end{figure}
Therefore, we have been led to consider, for the final state of the cluster, a \emph{model with infinite central density}. This can seem physically absurd at first sight. However, one should not forget that the density considered here is a probability density. The cluster is not made of a continuous medium, but rather of separated particles. For the model to have a physical meaning, the integral of the probability density must be finite over any finite volume; and we will see it is indeed the case.
In order to better emphasize this important point, we have created, from a table of random numbers, an ``artificial cluster'' , that follows the projected density law
\begin{equation}
\rho_P \propto \frac{1}{r}.
\end{equation}
\begin{figure}
\includegraphics[width=\columnwidth]{f03}
\caption{Artificial cluster with infinite central density.}
\label{fig:3}
\end{figure}
This cluster, shown in \reff{3}, looks like a real galactic cluster; the fact that its theoretical density is infinite in its center is not visible, and if we would plot the density profile from \reff{3}, as one would do for real cluster, we would likely state that the curve shows a finite maximum in the center!
The functional form of (4.1) is indeed the one we will find near the center of the model (see Equation 5.9). Note that the number of stars inside a circle of radius $r$ is proportional to $r$, and thus tends toward zero when the circle becomes smaller and smaller.
In Chapter~VIII we will study in detail how this central singularity of the density arises. For the time being, we only focus on the final state, and thus we assume that the singularity exists.
\subsection{Asymptotic expressions near the center}
$\rho$ being infinite at the center, the function $f$ must become infinite too, as shown in (2.6). As a consequence, the first two terms of the equation (2.53), which are of the order of $\fb{F}\fb{Q}$, become negligible with respect to the first one which is of the order of $\fb{F}^2\fb{Q}$, and we have:
\begin{equation}
\fb{S}'' = 0,
\end{equation}
i.e., after integration,
\begin{equation}
\fb{S} = \alpha_1 + \alpha_2\ \fb{E},
\end{equation}
where $\alpha_1$ and $\alpha_2$ are two constants.
Let's assume first that $\alpha_1$ and $\alpha_2$ are zero. In this case, using (2.27) and (2.28), (4.3) reduces to
\begin{equation}
\frac{\fb{F}}{\fb{F}^{(-1)}} = \frac{\fb{F}'}{\fb{F}},
\end{equation}
which immediately integrates as
\begin{equation}
\fb{F} = C_1\ \e{-C_2\fb{E}},
\end{equation}
where $C_1$ and $C_2$ are two constants, necessarily positive, because $\fb{F}$ must be positive and increasing for $\fb{E} \to -\infty$.
This way, we find a Maxwellian distribution of the energies near the center, independently of the particular functional form of $\fb{Q}$, i.e. of the structure of the cluster. This result is in agreement with what one could expect: in the central regions, which are very dense, the perturbations are very active and almost establish the Maxwellian equilibrium\footnote{often called ``isothermal equilibrium'', which is a misuse of language, because the concept of temperature is meaningless in stellar dynamics.}. We are now to set the two conditions that define the canonical form of the cluster (see Equation 2.31 and the following): we assume that $\fb{E}$ and $\fb{F}$ have been normalized so that the constants $C_1$ and $C_2$ equal unity. The distribution function near the center is then:
\begin{equation}
\fb{F} = \e{-\fb{E}}.
\end{equation}
This choice of the canonical conditions has the advantage of simplifying the formulae and the calculations a lot.
From (4.6), using successively (2.37a), (2.37b), (2.37c), we easily find the expressions of the various quantities near the center:
\begin{eqnarray}
\fb{D} &=& \left(\frac{\pi}{2}\right)^{1/2} \e{-\fb{U}},\\
\fb{R} &=& \left(\frac{8}{\pi}\right)^{1/4} \e{\fb{U}/2},\nonumber\\
\fb{Q} &=& K_D\ \e{3\fb{E}/2},\nonumber\\
\bigg(K_D &=& \frac{16}{9\sqrt{3}\ (8\pi)^{1/4}} = 0.45841\bigg).\nonumber
\end{eqnarray}
Note that (4.7b) is obtained as the unique solution of the differential equation (2.37b) for which $\fb{D}$ is infinite at the center of the cluster. We have, near the center:
\begin{equation}
\fb{D} = \frac{2}{\fb{R}^2}.
\end{equation}
Also note that the potential is, near the center:
\begin{equation}
\fb{U} = \frac{1}{2}\ln{\left(\frac{\pi}{8}\right)} + 2 \ln{(\fb{R})}.
\end{equation}
Therefore, it tends logarithmically toward $-\infty$ in the center: the cluster yields an infinite ``potential well''. As a consequence, the energy of the stars can take any value from $\fb{E}=-\infty$ to $\fb{E}=0$.
Finally, we have
\begin{eqnarray}
\int_{-\infty}^{\fb{E}} \fb{F}_1\fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E_1} &=& 2 K_D\ \e{\fb{E}/2},\\
\int_{-\infty}^{\fb{E}} \fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E_1} &=& 3 K_D\ \e{\fb{E}/2}.\nonumber
\end{eqnarray}
Let's go back to (4.3) in the general case where $\alpha_1$ and $\alpha_2$ are not zero. For small values of $\alpha_1$ and $\alpha_2$, the solution of (4.3) can be expanded as:
\begin{equation}
\fb{F} = \e{-\fb{E}} + \alpha_1 Y_1 + \alpha_2 Y_2,
\end{equation}
where $Y_1$ and $Y_2$ are function yet to be determined. By setting this in (4.3) and using (4.7c), we obtain two differential equations for $Y_1$ and $Y_2$:
\begin{eqnarray}
Y_1^{(-1)} + 2 Y_1 + Y_1' &=& \frac{\e{-\fb{E}/2}}{3K_D},\\
Y_2^{(-1)} + 2 Y_2 + Y_2' &=& \frac{(1+ \fb{E})\ \e{-\fb{E}/2}}{3K_D},\nonumber
\end{eqnarray}
which can be solved as
\begin{eqnarray}
Y_1 &=& C_{11} \e{-\fb{E}} + C_{12} \fb{E}\ \e{-\fb{E}} - \frac{2}{3K_D} \e{-\fb{E}/2},\\
Y_2 &=& C_{21} \e{-\fb{E}} + C_{22} \fb{E}\ \e{-\fb{E}} + \frac{2}{3K_D} (5-\bf{E})\ \e{-\fb{E}/2}.\nonumber
\end{eqnarray}
The $C_{ij}$ are arbitrary constants. The terms $C_{ij}$ are, in fact, parasite solutions coming from a bad normalization of $\fb{F}(\fb{E})$; we can get rid of them by modifying this normalization. Then, $\fb{F}$ reads
\begin{equation}
\fb{F} = \e{-\fb{E}} + \frac{2}{3K_D} \left(5\alpha_2 - \alpha_1 - \alpha_2\ \fb{E}\right) \e{-\fb{E}/2}.
\end{equation}
The second term of this expansion becomes negligible with respect to the first one, for all $\alpha_1$ and $\alpha_2$, when $\fb{E}$ is small enough. As a consequence, the expansion (4.14) is valid near the center, not only for small $\alpha_1$ and $\alpha_2$, but for any values they can take. Thus, we have found the general solution of (4.3), i.e. the functional form of the distribution function, for $\fb{E}\to -\infty$.\footnote{The expression (4.14) can be considered as an expansion of $\fb{F}$ up to the second order in $\e{\fb{E}/2}$; however, it has been obtained using the expansion of $\fb{Q}$ up to the first order, (4.7c). This method is valid because $\fb{S}$ cancels when $\fb{F}$ takes the form (4.6), for all $\fb{Q}$'s. We easily find out that the expansion of $\fb{S}$ up to the $n$-th order depends on the expansion of $\fb{F}$ up to the $n$-th order, and on the expansion of $\fb{Q}$ up to the $(n-1)$-th order only. Therefore, it is possible to consider one term less in the expansion of $\fb{Q}$ than in those of $\fb{F}$.}
\subsection{Flux of matter and flux of energy toward the center}
We are going to calculate the flux of matter and the flux of energy through the surface $E=$ cst, in the six-dimensional space of positions-velocities. This calculation will allow us to give a physical meaning to the constants $\alpha_1$ and $\alpha_2$, and to specify the conditions the model must fulfill near the center of the cluster. Let's go back a second to the study of the local properties and let's consider, in one point in the cluster, the subset of stars whose velocity is smaller than a given $v$. The density of this population is
\begin{eqnarray}
\rho_v &=& \int_0^\infty m\, \ensuremath{\mathrm{d}} m \int_0^v 4\pi a_1\ v_1^2\, \ensuremath{\mathrm{d}} v_1\\
&=& m_1 \int_0^v 4\pi A_1\ v_1^2\, \ensuremath{\mathrm{d}} v_1,\nonumber
\end{eqnarray}
where $a$ is the local distribution function, defined in (2.9). In analogy with (2.23), we set
\begin{equation}
a = \delta(m-m_1) A.
\end{equation}
From this, and using (2.10), we find
\begin{eqnarray}
\left(\frac{\partial \rho_v}{\partial t}\right)_p &=& 64 \pi^3\ G^2\ m_1^2\ \ln{(n)} \\
&& \bigg[ A \int_0^v A_1\ v_1^2\, \ensuremath{\mathrm{d}} v_1 + \frac{1}{3} \frac{\partial A}{\partial v} \bigg(\frac{1}{v} \int_0^v A_1\ v_1^4\, \ensuremath{\mathrm{d}} v_1\nonumber\\
&& + v^2 \int_0^v A_1\ v_1\, \ensuremath{\mathrm{d}} v_1 \bigg) \bigg]\nonumber.
\end{eqnarray}
This quantity can be physically interpreted as the flux of stars, in velocity-space, through the surface $v=$ cst.
The total mass of the stars in the subset $\mc{P}_E$ (stars whose total energy is less than a given $E$) is (see Equation 2.38):
\begin{equation}
\mc{M} = \int_0^{r_m} 4\pi \rho_v\ r^2\, \ensuremath{\mathrm{d}} r,
\end{equation}
which becomes, after differentiation while keeping $E$ constant:
\begin{equation}
\left(\frac{\partial \mc{M}}{\partial t}\right)_p = 4\pi \int_0^{r_m} \left[ \left(\frac{\partial \rho_v}{\partial t}\right)_p + \frac{\partial \rho_v}{\partial v} \left(\frac{\partial v}{\partial t}\right)_E\right] r^2\, \ensuremath{\mathrm{d}} r.
\end{equation}
This is the flux of the stars through the surface $E= $ cst, in the six-dimensional space. The subscript $p$ of the first term can be removed, because the equalization of the perturbation occurs between stars of same energy, and does not affect the value of $\mc{M}$. Furthermore, from (2.14):
\begin{equation}
\left(\frac{\partial v}{\partial t}\right)_E = -(2E - 2U)^{-1/2}\ \frac{\partial U}{\partial t}.
\end{equation}
Substituing into (4.19), then changing the variables, switching the operations the same way as in Chapter~II, and using (2.19), (2.20) and (2.21), we obtain:
\begin{eqnarray}
\frac{\partial \mc{M}}{\partial t} &=& 16 \pi^2\ F \frac{\partial q}{\partial t} + 256 \pi^4\ G^2\ m_1^2\ \ln{(n)}\\
&& \bigg[ F \int_{-\infty}^E F_1\ q_1'\, \ensuremath{\mathrm{d}} E_1 + F' \bigg(\int_{-\infty}^E F_1\ q_1\, \ensuremath{\mathrm{d}} E_1\nonumber\\
&& + q \int_E^{\infty} F_1\, \ensuremath{\mathrm{d}} E_1 \bigg) \bigg].\nonumber
\end{eqnarray}
We notice that when using (2.40a), we can easily go from this relation to the fundamental equation of the evolution (2.22), which is thus proved again. On the other hand, it is not possible to go from (2.22) to (4.21), because an integration constant remains undefined; that is why we had to go back to the basic equations to establish (4.21).
We switch to normalized variables thanks to (2.24) and (2.41a), which gives, using (2.29):
\begin{equation}
\frac{\partial M}{\partial T} = S' + F \frac{\partial Q}{\partial T}.
\end{equation}
Let's compute, in the same way, the flux of energy through a surface $E=$ cst. At a point in the cluster, the subset of stars which velocity is less than $v$ has a total energy density
\begin{equation}
h_v = m_1 \int_0^v 4\pi A\ Ev^2\, \ensuremath{\mathrm{d}} v,
\end{equation}
whose variation is
\begin{equation}
\left(\frac{\partial h_v}{\partial t}\right)_p = \int_0^v 4\pi \left[ \left(\frac{\partial A}{\partial t}\right)_p E + A \frac{\partial U}{\partial t} \right] v^2 \, \ensuremath{\mathrm{d}} v.
\end{equation}
The total energy of the subset $\mc{P}_E$ is
\begin{equation}
\mc{H} = \int_0^{r_m} 4\pi h_v\ r^2 \, \ensuremath{\mathrm{d}} r.
\end{equation}
We will not give the details of the calculation, which is similar to the previous one, but slightly longer; one must do several integrations by parts. We set, in analogy with (2.41b):
\begin{equation}
\mc{H} = (16\pi^2\ m_1)^{-1/2}\ G^{-3/2}\ H,
\end{equation}
and we obtain the expression of the flux of total energy through the surface $E=$ cst:
\begin{equation}
\frac{\partial H}{\partial T} = ES' - S + EF\ \frac{\partial Q}{\partial T} - \int_{-\infty}^E F_1 \frac{\partial Q_1}{\partial T} \, \ensuremath{\mathrm{d}} E_1.
\end{equation}
It is interesting to note that by computing the derivative (4.22) and (4.27) with respect to $E$, we find:
\begin{equation}
\frac{\partial H'}{\partial T} = E\ \frac{\partial M'}{\partial T}.
\end{equation}
This relation can be proved more directly. Indeed, let $\ensuremath{\mathrm{d}}\mc{P}_E$ be the subset of stars whose total energy is between $E$ and $E+\ensuremath{\mathrm{d}} E$; the mass of this subset is: $\ensuremath{\mathrm{d}} M = M' \ensuremath{\mathrm{d}} E$, and its total energy is: $\ensuremath{\mathrm{d}} H = H' \ensuremath{\mathrm{d}} E$. Thus comes
\begin{equation}
H' = E M',
\end{equation}
and the relation (4.28). However, here again, it is not possible to use (4.22) and (4.28) to demonstrate (4.27), because an integration constant would remain undefined.
(The quantity $\mc{H}$, that we called ``total energy of the subset $\mc{P}_E$'', has been computed by simply adding the total energies $E$ of its members, which is not correct, because by doing this, we count twice the mutual potential energy of the stars of $\mc{P}_E$. However, this effect is negligible if $\mc{P}_E$ only counts a small fraction of the stars of the cluster, which is the case for $E\to -\infty$.)
Near the center, one can neglect the last term of (4.22), and the two last terms of (4.27), which are of the order of $\e{E/2}$; we also replace $S$ with its value, given in (4.3) and (2.32f), and we get
\begin{eqnarray}
\frac{\partial M}{\partial T} &=& \beta^{3/4}\ \gamma^{1/2}\ \alpha_2,\\
\frac{\partial H}{\partial T} &=& -\beta^{7/4}\ \gamma^{1/2}\ \alpha_1.\nonumber
\end{eqnarray}
Thus, the two constants $\alpha_1$ and $\alpha_2$ represent, up to a factor, the fluxes of mass and of energy toward the center of the cluster. This result is unexpectedly simple.
Note that we can switch to the canonical variables thanks to (2.32e) and (2.43); $H$ has still the same dimension as $L$. By neglecting once more the terms in $\e{E/2}$, we obtain:
\begin{eqnarray}
\frac{\partial \fb{M}}{\partial \fb{T}} &=& \alpha_2,\\
\frac{\partial \fb{H}}{\partial \fb{T}} &=& -\alpha_1.\nonumber
\end{eqnarray}
We are now going to examine the physical meaning of these fluxes. Let's focus first on the flux of mass. (4.30a) shows that, near the center, it takes a constant value, independent of $E$; this means that a flow of matter enters or exits the cluster (depending on the sign of $\alpha_2$) through the central singularity. The fundamental equations do not rule out such a flow; one can even, as we will see, create an infinite number of models obeying the equations and yielding a non-zero flow of matter in the center. However, from a physical point of view, such a flow obviously does not make sense. This leads us to write the additional condition
\begin{equation}
\alpha_2 = 0.
\end{equation}
The need of writing a separate condition for the mass conservation in the center can be more easily explained: at the center of the cluster, the quantities become infinite and the fundamental equations are meaningless; thus an additional condition is required for this particular point. For that matter, we have already noted, just above, that it is impossible to go directly from the equation of evolution (2.22) to the Equation (4.21) which states the conversation of the mass; one constant is missing, which is precisely the value of the central flow.
However, the models with a non-zero central flow are useful: we will see them again in Chapter~VII, while studying the initial stages of the evolution, during which the central density slowly increases until it becomes infinite. Then, the flow of matter toward the center exists and simply corresponds to the slow ``filling'' of the central part of the density profile (see \reff{2}).
Let's consider now the flux of energy, given by (4.30b). This flux is also constant near the center: therefore the center of the cluster creates or absorbs energy. One could think that, in analogy with what is true for the mass, the flow of energy must be zero in the center. But the numerical computation shows (see the next Chapter) that the previously written condition (4.32) completed the definition of the solution of the system of equations, which is now unique; and this solution corresponds to a value of $\alpha_1$ that is positive and non zero. Thus, we have to admit that a flow of energy toward the center exists; more precisely: \emph{the center of the cluster absorbs some negative energy}.
We will come back with more details to this strange and very interesting phenomenon in the Chapter~V, and we will see how it can be physically interpreted.
\subsection{Follow-up on the expansions near the center}
To prepare the numerical computation, it is useful to push further the expansions of the various quantities near the center. Taking the relation (4.32) into account and setting
\begin{equation}
-\frac{2\alpha_1}{3K_D} = K,
\end{equation}
the expansion (4.14) of $\fb{F}$ reduces to
\begin{equation}
\fb{F} = \e{-\fb{E}} + K\ \e{-\fb{E}/2}.
\end{equation}
By using successively the fundamental equations (2.37a), (2.37b), (2.37c), we find out the two-terms expansions:
\begin{eqnarray}
\fb{D} &=& \left(\frac{\pi}{2}\right)^{1/2} \left(\e{-\fb{U}} + 2\sqrt{2}\ K\ \e{-\fb{U}/2} \right),\\
\fb{R} &=& \left(\frac{8}{\pi}\right)^{1/4} \left(\e{\fb{U}/2} - \frac{K}{\sqrt{2}} \ \e{\fb{U}} \right),\nonumber\\
\fb{Q} &=& K_D \left( \e{3\fb{E}/2} - \frac{9\sqrt{3}}{8\sqrt{2}}\ K\ \e{2\fb{E}}\right),\nonumber
\end{eqnarray}
that make (4.7) more precise. From a remark made above, we can compute the expansion of $\fb{F}$ at the third order by putting in (2.53) the expansion of $\fb{Q}$ at the second order only, given by (4.35c). The last two terms of (2.53) must now be taken into account. We find:
\begin{eqnarray}
\fb{F}^{(-1)} &=& -\e{-\fb{E}} - 2K\ \e{-\fb{E}/2} + \left(\frac{3}{2}b - \frac{3\sqrt{3}}{8\sqrt{2}}K^2 \right),\nonumber\\
\fb{F} &=& \e{-\fb{E}} + K\ \e{-\fb{E}/2} + 0.
\end{eqnarray}
The third term in the expansion of $\fb{F}$ is zero.
We stop the expansions here because the calculation of the terms of higher order is much more involved: the third term of $\fb{D}$ is not constant but proportional to $(-\fb{U})^{1/2}$ and its coefficient depends on the entire function $\fb{F}$. Therefore, we will use the expansions (4.35) for $\fb{D}$, $\fb{R}$, $\fb{Q}$. Finally, we note the expansions of the two integrals:
\begin{eqnarray}
\int_{-\infty}^\fb{E} \fb{F}_1 \fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1 = \\
K_D \left[2\e{\fb{E}/2} + \left(1-\frac{9\sqrt{3}}{8\sqrt{2}}\right)K\ \e{\fb{E}} - \frac{3\sqrt{3}}{4\sqrt{2}}K^2\ \e{3\fb{E}/2} \right],\nonumber\\
\int_{-\infty}^\fb{E} \fb{F}_1 \fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 = \nonumber\\
\frac{3}{2}K_D \left[2\e{\fb{E}/2} + \left(1-\frac{3\sqrt{3}}{2\sqrt{2}}\right)K\ \e{\fb{E}} - \frac{\sqrt{3}}{\sqrt{2}}K^2\ \e{3\fb{E}/2} \right].\nonumber
\end{eqnarray}
\section{The homologous model}
\subsection{Summary of the equations}
The model that we propose to compute is defined by the set of equations and conditions (2.37a, b, c, d), (2.53), (3.6), (3.13), (4.34), (4.35b) obtained in the previous Chapters. Some transformations are yet necessary to set the equations in a form that most favors the numerical computation:
\begin{enumerate}
\item In the differential equation (2.37b), we will consider $\fb{U}$ as the independent variable. Furthermore, $\fb{R}$ varies rapidly near the boundary, and it is useful to switch to a new variable $\fb{Z}$ defined as:
\begin{equation}
\fb{R} = \frac{1}{\fb{Z}}.
\end{equation}
\item The integral (2.37c) is transformed by means of an integration by parts.
\item The equation (2.53) is replaced with its integrated form with respect to $\fb{E}$; we have seen in Chapter~IV that the integration constant is zero.
\end{enumerate}
All the equations, modified this way, and the boundary conditions are gathered below:
\emph{Equations:}
\begin{eqnarray}
\fb{D} &=& \int_\fb{U}^\infty (2\fb{E}-2\fb{U})^{1/2} \fb{F}\, \ensuremath{\mathrm{d}}\fb{E}, \\
\frac{\ensuremath{\mathrm{d}}^2 \fb{Z}}{\ensuremath{\mathrm{d}}\fb{U}^2} &=& -\fb{D} \left(\frac{\ensuremath{\mathrm{d}}\fb{Z}}{\ensuremath{\mathrm{d}}\fb{U}}\right)^3 \fb{Z}^{-4},\nonumber\\
\fb{R} &=& \fb{Z}^{-1},\nonumber\\
\fb{Q} &=& \frac{1}{3}\int_{-\infty}^{\fb{E}} (2\fb{E}-2\fb{U})^{1/2}\ \fb{R}^3\, \ensuremath{\mathrm{d}}\fb{U},\nonumber\\
0 &=& \bigg\{ \fb{F} \int_{-\infty}^{\fb{E}} \fb{F}_1 \fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1\nonumber\\
&& + \fb{F}' \bigg( \int_{-\infty}^{\fb{E}} \fb{F}_1 \fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1 - \fb{Q}\fb{F}^{(-1)} \bigg) \nonumber\\
&& + \left( \frac{3}{4}b-\frac{3}{2}c \right) \fb{F}\fb{Q}\nonumber\\
&& + \left( \frac{1}{2}c-\frac{3}{4}b \right) \int_{-\infty}^{\fb{E}} \fb{F}_1 \fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 \bigg\},\nonumber\\
0 &=& (3\lambda + 1)b + (2-2\lambda) c.\nonumber
\end{eqnarray}
\emph{Boundary conditions:}
\begin{eqnarray}
\fb{F}(0) &=& 0,\\
\fb{F}^{(-1)}(0) &=& 0,\nonumber\\
\fb{F} &\simeq& \e{-\fb{E}} + K \e{-\fb{E}/2} \quad \textrm{for } \fb{E} \to -\infty,\nonumber\\
\fb{R} &\simeq& \left(\frac{8}{\pi}\right)^{1/4} \left(\e{\fb{U}/2} - \frac{K}{\sqrt{2}} \e{\fb{U}}\right) \quad \textrm{for } \fb{U} \to -\infty.\nonumber
\end{eqnarray}
The unknown functions are: $\fb{F}(\fb{E})$, $\fb{Q}(\fb{E})$, $\fb{D}(\fb{U})$, $\fb{Z}(\fb{U})$, $\fb{R}(\fb{U})$, defined from $-\infty$ to 0; the unknown constants are: $b$, $c$, $K$. $\lambda$ equals 1/3.
\subsection{Numerical solving method}
The form of the system does not allow the solution to be computed directly; one has to proceed by trial and error. Experience leads us to adopt the following iterative method:
\begin{enumerate}
\item Choose a temporary form for the function $\fb{F}$, that fulfills (5.3a) and (5.3c), with a temporary value of $K$;
\item compute $\fb{D}$ using (5.2a);
\item compute $\fb{Z}$ by integrating (5.2b) from the center to the boundary; the initial conditions are given in (5.3d);
\item compute $\fb{Q}$ by means of (5.2c) and (5.2d);
\item choose temporary values for $c$ and $K$; compute $b$ using (5.2f);
\item integrate (5.2e), which is equivalent to a differential system of the fourth order, from the center to the boundary; the initial conditions are given in (5.3c), (4.36a) and (4.37);
\item in general, this integration gives the final values of $\fb{F}(0)$ and $\fb{F}^{(-1)}(0)$ that are different from zero; go back to point [5.], modify $c$ or $K$ and begin the integration again; grope around this way for the values of $c$ and $K$ until $\fb{F}(0)$ and $\fb{F}^{(-1)}(0)$ vanish;
\item go back to point [2.] with the new function $\fb{F}$.
\end{enumerate}
This way, we obtain a series of approximations for $\fb{F}$; we stop when two successive approximations are equal, up to the desired precision. In practice, the convergence is quite fast: the errors are divided by about 5 at each iteration.
The computation has been done thanks to the IBM 650 device of the Observatoire de Meudon; 8 hours of computation are required to get the solution with a precision of 1/1000.
\subsection{Results: structure}
The four fundamental functions $\fb{F}$, $\fb{D}$, $\fb{R}$, $\fb{Q}$ are given in \reft{1} and plotted in \reff{4}. The table covers the range $(-5,0)$; below $\fb{E} = -5$ (or $\fb{U} = -5$), the functions are represented with a sufficient precision by the expressions (4.34) and (4.35). The values found for the constants are
\begin{eqnarray}
K &=& -0.9499,\\
c &=& + 0.4078,\nonumber\\
b &=& -\frac{2}{3}c = - 0.2719.\nonumber
\end{eqnarray}
\begin{figure}
\includegraphics[width=\columnwidth]{f04}
\caption{Homologous model: the four fundamental functions $\fb{F}(\fb{E})$, the distribution function; $\fb{D}(\fb{U})$, the spacial density; $\fb{R}(\fb{U})$, the distance to the center; $\fb{Q}(\fb{E})$, see (2.19).}
\label{fig:4}
\end{figure}
\begin{table2}
\caption{}
\label{tab:1}
\begin{tabular}{r@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}l}
\multicolumn{2}{c}{$\fb{E}$ or $\fb{U}$} & \multicolumn{2}{c}{$\fb{F}$} & \multicolumn{2}{c}{$\fb{D}$} & \multicolumn{2}{c}{$\fb{R}$} & \multicolumn{2}{c}{$\fb{Q}$} & \multicolumn{2}{c}{$\fb{D}_P$} & \multicolumn{2}{c}{$\fb{M}_P$} \\
\hline
\hline
-5& & 136&.96 & 148&.86 & 0&.1097 & 0&.0002805 & 44&.18 & 0&.2988 \\
-4&.9 & 123&.38 & 133&.12 & 0&.1156 & &\phantom{000.}3280 & & \\
-4&.8 & 111&.11 & 118&.97 & 0&.1220 & &\phantom{000.}3834 & 38&.80 & 0&.3282 \\
-4&.7 & 100&.05 & 106&.26 & 0&.1286 & &\phantom{000.}4487 & & \\
-4&.6 & 90&.06 & 94&.83 & 0&.1357 & &\phantom{000.}5249 & 34&.02 & 0&.3603 \\
-4&.5 & 81&.05 & 84&.57 & 0&.1432 & &\phantom{000.}6148 & & \\
-4&.4 & 72&.91 & 75&.36 & 0&.1511 & &\phantom{000.}7198 & 29&.73 & 0&.3954 \\
-4&.3 & 65&.57 & 67&.10 & 0&.1595 & &\phantom{000.}8436 & & \\
-4&.2 & 58&.95 & 59&.69 & 0&.1684 & &\phantom{000.}9885 & 25&.90 & 0&.4337 \\
-4&.1 & 52&.98 & 53&.04 & 0&.1778 & 0&.001159 & & \\
-4& & 47&.60 & 47&.09 & 0&.1878 & &\phantom{00.}1360 & 22&.48 & 0&.4754 \\
-3&.9 & 42&.75 & 41&.76 & 0&.1984 & &\phantom{00.}1596 & & \\
-3&.8 & 38&.37 & 36&.99 & 0&.2097 & &\phantom{00.}1874 & 19&.42 & 0&.5208 \\
-3&.7 & 34&.43 & 32&.73 & 0&.2217 & &\phantom{00.}2202 & & \\
-3&.6 & 30&.87 & 28&.91 & 0&.2345 & &\phantom{00.}2588 & 16&.70 & 0&.5702 \\
-3&.5 & 27&.67 & 25&.51 & 0&.2480 & &\phantom{00.}3045 & & \\
-3&.4 & 24&.79 & 22&.47 & 0&.2625 & &\phantom{00.}3584 & 14&.28 & 0&.6239 \\
-3&.3 & 22&.19 & 19&.76 & 0&.2779 & &\phantom{00.}4222 & & \\
-3&.2 & 19&.86 & 17&.35 & 0&.2943 & &\phantom{00.}4977 & 12&.13 & 0&.6821 \\
-3&.1 & 17&.75 & 15&.20 & 0&.3119 & &\phantom{00.}5872 & & \\
-3& & 15&.86 & 13&.30 & 0&.3306 & &\phantom{00.}6933 & 10&.24 & 0&.7452 \\
-2&.9 & 14&.16 & 11&.60 & 0&.3507 & &\phantom{00.}8195 & & \\
-2&.8 & 12&.63 & 10&.10 & 0&.3721 & &\phantom{00.}9695 & 8&.562 & 0&.8133 \\
-2&.7 & 11&.25 & 8&.769 & 0&.3951 & 0&.01148 & & \\
-2&.6 & 10&.02 & 7&.593 & 0&.4198 & &\phantom{0.}1362 & 7&.091 & 0&.8868 \\
-2&.5 & 8&.905 & 6&.554 & 0&.4464 & &\phantom{0.}1617 & & \\
-2&.4 & 7&.908 & 5&.639 & 0&.4750 & &\phantom{0.}1922 & 5&.806 & 0&.9658 \\
-2&.3 & 7&.013 & 4&.834 & 0&.5058 & &\phantom{0.}2288 & & \\
-2&.2 & 6&.211 & 4&.128 & 0&.5391 & &\phantom{0.}2729 & 4&.688 & 1&.0504 \\
-2&.1 & 5&.491 & 3&.509 & 0&.5751 & &\phantom{0.}3260 & & \\
-2& & 4&.847 & 2&.969 & 0&.6142 & &\phantom{0.}3901 & 3&.722 & 1&.1407 \\
-1&.9 & 4&.269 & 2&.498 & 0&.6566 & &\phantom{0.}4679 & & \\
-1&.8 & 3&.753 & 2&.090 & 0&.7029 & &\phantom{0.}5623 & 2&.894 & 1&.2884 \\
-1&.7 & 3&.290 & 1&.736 & 0&.7535 & &\phantom{0.}6774 & & \\
-1&.6 & 2&.877 & 1&.431 & 0&.8089 & &\phantom{0.}8183 & 2&.192 & 1&.3371 \\
-1&.5 & 2&.508 & 1&.170 & 0&.8699 & &\phantom{0.}9915 & & \\
-1&.4 & 2&.179 & 0&.9469 & 0&.9372 & 0&.1205 & 1&.605 & 1&.4420 \\
-1&.3 & 1&.885 & 0&.7576 & 1&.012 & &\phantom{.}1470 & & \\
-1&.2 & 1&.623 & 0&.5981 & 1&.095 & &\phantom{.}1801 & 1&.123 & 1&.5497 \\
-1&.1 & 1&.390 & 0&.4648 & 1&.189 & &\phantom{.}2216 & & \\
-1& & 1&.182 & 0&.3545 & 1&.294 & &\phantom{.}2741 & 0&.7369 & 1&.6578 \\
-0&.9 & 0&.9978 & 0&.2642 & 1&.414 & &\phantom{.}3410 & & \\
-0&.8 & 0&.8337 & 0&.1913 & 1&.552 & &\phantom{.}4273 & 0&.4392 & 1&.7622 \\
-0&.7 & 0&.6879 & 0&.1336 & 1&.712 & &\phantom{.}5396 & & \\
-0&.6 & 0&.5583 & 0&.08888 & 1&.900 & &\phantom{.}6880 & 0&.2232 & 1&.8571 \\
-0&.5 & 0&.4429 & 0&.05534 & 2&.125 & &\phantom{.}8873 & 0&.1442 & \\
-0&.4 & 0&.3397 & 0&.03128 & 2&.400 & 1&.161 & 0&.08554 & 1&.9319 \\
-0&.3 & 0&.2468 & 0&.01514 & 2&.744 & 1&.546 & 0&.04268 & 1&.9646 \\
-0&.2 & 0&.1619 & 0&.00550 & 3&.193 & 2&.107 & 0&.01526 & 1&.9861 \\
-0&.1 & 0&.0817 & 0&.00098 & 3&.805 & 2&.968 & 0&.00235 & 1&.9956 \\
0& & 0& & 0& & 4&.703 & 4&.384 & 0& & 1&.9968 \\
\hline
\end{tabular}
\end{table2}
Furthermore, the interesting physical quantities take the following values:
\begin{eqnarray}
\textrm{external radius:} & \fb{R}_e & = 4.703,\\
\textrm{total mass:} & \fb{M}_e & = 1.996,\nonumber\\
\textrm{total kinetic energy:} & \fb{L}_e & = 1.423.\nonumber
\end{eqnarray}
\begin{figure}
\includegraphics[width=\columnwidth]{f05}
\caption{Homologous model: spacial density and projected density as a function of the radius.}
\label{fig:5}
\end{figure}
The spacial density $\fb{D}$ is plotted in \reff{5} as a function of the radius. It increases very rapidly toward the center, as expected from (4.8). The structure of the model is perhaps better rendered in \reff{6}, which shows the mass fraction $\fb{M}_\fb{R}$ enclosed in a sphere of radius $\fb{R}$ (see Equation 2.45). Near the center, $\fb{M}_\fb{R}$ is proportional to $\fb{R}$. We note that half of the total mass is enclosed inside the radius $\fb{R} = 0.6800$, i.e. only about 1/7 of the external radius.
\begin{figure}
\includegraphics[width=\columnwidth]{f06}
\caption{Homologous model: mass enclosed in a sphere of radius $\fb{R}$.}
\label{fig:6}
\end{figure}
The potential $\fb{U}$ vanishes at the boundary of the cluster, according to our conventions; outside the cluster, its form is obtained by integrating (2.45):
\begin{equation}
\fb{U} = \fb{M}_e \left(\frac{1}{\fb{R}_e} - \frac{1}{\fb{R}}\right) \qquad \textrm{for } \fb{R} > \fb{R}_e.
\end{equation}
In particular, for $\fb{R} \to \infty$, the potential tends toward:
\begin{equation}
\fb{U}_\infty = \frac{\fb{M}_e}{\fb{R}_e} = 0.4243.
\end{equation}
(This is the potential created by the cluster only.)
The projected density, computed from (2.49) is given in \reft{1}, column 6, and plotted in \reff{5} as a function of the distance to the center (see also \reff{15} and \reff{16}). Near the center, from (4.35a) et (4.35b), the spatial density is expressed as a function of the radius as:
\begin{equation}
\fb{D} = \frac{2}{\fb{R}^2} + \frac{(8\pi)^{1/4}\ K}{\fb{R}}.
\end{equation}
We find
\begin{equation}
\fb{D}_P = \frac{2\pi}{\fb{R}} - 2(8\pi)^{1/4}\ K\ \ln{(\fb{R})} + K_P \qquad \textrm{for } \fb{R} \to 0,
\end{equation}
where $K_P$ is a constant which depends on the entire function \fb{D}. By linking the formula (5.9) with the values of \reft{1}, one finds $K_P \simeq -3.82$.
The projected mass (mass enclosed, in projection, inside a circle of radius $\fb{R}$), computed from the formula (2.51), is given in \reft{1}, column 7. In particular, we find out the value of the \emph{median radius} $\fb{R}_0$ of the cluster, defined as the radius of the circle which contains, in projection, half of the total mass. This quantity has the advantage of being easily measured for real clusters, while the external radius is, on the contrary, almost impossible to observe. We find
\begin{equation}
\fb{R}_0 = 0.4997.
\end{equation}
We note that this median radius $\fb{R}_0$ is about 10 times smaller than the external radius $\fb{R}_e$.
\subsection{Evolution}
When integrating (2.36), we get
\begin{eqnarray}
\beta &=& \beta_0\ \e{b\bf{T}},\\
\gamma &=& \gamma_0\ \e{c\bf{T}},\nonumber
\end{eqnarray}
The ``time'' $\fb{T}$ is defined by the differential equation (2.32e); it is not proportional to the physical time $T$. It is, somehow, the ``proper time'' of the cluster; its variation is measured with a scale which is proper to the cluster, and which always varies according to the evolution. To avoid any confusion, it is preferable to consider $\fb{T}$ as a simple parameter which measures the level of evolution of the cluster, as it is in (5.11).
The relation between $\fb{T}$ and the physical time $T$, found from (2.32e) and (5.11b), is
\begin{equation}
\fb{T} = -\frac{1}{c} \ln{(1-\gamma_0\ c\ T)},
\end{equation}
(by setting the origin of time at $T=0$ for $\fb{T}=0$). This relation is plotted in \reff{7}. We see that the evolution does not last forever, but rather ends abruptly at a time $T_1$, given by
\begin{equation}
T_1 = \frac{1}{\gamma_0 c}.
\end{equation}
\begin{figure}
\includegraphics[width=\columnwidth]{f07}
\caption{Homologous model: evolution of the parameter $\fb{T}$, as a function of time.}
\label{fig:7}
\end{figure}
For $T=T_1$, the mass of the cluster becomes zero, as we will see below; it is therefore the time when the cluster disappears, after the escape of the last stars.
However, the curve extends forever in the past: the age of the cluster can be anything, it is not possible to give it an upper limit.
When introducing (5.12) in (5.11), and taking (5.4c) into account, we obtain the variation laws of the two parameters of the homology:
\begin{eqnarray}
\beta &=& \beta_0 \left(1-\frac{T}{T_1}\right)^{2/3},\\
\gamma &=& \gamma_0 \left(1-\frac{T}{T_1}\right)^{-1}.\nonumber
\end{eqnarray}
We derive, from (2.32), (2.43), (2.48), the variations of the various physical quantities as functions of time. In particular, for the total mass, we get
\begin{equation}
M_e = M_{e_0} \left(1-\frac{T}{T_1}\right),
\end{equation}
which shows that \emph{the mass decreases linearly with time}. The absolute escape rate is thus constant.
The radius is proportional to $(1-T/T_1)^{1/3}$, thus decreases quite slowly. The density is constant, and as a consequence, the orbital period of the stars within the cluster is also constant. The velocities decrease as $(1-T/T_1)^{1/3}$. The total energy of the cluster decreases (in absolute value) as $(1-T/T_1)^{5/3}$, thus faster than the mass.
\subsection{Accumulation of negative energy in the center}
As we have seen in the previous Chapter, a non-zero value of $K$ implies the existence of a flow of energy toward the center. This energy cannot vanish; we must admit that it constitutes an energy reservoir.
Thus, the total energy of the cluster is always made of two parts:
\begin{itemize}
\item a ``point'' energy $H_1$, accumulated in the center of the cluster;
\item a ``diffuse'' energy $H_2$, distributed within the entire cluster.
\end{itemize}
Let's look at how these two fractions of the energy vary with time. The variation of $H_1$ equals the flux of energy toward the center, i.e., from (4.30b) and (4.33):
\begin{equation}
\frac{dH_1}{dT} = \beta^{7/4}\ \gamma^{1/2}\ \frac{3}{2} K_D\ K.
\end{equation}
The diffuse energy $H_2$ is, from the virial theorem:
\begin{equation}
H_2 = -L_e,
\end{equation}
where $L_e$ is the total kinetic energy of the cluster; this relation remains valid as soon as the potential is set to zero at infinity (instead of the convention $U_e = 0$ adopted up to now). Therefore, from (2.43b) and (5.2f), we have:
\begin{eqnarray}
\frac{\ensuremath{\mathrm{d}} H_2}{\ensuremath{\mathrm{d}} T} &=& -\gamma \frac{\ensuremath{\mathrm{d}} L_e}{\ensuremath{\mathrm{d}}\fb{T}} \\
&=& - \gamma\ L_e \left(\frac{7}{4}b-\frac{1}{2}c\right) = \beta^{7/4}\ \gamma^{1/2}\ \frac{4-2\lambda}{3\lambda+1}\ \fb{L}_e\ c.\nonumber
\end{eqnarray}
We can finally compute the total variation of energy $H_1+H_2$ of the cluster. This variation is only due to the fact that the stars escape taking some energy away with them. The difference of gravitational potential between the boundary of the cluster and infinity is $U_\infty$ given by (5.7). Therefore, each star that escapes takes away a mass $m_1$ and a negative energy, equal to $-m_1 U_\infty$. The escape rate is, from (2.43a) and (5.2f):
\begin{eqnarray}
\frac{\ensuremath{\mathrm{d}} M_e}{\ensuremath{\mathrm{d}} T} = \gamma \frac{\ensuremath{\mathrm{d}} M_e}{\ensuremath{\mathrm{d}}\fb{T}} &=& \gamma\ M_e \left(\frac{3}{4}b-\frac{1}{2}c\right)\\
&=& - \beta^{3/4}\ \gamma^{1/2}\ \frac{2}{3\lambda+1}\ \fb{M}_e\ c,\nonumber
\end{eqnarray}
and thus,
\begin{equation}
\frac{\ensuremath{\mathrm{d}}(H_1 + H_2)}{\ensuremath{\mathrm{d}} T} = \beta^{7/4}\ \gamma^{1/2}\ \frac{2}{3\lambda+1}\frac{\fb{M}_e^2}{\fb{R}_e}\ c.
\end{equation}
The comparison of (5.16), (5.18), (5.20) shows that we must have:
\begin{equation}
\frac{3}{2}K_D\ K + \frac{4-2\lambda}{3\lambda+1}\ \fb{L}_e\ c = \frac{2}{3\lambda+1}\ \frac{\fb{M}_e^2}{\fb{R}_e}\ c.
\end{equation}
which expresses the conservation of energy. This relation between the parameters of the model provides a useful verification of the calculations. When using the numerical values (5.4) and (5.5), and when omitting the factor $\beta^{7/4}\ \gamma^{1/2}$, we get:
\begin{eqnarray}
\frac{\ensuremath{\mathrm{d}} H_1}{\ensuremath{\mathrm{d}} T} &=& -0.6464,\\
\frac{\ensuremath{\mathrm{d}} H_2}{\ensuremath{\mathrm{d}} T} &=& +0.9672,\nonumber\\
\frac{\ensuremath{\mathrm{d}}(H_1+H_2)}{\ensuremath{\mathrm{d}} T} &=& +0.3455.\nonumber
\end{eqnarray}
The conservation of energy is quite well verified; the discrepancy that remains comes from the errors in the computation (mainly the error made in $\fb{R}_e$).
Thus, in the course of the evolution, the diffuse energy $H_2$, always negative according to (5.17), decreases in absolute value; the numerical values (5.22) show that about one third of the negative energy is taken away by the stars that escape, while the other two thirds go in the center.
It remains to explain the mechanism of this accumulation of energy in the center. It is not associated with an accumulation of matter; we have supposed that the flux of matter toward the center of the cluster is zero (and this is indeed necessary, because a central condensation of mass would create an additional potential that would modify the structure of the cluster; for example the potential would vary as $1/r$ and not $\ln{(r)}$ anymore, near the center). Thus, the negative energy $H_1$ must be stored without any increase of the number of stars. Apparently, there is only one process that allows this: \emph{the formation of tight binary or multiple stars in the center of the cluster}.
The direct observation cannot confirm the existence of this phenomenon: the images of the stars are sorely separated \tn{resolved} in the central region of the globular clusters, and it would be impossible to discover there the presence of a particularly compact group of stars. On the other hand, \citet{vonHoerner1960} has computed numerically the evolution of artificial clusters; by this means, it is possible to observe the mechanism of the evolution, in as much detail as desired. von Hoerner has indeed noted the frequent formation of binaries in the center of the cluster; in one case, a compact group of 4 stars appeared. This seems an excellent confirmation of the process that we have been led to admit.
Note again that this central accumulation of energy does not affect the structure of the cluster, except for the few stars that support it; thus, it can develop independently of the global evolution of the cluster. In fact, we have seen that the absolute value of the central energy increases, while that of the diffuse energy decreases. At the end of the evolution, about two thirds of the initial negative energy is in the central condensation, and one third only has left the cluster, carried by the stars. It would be very interesting, although not possible here, to study in more detail the process of accumulation and to answer in particular the following questions: how do multiple stars form? How many are there? What happens to them after the cluster has disappeared?
\section{Stars with different masses}
The case of a cluster with an arbitraty mass distribution seems much more involved than the simple case of equal masses that we have considered so far; thus we shall not seek to treat it in the general case. But we are going to see that some simplified cases allow easy calculations, and provide, not a complete and rigorous solution, but at least several indications about the effect of the dispersion of the masses. We will first assume that we add a small number of stars of different masses to a cluster that contains equal mass stars, and will study the behavior of this secondary population. Later, this study will help us to obtain an approximate solution to the general case.
\subsection{Simplified model made of two populations}
We suppose that the cluster is made of the mix of a main population 1 of stars of mass $m_1$, and of a secondary population 2, numerically negligible with respect to the first one, of stars of mass $m_2$. The distribution function is not (2.23) anymore but rather:
\begin{equation}
f(E,m,t) = \delta(m-m_1) F_1(E,t) + \delta(m-m_2) F_2(E,t),
\end{equation}
with:
\begin{equation}
F_2 \ll F_1.
\end{equation}
The population 1 is almost not affected by the presence of the population 2; it alone determines the structure and the evolution of the cluster; as a consequence, all the results obtained in the previous Chapters still work for it. We assume that the population 1 has reached the final state represented by the homologous model of the Chapter~V.
Furthermore, by setting $m=m_2$ in (2.22), one gets a new equation, that describes the evolution of $F_2$. One can neglect the perturbations between the stars of the population 2, and the equation becomes, after applying the transformation of (2.24):
\begin{eqnarray}
0 &=& \frac{\partial}{\partial E} \bigg[ \mu F_2 \int_{-\infty}^E F_1 Q_1'\, \ensuremath{\mathrm{d}} E_1 \\
&& + F_2' \left(\int_{-\infty}^E F_1 Q_1\, \ensuremath{\mathrm{d}} E_1 + Q \int_E^{\infty} F_1\, \ensuremath{\mathrm{d}} E_1 \right) \bigg] \nonumber\\
&& +F_2' \frac{\partial Q}{\partial T} - Q' \frac{\partial F_2}{\partial T},\nonumber
\end{eqnarray}
where we set:
\begin{equation}
\frac{m_2}{m_1} = \mu.
\end{equation}
This equation has the dimension of $F_2$; therefore, we can apply an homologous transformation to $F_2$, independently of those of $F_1$. Thus, we set
\begin{equation}
F_2 = \gamma_2\ \fb{F}_2.
\end{equation}
Therefore the homology now depends on the three parameters $\beta$, $\gamma$, $\gamma_2$. We set
\begin{equation}
\frac{1}{\gamma_2}\frac{\ensuremath{\mathrm{d}}\gamma_2}{\ensuremath{\mathrm{d}}\fb{T}} = c_2,
\end{equation}
and when continuing the calculation, as in the Chapter~II, we obtain the equation
\begin{eqnarray}
0 &=& \Bigg\{ \mu \fb{F}_2 \int_{-\infty}^{\fb{E}} \fb{F}_1 \fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 \\
&& + \fb{F}_2' \bigg( \int_{-\infty}^{\fb{E}} \fb{F}_1 \fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1 + \fb{Q} \int_\fb{E}^{\infty} F_1\, \ensuremath{\mathrm{d}} E_1 \bigg) \nonumber\\
&& + \left( \frac{3}{4}b-\frac{3}{2}c \right) \fb{F}_2\fb{Q}\nonumber\\
&& + \left( \frac{3}{2}c-\frac{3}{4}b - c_2 \right) \int_{-\infty}^{\fb{E}} \fb{F}_2 \fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1\Bigg\}.\nonumber
\end{eqnarray}
In this equation, $\fb{F}_1$, $\fb{Q}$ or $\fb{Q}_1$, $b$, $c$ are the functions and constants of the homologous model, given in the Chapter~V; the function $\fb{F}_2$ and the constant $c_2$ are unknown.
The boundary conditions are only:
\begin{equation}
\fb{F}_2(0) = 0,
\end{equation}
which tells us that the stars of the population 2 escape when they reach the boundary of the cluster.
Near the center, on the other hand, we can neglect the last terms of (6.7) in a first order approximation, and use the asymptotical expressions (4.6) and (4.7c) for $\fb{F}_1$ and $\fb{Q}$; the equation becomes:
\begin{equation}
\mu\ \fb{F}_2 + \fb{F}_2' = 0.
\end{equation}
and thus
\begin{equation}
\fb{F}_2 = C\ \e{-\mu \fb{E}},
\end{equation}
where $C$ is an arbitrary constant. The equation (6.10) is as expected: indeed, it shows that the Maxwellian equilibrium is realised in the central region of the cluster between the two populations.
\subsubsection{The $\mu \ge 3/2$ case}
This functional form for $\fb{F}_2$ has a strange consequence. The mass of the subset of stars of the population 2 whose energy is below $E$ is, from (2.40a):
\begin{equation}
\mc{M}_2 = 16 \pi^2\ m_2 \int_{-\infty}^E F_2 q_1'\, \ensuremath{\mathrm{d}} E_1,
\end{equation}
i.e., when switching to normalized variables, and then to the canonical variables,
\begin{equation}
\fb{M}_2 = \mu \int_{-\infty}^{\fb{E}} \fb{F}_2 \fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1.
\end{equation}
$\fb{Q}'$ is given near the center by (4.7c), i.e.,
\begin{equation}
\fb{Q}' = \frac{3}{2} K_D\ \e{3\fb{E}/2}.
\end{equation}
When using this form and (6.10), we immediately see that the integral (6.12) diverges if
\begin{equation}
\mu \ge \frac{3}{2}.
\end{equation}
In other words, in this case, the mass of the population 2 is infinite, because of a too rapid increase of the distribution function (and of the density) toward the center.
This anomaly is easy to explain. We have assumed in (6.2) that $F_2$ is always negligible with respect to $F_1$; however, when comparing the asymptotical forms (4.6) and (6.10) of these two functions, we can see that when $\mu$ is greater than unity, $F_2$ increases faster than $F_1$ for $E\to -\infty$, and that, a value of $E$ below which $F_2$ becomes larger than $F_1$ always exists, however small the constant $C$. This critical value is (from Equations 2.31b and 6.5)
\begin{equation}
\fb{E}_c = \frac{\ln{(C\ \gamma_2/\gamma)}}{\mu-1}.
\end{equation}
Using the formula (6.13) for $\fb{Q}'$ is only meaningful when $\fb{E} \gg \fb{E}_c$. On the contrary, for $\fb{E} \ll \fb{E}_c$, $\fb{F}_1$ becomes negligible with respect to $\fb{F}_2$. Then, we can obtain the new forms of $\fb{D}$, $\fb{R}$, $\fb{Q}$ by using (6.10) and the fundamental equations. In particular, we get
\begin{equation}
\fb{Q}' = \frac{C\ \gamma_2}{\gamma} \mu^{5/4}\ \frac{3}{2}K_D\ \e{3\mu\ \fb{E}/2} \qquad \textrm{for } \fb{E} \ll \fb{E}_c
\end{equation}
\tn{The $c$ subscript is missing in the original version.} and with this correct form, the integral (6.12) does not diverge anymore. The quantity $\fb{F}_2\fb{Q}'$ yields a maximum in the vicinity of $\fb{E} = \fb{E}_c$ and exponentially decreases in both sides. We can compute the mass $\fb{M}_2$ in an approximate way by assuming that $\fb{Q}'$ is given by (6.16) when $\fb{E}<\fb{E}_c$ and by (6.13) when $\fb{E}>\fb{E}_c$; we find
\begin{eqnarray}
\fb{M}_2 &=& 3 K_D \bigg( \mu^{-5/4}\\
&& + \frac{\mu}{2\mu - 3}\bigg) \frac{\gamma}{\gamma_2} \left(\frac{C\ \gamma_2}{\gamma}\right)^{1/(2\mu-2)} \qquad \textrm{for } \fb{E} \gg \fb{E}_c,\nonumber
\end{eqnarray}
$\fb{M}_2$ does not depend on $\fb{E}$ anymore; indeed, because of the form of $\fb{F}_2\fb{Q}'$, almost all the stars of the population 2 have an energy of the order of magnitude of $\fb{E}_c$, therefore are part of $\fb{M}_2$ for $\fb{E} \gg \fb{E}_c$.
We focus here on the extreme case where the population 2 is negligible with respect to the population 1; therefore $C$ must be very small compared to unity. For $C\to 0$, (6.15) shows that $\fb{E}_c \to -\infty$. This way, we obtain the following result: \emph{the stars whose mass is greater than $3/2$ times the mean mass are almost all gathered near the center of the cluster}. Obviously, this conclusion is related to the simplified mass distribution that we have adopted, and should not be extended without further study to the case of an arbitrary mass distribution.
We also notice that $\fb{M}_2$ is not proportional to $C$: the power of $C$ in (6.17) is less than unity. As a consequence, for $C\to 0$, the last term of the equation (6.7) decreases slower than the others, which are proportional to $C$. In the limit, the equation becomes
\begin{equation}
\frac{3}{2}c-\frac{3}{4}b - c_2 = 0.
\end{equation}
Let's divide all the terms of (6.7) by $C$; the last term is then, for $C \to 0$, an indeterminate form $0\times \infty$. Let $p$ be its value; $p$ is a new constant yet to be determined, which replaces $c_2$.
\subsubsection{Expansion near the center}
Substituting in the equation for the expansions (4.35c), (4.36a), (4.37), we get, after some calculations, the expansion of $\fb{F}_2$ near the center. For $\mu < 3/2$:
\begin{eqnarray}
\fb{F}_2 &=& C\ \e{-\mu\fb{E}} \bigg\{ 1+ \mu K\ \e{\fb{E}/2}\\
&& + \bigg[ \frac{\mu(\mu-1)}{2}K^2 + \frac{c_2 - \mu c + \mu (\mu-1) b}{3-2\mu} \bigg] \e{\fb{E}} \bigg\}.\nonumber
\end{eqnarray}
For $\mu > 3/2$:
\begin{eqnarray}
\fb{F}_2 &=& C\ \e{-\mu\fb{E}} \bigg\{ 1+ \mu K\ \e{\fb{E}/2}\\
&& + \bigg[ \frac{\mu(\mu-1)}{2}K^2 + \frac{1}{2}c - \frac{1}{2} \bigg(\mu+\frac{1}{2}\bigg)b \bigg] \e{\fb{E}} \bigg\}.\nonumber
\end{eqnarray}
Finally, for the special case $\mu = 3/2$:
\begin{eqnarray}
\fb{F}_2 &=& C\ \e{-3\fb{E}/2} \bigg\{ 1+ \frac{3}{2} K\ \e{\fb{E}/2}\\
&& + \bigg[ \frac{3}{8}K^2 + \frac{1}{2}c - b - \frac{p}{3K_D\ C} \bigg] \e{\fb{E}} \bigg\}.\nonumber
\end{eqnarray}
\subsubsection{Method of solution}
Equation (6.7) is integrated from the center toward the boundary. The initial conditions are given by one of the expansions (6.19) to (6.21). The value of the factor $C$ does not matter here; in practice, one sets $C=1$. Proceeding by trial and error, we find the value of the parameter $c_2$ or $p$ for which the condition (6.8) is fulfilled at the boundary.
\subsection{Results: structure}
The calculation has been done for several values of the relative mass $\mu$. The value of the parameter is given in \reft{2}, column 2 or 3. \reff{8} plots the ratio of the distribution function $\fb{F}_2$ to its asymptotical form $C\ \e{-\mu\fb{E}}$. This ratio tends toward unity when $\fb{E}$ goes to $-\infty$; toward the boundary, it becomes smaller and smaller, which expresses the discrepancy between the real distribution and a Maxwellian distribution\footnote{\citet{Spitzer1958b} have computed these functions in a simpler case: the structure of the cluster was supposed to remain constant and was represented by a constant potential inside the cluster, zero outside; furthermore, the distribution function $F_1$ of the main population was supposed to be Maxwellian. The curves obtained by these authors have some similarities with ours but are arranged in reverse order! This peculiarity is probably linked to the too rough approximation made to the potential.}.
\begin{table}
\caption{}
\label{tab:2}
\begin{tabular}{r@{}lccr@{}l}
\multicolumn{2}{c}{$\mu$} & $c_2$ & $p$ & \multicolumn{2}{c}{$\theta$} \\
\hline
\hline
0& & -0.6240& & 1&.4396\\
0&.2& -0.3998& & 1&.2154\\
0&.4& -0.1838& & 0&.9994\\
0&.6& +0.0233& & 0&.7923\\
0&.8& +0.2208& & 0&.5948\\
1& & +0.4078& & 0&.4078\\
1&.2& +0.5829& & 0&.2327\\
1&.4& +0.7430& & 0&.0726\\
1&.5& & 0.4796& 0 & \\
1&.6& & 0.4670& 0 & \\
1&.8& & 0.4307& 0 & \\
2& & & 0.3859& 0 & \\
2&.5& & 0.2685& 0 & \\
3& & & 0.1726& 0 & \\
4& & & 0.0685& 0 & \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\includegraphics[width=\columnwidth]{f08}
\caption{Stars with different masses: distribution functions.}
\label{fig:8}
\end{figure}
\reff{9} and \reft{3} give, for several values of $\mu$, the product $\fb{D}_P\fb{R}$ of the projected density and the radius, as a function of the radius; we will see in the Chapter~VIII that this function is the one that best allows for a comparison with the observations. The factors have been adjusted so that all the curves cross at the same point: $\fb{D}_P\fb{R} = 1$ for $\fb{R} = 1$. The values of $C$ are given in the last row of \reft{3}.
\begin{figure}
\includegraphics[width=\columnwidth]{f09}
\caption{Stars with different masses: projected density times radius, as a function of the radius.}
\label{fig:9}
\end{figure}
\begin{table2}
\caption{$\log{(\fb{D}_P \fb{R})}$}
\label{tab:3}
\begin{tabular}{r@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}lr@{}l}
\multicolumn{2}{c}{$\fb{R}$} & \multicolumn{2}{c}{$\mu = 0$} & \multicolumn{2}{c}{0.2} & \multicolumn{2}{c}{0.4} & \multicolumn{2}{c}{0.6} & \multicolumn{2}{c}{0.8} & \multicolumn{2}{c}{1} & \multicolumn{2}{c}{1.2} & \multicolumn{2}{c}{1.4} & \multicolumn{2}{c}{1.6} & \multicolumn{2}{c}{1.8} & \multicolumn{2}{c}{2} & \multicolumn{2}{c}{2.5} & \multicolumn{2}{c}{3} & \multicolumn{2}{c}{4} \\
\hline
\hline
0& & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & 0.&655 & \multicolumn{2}{c}{$+\infty$} & \multicolumn{2}{c}{$+\infty$} & \multicolumn{2}{c}{$+\infty$} & \multicolumn{2}{c}{$+\infty$} & \multicolumn{2}{c}{$+\infty$} & \multicolumn{2}{c}{$+\infty$} & \multicolumn{2}{c}{$+\infty$} & \multicolumn{2}{c}{$+\infty$}\\
0&.110 & 1.&645 & 1.&763 & 1.&910 & 0.&090 & 0.&301 & & 542 & 0.&809 & 1.&094 & 1.&376 & 1.&659 & 1.&963 & 2.&783 & 3.&652 & 5.&433 \\
0&.168 & & 784 & & 883 & 0.&002 & & 145 & & 310 & & 497 & & 703 & 0.&923 & & 139 & & 355 & & 589 & & 228 & 2.&915 & 4.&332 \\
0&.262 & & 906 & & 983 & & 073 & & 178 & & 297 & & 431 & & 578 & & 735 & 0.&887 & & 036 & & 200 & 1.&657 & & 158 & 3.&203\\
0&.331 & & 958 & 0.&022 & & 097 & & 182 & & 279 & & 386 & & 505 & & 631 & & 751 & 0.&869 & 0.&999 & & 364 & 1.&769 & 2.&623\\
0&.420 & 0.&000 & & 051 & & 109 & & 175 & & 249 & & 331 & & 421 & & 516 & & 606 & & 693 & & 789 & & 063 & & 372 & & 030\\
0&.539 & & 029 & & 065 & & 106 & & 152 & & 204 & & 260 & & 321 & & 386 & & 447 & & 505 & & 569 & 0.&753 & 0.&965 & 1.&422\\
0&.703 & & 037 & & 058 & & 081 & & 107 & & 135 & & 165 & & 199 & & 234 & & 267 & & 297 & & 331 & & 429 & & 543 & 0.&794\\
0&.937 & & 012 & & 016 & & 020 & & 025 & & 029 & & 034 & & 041 & & 047 & & 052 & & 057 & & 063 & & 080 & & 100 & & 143\\
1&.095 & 1.&979 & 1.&974 & 1.&968 & 1.&962 & 1.&954 & 1.&947 & 1.&940 & 1.&931 & 1.&924 & 1.&917 & 1.&910 & 1.&889 & 1.&864 & 1.&806\\
1&.294 & & 924 & & 909 & & 893 & & 876 & & 857 & & 836 & & 816 & & 793 & & 774 & & 755 & & 736 & & 679 & & 613 & & 457\\
1&.552 & & 835 & & 811 & & 784 & & 755 & & 724 & & 691 & & 657 & & 620 & & 589 & & 560 & & 529 & & 439 & & 334 & & 087\\
1&.900 & & 688 & & 652 & & 614 & & 574 & & 530 & & 484 & & 438 & & 388 & & 344 & & 305 & & 263 & & 144 & & 003 & 2.&676\\
2&.125 & & 568 & & 528 & & 487 & & 443 & & 393 & & 343 & & 293 & & 239 & & 188 & & 146 & & 101 & 2.&968 & 2.&812 & & 455\\
2&.400 & & 413 & & 369 & & 325 & & 278 & & 224 & & 169 & & 113 & & 054 & & 001 & 2.&958 & 2.&908 & & 764 & & 594 & & 208\\
2&.744 & & 177 & & 133 & & 086 & & 036 & 2.&982 & 2.&925 & 2.&867 & 2.&805 & 2.&751 & & 704 & & 652 & & 505 & & 333 & 3.&930\\
3&.193 & 2.&797 & 2.&753 & 2.&706 & 2.&656 & & 602 & & 545 & & 487 & & 425 & & 371 & & 324 & & 272 & & 125 & 3.&953 & & 550\\
3&.805 & & 060 & & 016 & 3.&969 & 3.&919 & 3.&865 & 3.&808 & 3.&750 & 3.&688 & 3.&634 & 3.&587 & 3.&535 & 3.&388 & & 216 & 4.&813\\
4&.703 & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$} & \multicolumn{2}{c}{$-\infty$}\\
\multicolumn{2}{c}{$\log{(C)}$} & 1.&737 & 1.&769 & 1.&797 & 1.&821 & 1.&841 & 1.&857 & 1.&870 & 1.&877 & 1.&858 & 1.&827 & 1.&805 & 1.&769 & 1.&741 & 1.&651\\
\hline
\end{tabular}
\end{table2}
For $\mu=1$, we naturally retrieve the distribution of the homologous model; near the center, the quantity $\fb{D}_p\fb{R}$ tends toward a constant (see Equation 5.9). For $\mu <1$, it tends toward zero; for $\mu > 1$, it tends to infinity. Toward the boundary, the curves tend to become parallel.
\subsection{Escape rate}
From the formula (6.11) we find out that the total mass $\mc{M}_{2e}$ of the population 2 is proportional to
\[
\beta^{3/4}\ \gamma^{-3/2}\ \gamma_2.
\]
As a consequence, the relative escape rate of the stars of the population 2 is given by:
\begin{equation}
\theta= -\frac{1}{\mc{M}_{2e}}\frac{\ensuremath{\mathrm{d}}\mc{M}_{2e}}{\ensuremath{\mathrm{d}}\fb{T}} = -\frac{3}{4}b + \frac{3}{2}c - c_2.
\end{equation}
$\theta$ is given in \reft{2} column 4, and plotted in \reff{10} as a function of $\mu$. From (6.18), we have
\begin{equation}
\theta= 0 \qquad \textrm{for } \mu \ge \frac{3}{2}.
\end{equation}
Therefore, in our simplified model, \emph{the stars more massive than $3/2$ time the mean mass do not escape}.
\begin{figure}
\includegraphics[width=\columnwidth]{f10}
\caption{Escape rate as a function of mass. \tn{The legend reads \emph{Pr\'esent mod\`ele}: this model.}}
\label{fig:10}
\end{figure}
Between $\mu=0$ and $\mu=3/2$, the escape rate decreases regularly, as one would expect. The very low-mass stars ($\mu \simeq 0$) escape about 3.5 times faster than the mean-mass stars.
The escape rates computed by \citet{Chandrasekhar1942} and by \citet{Spitzer1958b}, normalized so that they all take the same value for $\mu=1$, are also plotted in \reff{10}. We can see that the three curves are very different, especially for the low masses. This shows strikingly the lack of rigor of the cluster theory, in its present state.
It is quite obvious, however, that the less massive stars must escape faster. The fact that the curve from Chandrasekhar yields a maximum and then decreases for the low-mass stars is due to excessive simplifications. Indeed, in his computation he assumes that the stars must first adopt a Maxwellian distribution (and thus, high velocities for the low-mass stars), before being able to escape. In reality, as we have seen, the stars escape almost as soon as they have reach their escape velocity.
Previous works \citep{vandenBergh1957, Takase1960} used the escape rate provided by Chandrasekhar to calculate the initial mass function of open clusters from their present-day mass function; it would be advisable to do these calculations again with a more exact escape rate. In particular, the abnormal lack of low-mass stars, found by Takase in the initial distribution of the Pleiades, is likely to disappear.
\subsection{Approximation in the general case}
We are going to show that the results obtained above in the case of the very special mass distribution (6.1) allow us to solve in an approximate way the general case of any mass distribution.
In general, the distribution function takes the form (2.4): $f(E,m,t)$. Let $n_m\ \ensuremath{\mathrm{d}} m$ be the number of stars of the cluster whose mass is between $m$ and $m+\ensuremath{\mathrm{d}} m$; we easily derive (see Equations 2.38 and 2.40)
\begin{equation}
n_m = 16 \pi^2 \int_{-\infty}^0 f q'\, \ensuremath{\mathrm{d}} E.
\end{equation}
$n_m$ depends on $m$ and also on $t$, because of the escape.
The total number of stars and the total mass are given by
\begin{eqnarray}
n &=& \int_0^\infty n_m\, \ensuremath{\mathrm{d}} m,\\
\mc{M}_e &=& \int_0^\infty n_m\ m\, \ensuremath{\mathrm{d}} m.\nonumber
\end{eqnarray}
The mean mass and the quadratic mean mass are defined by means of the classical formulae:
\begin{eqnarray}
\overline{m} &=& \frac{1}{n} \int_0^\infty n_m\, \ensuremath{\mathrm{d}} m = \frac{\mc{M}_e}{n},\\
\overline{m^2} &=& \frac{1}{n}\int_0^\infty n_m\ m^2\, \ensuremath{\mathrm{d}} m.\nonumber
\end{eqnarray}
This done, let's consider the fundamental equations (2.6) and (2.22). We can switch the order of integration with respect to $m$ and $E$. Then, we notice that $m$ can be removed if we introduce the two functions:
\begin{eqnarray}
\int_0^\infty f m\, \ensuremath{\mathrm{d}} m &=& J_1(E,t),\\
\int_0^\infty f m^2\, \ensuremath{\mathrm{d}} m &=& J_2(E,t).\nonumber
\end{eqnarray}
$J_1$ and $J_2$ are ``mean distribution functions'', obtained by weighting in two ways the distribution functions corresponding to the different values of the mass. To come back to equations already solved, we have to make an approximation: we suppose that theses two mean distribution functions are similar up to a factor. In this case, the equations (6.24) and (6.26) show that $J_1$ and $J_2$ must be proportional to $\overline{m}$ and $\overline{m^2}$ respectively, and thus we set
\begin{eqnarray}
J_1(E,t) &=& \overline{m}\ F(E,t),\\
J_2(E,t) &=& \overline{m^2}\ F(E,t).\nonumber
\end{eqnarray}
When using these expressions in the fundamental equations, they become
\begin{eqnarray}
\rho &=& 4\pi \overline{m} \int_U^{\infty} (2E-2U)^{1/2} F\, \ensuremath{\mathrm{d}} E\\
0 &=& \Bigg\{ 16 \pi^2\ G^2\ \ln{(n)} \frac{\partial}{\partial E} \bigg[ \\
&& \overline{m} mf \int_{-\infty}^{E} F_1 q_1'\, \ensuremath{\mathrm{d}} E_1\nonumber\\
&& + \overline{m^2} f' \bigg( \int_{-\infty}^{E} F_1 q_1\, \ensuremath{\mathrm{d}} E_1 \nonumber \\
&& + q \int_E^{\infty} F_1\, \ensuremath{\mathrm{d}} E_1 \bigg) \bigg] + f' \frac{\partial q}{\partial t}-q'\frac{\partial f}{\partial t}\Bigg\}.\nonumber
\end{eqnarray}
The other two fundamental equations, (2.7) and (2.19), do not involve the mass, and thus are not modified. We now define the normalized variables by means of transformations, slightly different from (2.24):
\begin{eqnarray}
\rho &=& 4\pi\ \overline{m}\ D\\
r &=& (16 \pi^2\ G\ \overline{m})^{-1/2}\ R\nonumber\\
q &=& (16 \pi^2\ G\ \overline{m})^{-3/2}\ Q\nonumber\\
\ensuremath{\mathrm{d}} t &=& [16 \pi^2 G^2\ \overline{m^2}\ \ln{(n)}]^{-1}\ \ensuremath{\mathrm{d}} T.\nonumber
\end{eqnarray}
We easily check that, this way, we retrieve the equations (2.25) which have been obtained for the case of equal mass stars. (The equation 6.30 must be multiplied by $m\, \ensuremath{\mathrm{d}} m$ and integrated.) As a consequence, the structure of the cluster, given by the functions $D$, $R$, $Q$, is that of the homologous model, and the mean distribution function $F$, introduced in (6.28), also matches that of the homologous model.
Furthermore, when setting
\begin{equation}
\frac{\overline{m}}{\overline{m^2}}\ m = \mu
\end{equation}
and replacing $f$ with $F_2$, we find that the equation (6.30) transforms into (6.3), i.e. the equation obtained at the beginning of this Chapter for the simplified model. It follows that the detailed distribution function $f$ is represented, for the different masses, by the solutions $F_2$ of the simplified model taking (6.32) into account.
From the previous relation, we note that $\mu$ is the relative mass computed by considering the mass unit to be, not the mean mass $\overline{m}$, but rather
\begin{equation}
m_0 = \frac{\overline{m^2}}{\overline{m}},
\end{equation}
which is different, in practice, by a factor greater than 2 (see Equation 8.9, below).
Finally, we note that the transformation equations (2.41) and (2.47) must be replaced with:
\begin{eqnarray}
\mc{M} &=& (16 \pi^2\ \overline{m})^{-1/2}\ G^{-3/2}\ M\\
\mc{L} &=& (16 \pi^2\ \overline{m})^{-1/2}\ G^{-3/2}\ L\nonumber\\
\rho_P &=& \overline{m}^{1/2}\ G^{-1/2}\ D_P.\nonumber
\end{eqnarray}
\section{Approach of the homologous model}
In this Chapter, we are back to the hypothesis of equal masses, and we are going to try to extend the results of the homologous model in another direction: we will study the evolution of a cluster whose shape is close to those of the homologous model, but not identical. We will first consider the case of a cluster that differs from the homologous model because its central density is finite; then the case of a cluster with infinite central density but with differences in the global structure. In both cases, the calculation will only be approximate. Finally, the results will be combined to draw a general picture of the evolution of the cluster.
\subsection{Formation of the central singularity}
Let's consider a cluster that matches the homologous model everywhere, except in a small central region, where it differs so that its central density is finite. The central potential is then also finite; let's call it $\fb{U}_0$. Near the center, the radius $\fb{R}$ is proportional to $(\fb{U} - \fb{U}_0)^{1/2}$; From (5.2d), we easily derive that $\fb{Q}$ is proportional to $(\fb{E} - \fb{U}_0)^3$. If we assume that the distribution function $\fb{F}$ also remains finite in the center, we find that $\fb{F}\fb{Q}'$ is proportional to $(\fb{E} - \fb{U}_0)^2$ near the center. It will be useful, for the next calculations, to have a formula for $\fb{F}\fb{Q}'$ as simple as possible; therefore, we define a simplified model by means of the following conditions: below a given value $\fb{E}=\fb{E}_1$, $\fb{F}\fb{Q}'$ is proportional to $(\fb{E} - \fb{U}_0)^2$; above this value, it is as the homologous model. In addition, the function and its derivative must be continuous for $\fb{E}=\fb{E}_1$ (\reff{11}). These conditions translate into:
\begin{eqnarray}
\fb{E}_1 &=& \fb{U}_0 + 4\\
\fb{F}\fb{Q}' &=& \left\{\begin{array}{ll}
\displaystyle
\frac{3}{32}K_D\ \e{\fb{E}_1/2}\ (\fb{E}-\fb{U}_0)^2 & \textrm{for } \fb{U}_0<\fb{E}<\fb{E}_1\\
\\
\displaystyle
\frac{3}{2}K_D\ \e{\fb{E}/2} & \textrm{for } \fb{E}>\fb{E}_1
\end{array}\right.\nonumber
\end{eqnarray}
\begin{figure}
\includegraphics[width=\columnwidth]{f11}
\caption{Homologous model (dashed line) and simplified model, with finite central density (solid line).}
\label{fig:11}
\end{figure}
We now suppose that the formula (7.1) is valid not only initially, but always, although the central potential $\fb{U}_0$ is a function of the time. Thus the evolution of the cluster consists in two phenomena that overlap: the normal homologous evolution, and the variation of the structure of a small central region. We will see that the latter can quite easily be computed when considering the fluxes of mass and energy toward the center of the cluster.
The partial mass of the stars whose energy is less than $\fb{E}_1$ is, for the model considered here (see Equation 2.42)
\begin{equation}
\fb{M} = \int_{\fb{U}_0}^{\fb{E}_1} \fb{F}\fb{Q}'\, \ensuremath{\mathrm{d}}\fb{E} = 2 K_D\ \e{\fb{E}_1/2}.
\end{equation}
For the homologous model, this mass would be, from (4.10b)
\begin{equation}
\fb{M}_0 = 3K_D\ \e{\fb{E}_1/2}.
\end{equation}
Furthermore, the mass of the stars whose energy is greater than $\fb{E}_1$ is the same for both models. The difference of mass between the present model and the homologous one is therefore
\begin{equation}
\Delta\fb{M} = \fb{M} - \fb{M}_0 = -K_D\ \e{\fb{E}_1/2}.
\end{equation}
In the same way, we compute the total energy of the stars whose \tn{individual} energy is less than $\fb{E}_1$:
\begin{eqnarray}
\fb{H} &=& \int_{\fb{U}_0}^{\fb{E}_1} \fb{E}\fb{F}\fb{Q}' \, \ensuremath{\mathrm{d}}\fb{E} = (6 + 2\fb{U}_0) K_D\ \e{\fb{E}_1/2}\\
\fb{H}_0 &=& (6 + 3\fb{U}_0) K_D\ \e{\fb{E}_1/2},\nonumber
\end{eqnarray}
thus,
\begin{equation}
\Delta\fb{H} = \fb{H} - \fb{H}_0 = -K_D\ \fb{U}_0\ \e{\fb{E}_1/2}.
\end{equation}
The definition (7.1) of the model is partly arbitrary. One can redo the calculations above for other definitions (for example, when supposing that $\fb{Q}'$ rather than $\fb{F}\fb{Q}'$ is represented by a parabola near the center); we note that the results are always
\begin{eqnarray}
\Delta\fb{M} &=& k_1\ \e{\fb{U}_0/2}\\
\Delta\fb{H} &=& (k_1 \fb{U}_0 + k_2) \e{\fb{U}_0/2},\nonumber
\end{eqnarray}
where $k_1$ and $k_2$ are two numerical constants, the values of which are slightly different from one model to the other. In the present case, we have
\begin{eqnarray}
k_1 &=& - K_D\ \e{2} = -3.387\\
k_2 &=& 0.\nonumber
\end{eqnarray}
By computing the derivative of (7.7) with respect to time, we obtain the fluxes of mass and energy toward the center:
\begin{eqnarray}
\frac{\partial \fb{M}}{\partial \fb{T}} &=& \frac{k_1}{2} \e{\fb{U}_0/2}\ \frac{\ensuremath{\mathrm{d}} \fb{U}_0}{\ensuremath{\mathrm{d}}\fb{T}}\\
\frac{\partial \fb{H}}{\partial \fb{T}} &=& \frac{1}{2} (k_1 \fb{U}_0 + 2k_1 + k_2) \e{\fb{U}_0/2}\ \frac{\ensuremath{\mathrm{d}}\fb{U}_0}{\ensuremath{\mathrm{d}}\fb{T}}.\nonumber
\end{eqnarray}
These fluxes are related, through the equations (4.31), to the factors $\alpha_1$ and $\alpha_2$ that appear in the expansion (4.14) of $\fb{F}$. It is more handy to write this expansion as
\begin{equation}
\fb{F} = \e{-\fb{E}} + (K+K_2 \fb{E})\ \e{-\fb{E}/2},
\end{equation}
by setting
\begin{eqnarray}
\frac{2}{3K_D}(5\alpha_2-\alpha_1) &=& K\\
-\frac{2}{3K_D} \alpha_2 &=& K_2.\nonumber
\end{eqnarray}
Then, we obtain the relations:
\begin{eqnarray}
K &=& \frac{k_1 \fb{U}_0 + 7 k_1 + k_2}{3K_D} \e{\fb{U}_0/2}\ \frac{\ensuremath{\mathrm{d}}\fb{U}_0}{\ensuremath{\mathrm{d}}\fb{T}}\\
K_2 &=& -\frac{k_1}{3K_D} \e{\fb{U}_0/2}\ \frac{\ensuremath{\mathrm{d}}\fb{U}_0}{\ensuremath{\mathrm{d}}\fb{T}}\nonumber
\end{eqnarray}
Furthermore, we have seen in Chapter~V that the boundary conditions require two relations for the parameters of the model. In the case of the homologous model, $K_2$ is zero, and the other two parameters $K$ and $c$ are determined uniquely. Here, there are three parameters: $K$, $K_2$ and $c$. Therefore, the boundary conditions require, after canceling of $c$, a relation between $K$ and $K_2$. As the models considered here are close to the homologous one, $K_2$ is slightly different from zero, and we can assume that the relation is linear, i.e.
\begin{equation}
K = K_0 + d\ K_2,
\end{equation}
where $K_0$ is the value taken by $K$ in the homologous model, and $d$ is a constant.
To find this constant, two models have been computed with a non-zero $K_2$. The procedure is as in Chapter~V, but the initial condition (5.3c) has to be replaced with (7.10); the other initial expansions have, of course, to be modified accordingly. While doing this, we assume that the first order expansions (4.7) are still valid; indeed these new models differ significantly from the homologous model only in the small central region, i.e. for very small values of $\fb{E}$ or $\fb{U}$. Near the boundary, the modification only affects the second order terms in the expansions, as visible for example in (7.10).
The results about the relation between $K$ and $K_2$ are:
\begin{center}
\begin{tabular}{r@{}lr@{}l}
\multicolumn{2}{c}{$K_2$} & \multicolumn{2}{c}{$K$} \\
\hline
\hline
0&& -0.&938464 \\
-0&.01& -0.&931757 \\
-0&.1& -0.&87028 \\
\hline
\end{tabular}
\end{center}
($K_2 = 0$ corresponds to the homologous model; the value obtained here is slightly different from that of Chapter~V, Equation 5.4a, because of a larger integration step). We verify that these figures match well a linear relation of the type (7.13) with
\begin{equation}
d = -0.6707.
\end{equation}
When combining the relations (7.12) and (7.13), we get
\begin{eqnarray}
\frac{\ensuremath{\mathrm{d}}\fb{U}_0}{\ensuremath{\mathrm{d}}\fb{T}} &=& \frac{3K_D\ K_0}{k_1}\ \frac{\e{-\fb{U}_0/2}}{\fb{U}_0 + 7 + d + k_2/k_1}\\
&=& \frac{0.3816\ \e{-\fb{U}_0/2}}{\fb{U}_0 + 6.329}.\nonumber
\end{eqnarray}
This differential equation allows to compute the variation of the central potential $\fb{U_0}$ as a function of time, and thus to solve our problem. Its explicit solution is:
\begin{equation}
\fb{T}-\fb{T}_2 = 5.241 (\fb{U}_0 + 4.329)\ \e{\fb{U}_0/2},
\end{equation}
where $\fb{T}_2$ is a constant. This relation is plotted in \reff{12}. We first note that the central potential $\fb{U}_0$ decreases with time, which corresponds to an increase of the central density, and to getting closer to the homologous model. Furthermore, this decrease gets faster and faster, so that the central potential becomes $-\infty$ after a finite timelapse: the cluster reaches the structure of the homologous model at a given time $\fb{T} = \fb{T}_2$, and keeps it afterwards.
\begin{figure}
\includegraphics[width=\columnwidth]{f12}
\caption{Evolution of the central potential.}
\label{fig:12}
\end{figure}
Substituting (7.15) into (7.9), we obtain the explicit expressions of the fluxes near the center:
\begin{eqnarray}
\frac{\partial \fb{M}}{\partial \fb{T}} &=& -\frac{0.6464}{\fb{U}_0 + 6.329}\\
\frac{\partial \fb{H}}{\partial \fb{T}} &=& -\frac{0.6464\ (\fb{U}_0 + 2)}{\fb{U}_0 + 6.329}\nonumber
\end{eqnarray}
The flux of mass is first positive and then decreases; it becomes zero at the time $\fb{T}_2$ and remains null afterwards. The flux of energy is always negative; its absolute value is first decreasing; at $\fb{T}_2$, it reaches the value of -0.6464, that its keeps afterwards (see 5.22a). These results confirm the reasoning made in Chapter~IV.
At the beginning of its life, the cluster is likely not very concentrated; let's assume that $\fb{U}_0$ has initially the largest value allowed by (7.15), i.e. $\fb{U}_0 = -6.329$. By assuming that $\fb{T}=0$ at the initial time, we find
\begin{equation}
\fb{T}_2= 0.4427.
\end{equation}
$\fb{T}_2$ is, in canonical variables, the time needed to form the central singularity.
\subsection{Proper differences}
We now suppose that the central density has already become infinite, but that the cluster as a whole has not completely reached the final state represented by the homologous model.
Let $\fb{F}_0$ be the distribution function of the homologous model; the distribution function of the cluster considered here would be
\begin{equation}
\fb{F} = \fb{F}_0 + \Delta\fb{F},
\end{equation}
where $\Delta\fb{F}$ is small with respect to $\fb{F}_0$. $\Delta\fb{F}$ is time-dependent. In the same way, we set
\begin{equation}
\fb{Q} = \fb{Q}_0 + \Delta\fb{Q}, \textrm{ etc.}
\end{equation}
The function $\fb{F}$ is assumed to be in the usual canonical form, defined by (4.6). The parameters $K$, $b$, $c$, would all have slightly different values than those of the homologous model:
\begin{equation}
K = K_0 + \Delta K, \textrm{ etc.}
\end{equation}
The cluster obeys the relations (5.2) and (5.3), as soon as the terms including time derivatives, which were zero for the homologous model (see Equations 2.37e, 2.53, 3.12 and 3.13), are put back in (5.2e) and (5.2f). These equations become
\begin{eqnarray}
0 &=& S' + \left(\frac{3}{4}b-\frac{3}{2}c\right) \fb{F}\fb{Q} + \left(\frac{1}{2}c-\frac{3}{4}b\right)\\
&&\int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 + \int_{-\infty}^{\fb{E}} \left( \fb{F}_1'\frac{\partial \fb{Q}_1}{\partial \fb{T}} - \fb{Q}_1'\frac{\partial \fb{F}_1}{\partial \fb{T}}\right)\, \ensuremath{\mathrm{d}}\fb{E}_1,\nonumber\\
0 &=& (3\lambda + 1) b + (2-2\lambda)c\nonumber\\
&&+4\lambda\frac{\ensuremath{\mathrm{d}} \ln{(\fb{M}_e)}}{\ensuremath{\mathrm{d}}\fb{T}}-4\frac{\ensuremath{\mathrm{d}} \ln{(\fb{R}_e)}}{\ensuremath{\mathrm{d}}\fb{T}}.\nonumber
\end{eqnarray}
By substituting the expressions (7.19) to (7.21) into the previous equations, and neglecting the second order terms, we obtain a set of linear equations that have the dimensions of $\Delta\fb{F}$, $\Delta\fb{Q}$, etc., and their derivatives with respect to time. We know that the general solution of such a system is (unless a degeneracy exists) a linear combination of particular solutions of the form
\begin{eqnarray}
\Delta\fb{F} &=& \Delta\fb{F}_0\ \e{-s \fb{T}},\\
\Delta\fb{Q} &=& \Delta\fb{Q}_0\ \e{-s \fb{T}}, \textrm{ etc},\nonumber
\end{eqnarray}
where $\Delta\fb{F}_0$, $\Delta\fb{Q}_0$, ... do not depend on the time, and $s$ is a constant, real or complex. As in the classical terminology, we shall call such a solution \emph{proper difference}, and $s$, the associated \emph{proper value} \tn{eigenvalue}. By substituting (7.23) into the equations, and dividing by $\e{-s\fb{T}}$, we obtain a system that is independent of time, but which involves a new parameter: $s$.
\subsubsection{Method of solution}
In practice, instead of writing and solving the system of equations of $\Delta\fb{F}$ and $\Delta\fb{Q}$, etc ..., it is easier to keep the equations of $\fb{F}$, $\fb{Q}$, etc ..., to compute a slightly different model from the homologous one by means of these equations, and to obtain the differences by simply computing the differences between the values of the two models. The differences should not be too large (so that the second order terms are indeed negligible), nor too small (so that one can get them with a sufficient accuracy); experiment led us to fix the amplitude of these differences by setting
\begin{equation}
\Delta K = -0.001.
\end{equation}
Taking into account that the device we use computes with 8 significant digits, we can obtain the differences with a precision of the order of 1/1000.
Hence, the system to solve is as (5.2) and (5.3), as soon as (5.2e) and (5.2f) are replaced with
\begin{eqnarray}
0 &=& \Bigg\{ \fb{F} \int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1\\
&& + \fb{F}'\left(\int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1 - \fb{Q}\fb{F}^{(-1)}\right) \nonumber\\
&& + \left(\frac{3}{4}b-\frac{3}{2}c\right)\fb{F}\fb{Q}\nonumber\\
&& +\left(\frac{1}{2}c-\frac{3}{4}b\right) \int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 \nonumber\\
&& + s\int_{-\infty}^{\fb{E}} [ \fb{Q}_1' (\fb{F}_1-\fb{F}_0) - \fb{F}_1' (\fb{Q}_1 - \fb{Q}_0)]\, \ensuremath{\mathrm{d}}\fb{E}_1 \Bigg\}\nonumber\\
0 &=& (3\lambda + 1)b + (2-2\lambda)c \\
&& + 4 s\left( \frac{\fb{R}_e-\fb{R}_{e_0}}{\fb{R}_{e_0}} - \lambda \frac{\fb{M}_e-\fb{M}_{e_0}}{\fb{M}_{e_0}}\right)\nonumber
\end{eqnarray}
and when (7.24) is added.
The method of solution is an extension of the one used for the homologous model (Chapter~V):
\begin{enumerate}
\item choose a value for $s$;
\item choose a temporary form for the function $\fb{F}$ that satisfies (5.3a) and (5.3c), with $K$ set by (7.21) and (7.24);
\item compute $\fb{D}$, $\fb{Z}$, $\fb{R}$, $\fb{Q}$;
\item adopt temporary value for $b$ and $c$, and integrate (7.25);
\item re-do after changing $b$ and $c$, until the final conditions (5.3a) and (5.3b) are fulfilled;
\item go back to point 3 with the new function $\fb{F}$;
\item when two consecutive approximations of $\fb{F}$ are equal: compute the right-hand side term of (7.26). It is not zero in general. Modify the value of $s$, go back to point 2; grope around this way with $s$ until (7.26) is true.
\end{enumerate}
The computation is quite long, because of the trial and error required on three parameters: $b$, $c$, $s$; that is why we limited ourselves to the exact computation of the proper differences corresponding to the two smallest proper values $s$. (We will see later that the possible values of $s$ are real, positive and form a discrete series.) These differences are the most interesting in practice, because they are those which decay the slowest. The other differences will be treated in a more approximate way later in this Chapter.
\subsubsection{Results: first proper difference}
We find that the smallest proper value is:
\begin{equation}
s = 1.81.
\end{equation}
The differences $\Delta\fb{F}$, $\Delta\fb{D}$, $\Delta\fb{R}$, $\Delta\fb{Q}$, $\Delta\fb{D}_P$ of the four fundamental functions and the projected density are given in \reft{4}. The differences of the parameters and the characteristic quantities are:
\begin{eqnarray}
\frac{\Delta b}{\Delta K} = +1.30 \qquad \frac{\Delta c}{\Delta K} = +5.22,\qquad\\
\frac{\Delta \fb{R}_e}{\Delta K} = -6.62 \quad \frac{\Delta \fb{M}_e}{\Delta K} = -0.537 \quad \frac{\Delta \fb{L}_e}{\Delta K} = +0.065.\nonumber
\end{eqnarray}
\begin{table2}
\caption{}
\label{tab:4}
\begin{tabular}{r@{}lr@{}lr@{}lr@{}lr@{}lr@{}l}
\multicolumn{2}{c}{$\fb{E}$ or $\fb{U}$} & \multicolumn{2}{c}{$\frac{\Delta \fb{F}}{\Delta K}$} & \multicolumn{2}{c}{$\frac{\Delta \fb{D}}{\Delta K}$} & \multicolumn{2}{c}{$-\frac{\Delta \fb{R}}{\Delta K}$} & \multicolumn{2}{c}{$-\frac{\Delta \fb{Q}}{\Delta K}$} & \multicolumn{2}{c}{$\frac{\Delta \fb{D_P}}{\Delta K}$} \vspace{0.1cm}\\
\hline
\hline
-5& & 12&.18 & 33&.4 & 0&.00673 & 0&.0000287 & 8&.84\\
-4&.9 & 11&.62 & 31&.2 & &\phantom{.00}748 & &\phantom{.0000}361 & &\\
-4&.8 & 11&.08 & 29&.4 & &\phantom{.00}831 & &\phantom{.0000}448 & 8&.42\\
-4&.7 & 10&.55 & 27&.6 & &\phantom{.00}923 & &\phantom{.0000}565 & &\\
-4&.6 & 10&.04 & 25&.8 & 0&.0102 & &\phantom{.0000}703 & 7&.98\\
-4&.5 & 9&.56 & 24&.2 & &\phantom{.0}114 & &\phantom{.0000}885 & &\\
-4&.4 & 9&.10 & 22&.6 & &\phantom{.0}126 & 0&.000110 & 7&.54\\
-4&.3 & 8&.66 & 21&.1 & &\phantom{.0}140 & &\phantom{.000}138 & &\\
-4&.2 & 8&.24 & 19&.7 & &\phantom{.0}155 & &\phantom{.000}171 & 7&.10\\
-4&.1 & 7&.84 & 18&.4 & &\phantom{.0}172 & &\phantom{.000}213 & &\\
-4& & 7&.45 & 17&.2 & &\phantom{.0}191 & &\phantom{.000}265 & 6&.66\\
-3&.9 & 7&.08 & 16&.0 & &\phantom{.0}212 & &\phantom{.000}329 & &\\
-3&.8 & 6&.73 & 14&.8 & &\phantom{.0}235 & &\phantom{.000}408 & 6&.21\\
-3&.7 & 6&.40 & 13&.8 & &\phantom{.0}261 & &\phantom{.000}508 & &\\
-3&.6 & 6&.07 & 12&.8 & &\phantom{.0}290 & &\phantom{.000}629 & 5&.77\\
-3&.5 & 5&.77 & 11&.9 & &\phantom{.0}323 & &\phantom{.000}782 & &\\
-3&.4 & 5&.47 & 11&.0 & &\phantom{.0}359 & &\phantom{.000}970 & 5&.33\\
-3&.3 & 5&.20 & 10&.1 & &\phantom{.0}399 & 0&.00121 & &\\
-3&.2 & 4&.93 & 9&.33 & &\phantom{.0}445 & &\phantom{.00}150 & 4&.89\\
-3&.1 & 4&.68 & 8&.59 & &\phantom{.0}495 & &\phantom{.00}186 & &\\
-3& & 4&.44 & 7&.89 & &\phantom{.0}552 & &\phantom{.00}232 & 4&.45\\
-2&.9 & 4&.21 & 7&.23 & &\phantom{.0}616 & &\phantom{.00}289 & &\\
-2&.8 & 3&.99 & 6&.61 & &\phantom{.0}688 & &\phantom{.00}360 & 4&.02\\
-2&.7 & 3&.78 & 6&.03 & &\phantom{.0}769 & &\phantom{.00}449 & &\\
-2&.6 & 3&.58 & 5&.49 & &\phantom{.0}861 & &\phantom{.00}562 & 3&.59\\
-2&.5 & 3&.39 & 4&.98 & &\phantom{.0}965 & &\phantom{.00}703 & &\\
-2&.4 & 3&.21 & 4&.50 & 0&.108 & &\phantom{.00}882 & 3&.18\\
-2&.3 & 3&.04 & 4&.06 & &122 & 0&.0111 & &\\
-2&.2 & 2&.87 & 3&.64 & &137 & &\phantom{.0}139 & 2&.77\\
-2&.1 & 2&.71 & 3&.25 & &154 & &\phantom{.0}176 & &\\
-2& & 2&.56 & 2&.90 & &174 & &\phantom{.0}223 & 2&.37\\
-1&.9 & 2&.42 & 2&.56 & &197 & &\phantom{.0}282 & &\\
-1&.8 & 2&.28 & 2&.25 & &223 & &\phantom{.0}359 & 1&.98\\
-1&.7 & 2&.15 & 1&.97 & &253 & &\phantom{.0}458 & &\\
-1&.6 & 2&.02 & 1&.71 & &288 & &\phantom{.0}587 & 1&.62\\
-1&.5 & 1&.90 & 1&.47 & &329 & &\phantom{.0}755 & &\\
-1&.4 & 1&.78 & 1&.25 & &377 & &\phantom{.0}975 & 1&.27\\
-1&.3 & 1&.66 & 1&.06 & &433 & 0&.127 & &\\
-1&.2 & 1&.55 & 0&.877 & &500 & &165 & 0&.95\\
-1&.1 & 1&.44 & 0&.718 & &580 & &217 & &\\
-1& & 1&.33 & 0&.576 & &675 & &287 & 0&.665\\
-0&.9 & 1&.22 & 0&.452 & &791 & &382 & &\\
-0&.8 & 1&.11 & 0&.345 & &933 & &514 & 0&.418\\
-0&.7 & 1&.00 & 0&.253 & 1&.11 & &700 & &\\
-0&.6 & 0&.886 & 0&.177 & 1&.34 & &965 & 0&.220\\
-0&.5 & 0&.769 & 0&.116 & 1&.63 & 1&.35 & &\\
-0&.4 & 0&.646 & 0&.0684 & 2&.01 & 1&.93 & 0&.0816\\
-0&.3 & 0&.512 & 0&.0346 & 2&.55 & 2&.83 & &\\
-0&.2 & 0&.364 & 0&.0130 & 3&.32 & 4&.28 & 0&.0111\\
-0&.1 & 0&.195 & 0&.0024 & 4&.53 & 6&.80 & &\\
0& & 0 && 0 && 6&.62 & 11&.61 & 0&\\
\hline
\end{tabular}
\end{table2}
\reff{13} compares the projected densities of the usual homologous model ($\Delta K=0$) with those of the homologous model modified by the first proper difference; here we have set $\Delta K=0.2$ so that the difference is well-visible. (As for its amplitude, the sign of $\Delta K$ is arbitrary; we could have chosen a negative $\Delta K$; in this case, the difference with the homologous model would have been in the opposite sense.) We see that the difference mostly affects the external regions of the cluster: \emph{the first proper difference mainly consists in a variation of the external radius of the cluster}. The structure of the internal region differs little; the total mass and kinetic energy vary much less than the radius, as shown by the values (7.28).
\begin{figure}
\includegraphics[width=\columnwidth]{f13}
\caption{Usual homologous model ($\Delta K=0$) and homologous model modified by the first proper difference ($\Delta K=0.2$).}
\label{fig:13}
\end{figure}
This is naturally explained by the fact that the perturbations between the stars are more efficient in the center of the cluster, where the density is higher; as a consequence this region becomes close to the final state, represented by the homologous model, earlier.
\subsubsection{Results: second proper difference}
The second proper value is:
\begin{equation}
s = 5.12.
\end{equation}
We only give the differences of the parameters:
\begin{eqnarray}
\frac{\Delta b}{\Delta K} = +1.71 \qquad \frac{\Delta c}{\Delta K} = +6.90,\qquad\\
\frac{\Delta \fb{R}_e}{\Delta K} = -3.48 \quad \frac{\Delta \fb{M}_e}{\Delta K} = -0.760 \quad \frac{\Delta \fb{L}_e}{\Delta K} = +0.182.\nonumber
\end{eqnarray}
The profile of the projected density (not shown here) shows that the second proper difference, like the first one, mostly consists in a variation of the external radius.
\subsection{Stability of the homologous model}
Up to now, we have implicitly assumed that the clusters naturally tend toward the homologous model. For this to be true, the model has to be stable, i.e. any difference with respect to this model is decreasing. On the other hand, the stability of the homologous model, if demonstrated, will be, if not a rigorous proof, at least a very strong clue that the model is indeed the final state toward which all clusters tend.
For the homologous model to be stable, it is necessary and sufficient that all the proper values $s$ yield a positive real part. Thus, we are going to study the complete family of the proper values $s$, thanks to a very approximate calculation, yet sufficient for the goal we seek.
We first assume that the central expansion (4.34) of $\fb{F}$ is valid up to $\fb{E}=-1$. (This approximation, as all which will follow, has been suggested and checked by the exact computation of the first two proper functions.) Hence, for $\fb{E}=-1$, we have:
\begin{eqnarray}
\frac{\Delta\fb{F}}{\Delta K}(-1) &=& \e{1/2},\\
\frac{\Delta\fb{F}'}{\Delta K}(-1) &=& -\frac{1}{2}\e{1/2}.\nonumber
\end{eqnarray}
Between $\fb{E}=-1$ and $\fb{E}=0$, we consider the exact equation (7.25). By calculating the derivative of this equation, neglecting the terms in $\fb{F}^{(-1)}$ and $\fb{F}^2$ (because $\fb{F}$ vanishes at the boundary), replacing (7.19) etc ..., and subtracting the equation of the homologous model, we obtain
\begin{eqnarray}
0 &=& \Bigg\{ \int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1 \cdot \Delta\fb{F}'' \\
&& + \fb{F}''\Delta\left(\int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1\right) + \int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 \cdot \Delta\fb{F}'\nonumber\\
&& + \fb{F}' \Delta\left(\int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1\right)\nonumber\\
&& + \fb{F}'\fb{Q}\Delta\left(\frac{3}{4}b-\frac{3}{2}c\right) + \left(\frac{3}{4}b-\frac{3}{2}c\right) \fb{Q}\Delta\fb{F}' \nonumber \\
&& + \left(\frac{3}{4}b-\frac{3}{2}c\right) \fb{F}'\Delta\fb{Q} - \fb{F}\fb{Q}'\Delta c - c\fb{Q}'\Delta\fb{F} \nonumber\\
&& - c\fb{F}\Delta\fb{Q}' + s\fb{Q}'\Delta\fb{F} - s\fb{F}'\Delta\fb{Q}\Bigg\}\nonumber
\end{eqnarray}
\tn{The upper limit of the first integral is missing in the original version}
Furthermore (7.26) becomes, when neglecting the term in $\Delta\fb{M}_e$:
\begin{equation}
0 = (3\lambda + 1)\Delta b + (2-2\lambda)\Delta c + 4s \frac{\Delta \fb{R}_e}{\fb{R}_e}.
\end{equation}
The expansions (4.35) show that we have, near the center,
\begin{eqnarray}
\frac{\Delta\fb{R}}{\fb{R}} &=& -\frac{1}{\sqrt{2}}\e{\fb{U}/2}\ \Delta K,\\
\frac{\Delta\fb{Q}}{\fb{Q}} &=& -\frac{9\sqrt{3}}{8\sqrt{2}}\e{\fb{E}/2}\ \Delta K,\nonumber\\
\frac{\Delta\fb{Q}'}{\fb{Q}'} &=& -\frac{3\sqrt{3}}{2\sqrt{2}}\e{\fb{E}/2}\ \Delta K.\nonumber
\end{eqnarray}
We shall suppose that these expressions are valid up to the boundary. In (7.32), we can neglect the second and the fourth terms, because the integrals do not change much. We also neglect the terms in $\Delta b$, as they are small with respect to $\Delta c$. By taking $\Delta c$ from (7.33) and substituting it in (7.32), we obtain
\begin{eqnarray}
0 &=& \Bigg\{ \int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1\, \ensuremath{\mathrm{d}}\fb{E}_1 \cdot \Delta\fb{F}'' \\
&& + \left( \int_{-\infty}^{\fb{E}}\fb{F}_1\fb{Q}_1'\, \ensuremath{\mathrm{d}}\fb{E}_1 -\frac{3\lambda+3}{3\lambda+1} c\ \fb{Q}\right) \Delta\fb{F}' \nonumber\\
&& + (s-c)\fb{Q} \Delta\fb{F} + (s-c) \bigg[\frac{9\sqrt{3}}{8\sqrt{2}}\ \e{\fb{E}/2}\ \fb{F}'\fb{Q} \nonumber\\
&& - \frac{\sqrt{2}}{1-\lambda}\left(\fb{F}\fb{Q}'+\frac{3}{2}\fb{F}'\fb{Q}\right)\bigg]\Delta K\nonumber\\
&& + c \bigg[\frac{6\lambda+4}{3\lambda+1}\frac{9\sqrt{3}}{8\sqrt{2}}\ \e{\fb{E}/2}\ \fb{F}'\fb{Q} + \frac{3\sqrt{3}}{2\sqrt{2}}\ \e{\fb{E}/2}\ \fb{F}\fb{Q}'\nonumber\\
&& - \frac{\sqrt{2}}{1-\lambda}\left(\fb{F}\fb{Q}'+\frac{3}{2}\fb{F}'\fb{Q}\right)\bigg]\Delta K\Bigg\}\nonumber
\end{eqnarray}
Finally, we neglect the term in $\Delta\fb{F}'$, whose factor is relatively small, and the term in $c$. The equation shrinks to:
\begin{equation}
0 = \Delta\fb{F}'' + (s-c) B_1\ \Delta F + (s-c) B_2\ \Delta K,
\end{equation}
where $B_1$ and $B_2$ are two functions of $\fb{E}$. This is a second order differential equation for $\Delta\fb{F}$; it must fulfill the boundary conditions (7.31), as well as the condition
\begin{equation}
\Delta\fb{F}(0) = 0.
\end{equation}
These three conditions would be simultaneously fulfilled only for certain values of $s$, which are the proper values we seek.
The numerical computation of the functions $B_1$ and $B_2$ shows that they can be quite well fitted, in the range (-1,0), with the expressions
\begin{eqnarray}
B_1 & \simeq & \frac{2.10}{(0.332-\fb{E})^2},\\
B_2 & \simeq & (0.265 + 0.643\ \fb{E})\ B_1.\nonumber
\end{eqnarray}
By replacing this in (7.36), we find that the general solution of this equation is
\begin{eqnarray}
\frac{\Delta\fb{F}}{\Delta K} &=& -0.265 - 0.643\ \fb{E}\\
&& + C_1 (0.332 - \fb{E})^{p_1} + C_2 (0.332 - \fb{E})^{p_2},\nonumber
\end{eqnarray}
where $C_1$ and $C_2$ are two arbitrary constants, and $p_1$, $p_2$ are the roots of the equation:
\begin{equation}
p^2 - p + 2.10 (s-c) =0,
\end{equation}
so that we have
\begin{eqnarray}
p_1+p_2 &=& 1,\\
p_1\ p_2 &=& 2.10 (s-c).\nonumber
\end{eqnarray}
By writing the boundary conditions (7.31) and (7.37) for the formula (7.39), and then eliminating $C_1$ and $C_2$, we obtain
\begin{equation}
\left|\begin{array}{rrr}
(1.332)^{p_1} & (1.332)^{p_2} & 1.271 \\
p_1 (1.332)^{p_1} & p_2 (1.332)^{p_2} & -0.181 \\
(0.332)^{p_1} & (0.332)^{p_2} & 0.265
\end{array}\right|=0.
\end{equation}
The equations (7.41a) and (7.42) make a system of equations for $p_1$ and $p_2$ that can be solved numerically. We do not give the details of this solution, which is long but without difficulty; we find that $p_1$ and $p_2$ are necessarily of the form
\begin{eqnarray}
p_1 = \frac{1}{2}+\xi i,\\
p_1 = \frac{1}{2}-\xi i,\nonumber
\end{eqnarray}
where $\xi$ is real, and $i = \sqrt{-1}$. The possible values of $\xi$ constitute an infinite series; the first ones are given in the table below, as well as the corresponding values of $s$, found from (7.41b).
\begin{center}
\begin{tabular}{r@{}lr@{}l}
\multicolumn{2}{c}{$\xi$} & \multicolumn{2}{c}{$s$} \\
\hline
\hline
1&.21 & 1&.22\\
3&.82 & 7&.47\\
5&.43 & 14&.6\\
8&.28 & 33&.2\\
9&.91 & 47&.3\\
12&.8 & 78&.3\\
14&.4 & 99&.6\\
17&.3 & 143&\\
18&.9 & 171&\\
21&.8 & 227&\\
\hline
\end{tabular}
\end{center}
That is, all the proper values $s$ are real and positive; therefore, \emph{the homologous model is stable}, and it is likely the final state toward which all the clusters tend.
The fact that $s$ never yields an imaginary part shows that the clusters tend toward the homologous model through a simple ``relaxation'', with no oscillations.
The first two values of $s$ are in rough agreement with the exact values (7.27) and (7.29); the differences are not surprising, given the great number of approximations we made.
Note that the values of $s$ increase very fast which shows indeed that only the first values are interesting, in practice; the next ones corresponds to differences that vanish very quickly.
Finally, we note that these results allow us to define rigorously the ``relaxation time'' for the entire cluster: this would be the time required for the amplitude of the first proper difference to be divided by $e$.
\subsection{General picture of the evolution}
We have identified three distinct evolutionary phenomena:
\begin{enumerate}
\item the normal homologous evolution;
\item the formation of the central singularity;
\item the decrease of the differences.
\end{enumerate}
We are now going to compare their timescales.
The normal homologous evolution can be illustrated through the variation of the total mass given by (5.15) and plotted as a straight line on \reff{14}. The cluster disappears after a time $T_1$ given in (5.13):
\begin{equation}
T_1 = \frac{1}{\gamma_0 c} = \frac{2.452}{\gamma_0}.
\end{equation}
The formation of the central singularity can be seen as the variation of $\Delta M$, difference in mass with respect to the homologous model. This variation is given in (7.7a) and (7.16), in homologous variables; we switch to non-homologous variables by means of (5.12) and (5.15). The resulting curve is plotted in \reff{14} (for all the curves of this figure, the mass or the mass difference is normalized with respect to its own value at $T=0$).
\begin{figure}
\includegraphics[width=\columnwidth]{f14}
\caption{Compared evolutions: EH = homologous evolution \tn{\emph{\'evolution homologique}}; EP1 = first proper difference; EP2 = second proper difference \tn{\emph{\'ecart propre}}; C = formation of the central singularity.}
\label{fig:14}
\end{figure}
We immediately see that the formation of the central singularity is relatively rapid: it ends at the time $T_2$, computed from (7.18) and (5.12):
\begin{equation}
T_2 = \frac{0.4051}{\gamma_0}
\end{equation}
and thus occurs for only the first sixth of the lifetime of the cluster.
Finally, the decrease of the proper differences will also be represented as variations of the mass differences. In homologous variables, we have, from (7.23),
\begin{equation}
\frac{\Delta \fb{M}_e}{\Delta \fb{M}_{e_0}} = \e{-s\fb{T}},
\end{equation}
and we switch back to non-homologous variables by means of (5.12) and (5.15), which gives
\begin{equation}
\frac{\Delta M_e}{\Delta M_{e_0}} = \left(1-\frac{T}{T_1}\right)^{1+s/c}.
\end{equation}
This variation is plotted in \reff{14}, for the first two proper differences. We see once more, that the decrease of the differences is rapid with respect to the homologous evolution.
In reality, the three evolutionary phenomena overlap. The initial amplitude of the differences with respect to the homologous model cannot be determined in the framework of this theory; it could only be found thanks to a detailed study of the formation and the initial phases of the evolution (see Chapter~IX). However, there is no reason for the initial state of the cluster to already be close to the homologous model, and thus the initial differences in the various quantities are likely of the same order of magnitude as the quantities themselves. Therefore, we can suppose that the curves of \reff{14}, normalized to unity for $T=0$, roughly represent the relative importance of the various effects.
The evolution of a cluster can be describe in a broad outline as follows: the central density first increases, faster and faster; simultaneously, the differences with respect to the homologous model decrease everywhere in the cluster. After a time $T_1/6$, the central density has became infinite; the remaining difference is practically limited to the first proper difference; the cluster has only lost 1/6 of its mass, through escapes. Then, the first proper difference continues to decrease; after a time $T_1/3$, it has almost vanished; the cluster takes the form of the homologous model and keeps it until it disappears, at the time $T_1$.
We will see in the next Chapter how these results can be used to estimate the age of globular clusters.
\section{Application to globular clusters}
We are going to compare the theoretical results from the previous Chapters to observational data of globular clusters. The first goal of this comparison is to check that the theory is compatible with the facts (but this would not be a proof of the correctness of the theory, for the reasons detailed in Chapter~I); the comparison will give us information that is out of reach of the direct observation, in particular about the mass-luminosity relation and the age of the clusters.
\subsection{Artificial cluster}
In order to get a concrete picture of the theoretical model, an ``artificial cluster'' has been built by calculating the spacial coordinates of the stars, from a table of random numbers, so that it reproduces the density law of the homologous model. More precisely, let $n_1$, $n_2$, $n_3$ be three random number, taken between 0 and 1, with an uniform probability density function; the spherical coordinates $\fb{R}$, $\theta$, $\varphi$ of the star are computed by means of
\begin{eqnarray}
\fb{M}_\fb{R} &=& n_1 \fb{M}_e,\\
\cos{\theta} &=& 2n_2 -1,\nonumber\\
\varphi &=& 2\pi n_3.\nonumber
\end{eqnarray}
\reff{15} plots the two-dimensional projection of the cluster obtained. The number of stars is 1320. The figure covers only a fraction of the surface of the cluster (see the scale at the bottom of the figure), but, however, contains all the stars; this is because the projected density is extremely small in the external regions.
\begin{figure2}
\includegraphics[width=16cm]{f15}
\caption{Homologous model: artificial cluster.}
\label{fig:15}
\end{figure2}
\reff{16} is a zoom-in on the central region of the cluster.
\begin{figure2}
\includegraphics[width=16cm]{f16}
\caption{Zoom-in on the center of \reff{15}.}
\label{fig:16}
\end{figure2}
Globally, this artificial cluster resembles well the real clusters; the similarity would not be perfect because the artificial cluster of \reff{15} corresponds to the case of equal masses, while the real clusters yields an extended mass spectrum, of which we only observed the upper end. The central condensation in \reff{16} is likely to be too high \citep{King1961a}.
\subsection{Comparisons of the projected densities}
A more precise comparison of the homologous model with real clusters can be obtained when plotting the density profiles. It seems better to do the comparison for the projected densities, and not the spacial densities. Indeed, the theoretical curves can be obtained with an arbitrary accuracy and going from the theoretical spacial density to the theoretical projected density is errorless; on the contrary, going from the observed projected density to the spacial density strongly amplifies the errors, that were already not negligible. This is shown in \reff{17} where the spacial density of the cluster M3, derived from the observations thanks to two different methods \citep{Kholopov1955, Oort1959} is plotted.
\begin{figure}
\includegraphics[width=\columnwidth]{f17}
\caption{Spacial density of the cluster M3, from Kholopov (top) and Oort \& van Herk (bottom). The curves are vertically shifted.}
\label{fig:17}
\end{figure}
Generally, it seems better to transform the observational data as little as possible, therefore to compare theory and observations not at an intermediate level, but rather at the very level of the observations.
We shall use, not the projected density itself, but its product with the distance to the center: $\rho_P r$. This has several advantages:
\begin{enumerate}
\item in the theoretical model (case of equal mass stars), this quantity tends to a finite value in the center, while the projected density becomes infinite.
\item $\rho_P r$ varies more slowly than the projected density (recall \reff{5} and \reff{9});
\item the observations are often counts of stars in concentric rings of constant width. The number of stars in a ring is
\begin{equation}
\Delta n = \frac{1}{m} \int_{r_1}^{r_2} 2\pi \rho_P\ r \, \ensuremath{\mathrm{d}} r.
\end{equation}
Because this function is smooth and does not yield a strong curvature, we can write
\begin{equation}
\Delta n = \frac{2\pi}{m} \Delta r\ (\rho_P r)_{\overline{r}},
\end{equation}
with:
\begin{equation}
\Delta r = r_2 - r_1, \qquad \overline{r} = \frac{r_1 + r_2}{2}.
\end{equation}
That is, the observed numbers $\Delta n$ immediately give, up to a factor, the values of $\rho_P r$.
\end{enumerate}
\subsubsection{Projected density: star counts (M3)}
\citet{Sandage1954, Sandage1957a} has made detailed counts of stars in the cluster M3 = NGC~5272; the results have been published by \citet{Oort1959}. These counts corresponds to the stars brighter than a given magnitude, and situated in concentric rings. It would be necessary, in principle, to derive the numbers of stars in successive magnitude ranges by subtraction; however, attempts showed that this strongly increases the spreading of the points, because the errors accumulate, while the number decreases. Furthermore, because the luminosity function rapidly grows with the magnitude \citep[see][]{Sandage1957a}, the stars brighter than a given magnitude are almost all gathered in a small range of magnitude, and thus all have similar masses (we assume a well-defined mass-luminosity relation). It is therefore better and allowed to accept that the observed numbers correspond, for each limit in magnitude, to stars of a given mass.
\begin{figure2}
\includegraphics[width=16cm]{f18}
\caption{M3: comparison between observations (points) and theory (lines). \tn{The scale reads \emph{Unit\'es canoniques}: canonical units.}}
\label{fig:18}
\end{figure2}
\begin{figure2}
\includegraphics[width=16cm]{f19}
\caption{Continuation of \reff{18}.}
\label{fig:19}
\end{figure2}
That is, the points on \reff{18} mark the observed numbers of stars brighter than a given magnitude $M_V$, in rings of width 30'' situated at the mean distance $r$ from the center. These numbers must be compared to the theoretical curves obtained in Chapter~VI (\reff{9}). By trial and error, we find that the best match is obtained for
\begin{equation}
1' \sim 0.21 \textrm{ canonical units}.
\end{equation}
We note that this gives out an external radius
\begin{equation}
r_e = 4.703 \textrm{ canonical units} \sim 22.4' \sim 79 \textrm{ pc}.
\end{equation}
For each group, the curve is fitted, by playing with two parameters: the relative mass $\mu$, and factor $C$, i.e. the number of stars in the group; a variation of $C$ corresponds simply to a vertical shift of the curve.
The values of $\mu$ and $C$ determined this way are given in \reft{5}, columns 4 and 5. The curves are plotted in \reff{18}; we see that the agreement is as good as possible, given the accuracy of the observations. Only the brightest stars are not well-retrieved near the center of the cluster, which can easily be explained by two reasons: (1) we have seen in Chapter~VI that the simplified theoretical model is likely to be wrong near the center, where it predicts a too high density of massive stars; (2) the observation underestimates the number of stars in the very populated central regions.
\begin{table}
\caption{}
\label{tab:5}
\begin{tabular}{ccccccc}
$M_V$ & $M_V$ & $\overline{M_V}$ & $\mu$ & $\log{(C)}$ & $\log{\left(\frac{m}{\U{M}_{\odot}}\right)}$ & $M_\textrm{bol}$ \\
min & max & &&&&\\
\hline
\hline
$-\infty$ & 0.4 & -0.5 & 1.55 & 1.328 & 0.090 & -0.6\\
$-\infty$ & 1.2 & -0.1 & 1.50 & 1.538 & 0.076 & -0.2\\
$-\infty$ & 1.9 & 0.5 & 1.45 & 1.698 & 0.061 & 0.4\\
$-\infty$ & 3.5 & 2.2 & 1.42 & 2.472 & 0.052 & 2.1\\
$-\infty$ & 4.1 & 2.8 & 1.40 & 2.672 & 0.046 & 2.7\\
$-\infty$ & 4.6 & 3.3 & 1.10 & 2.894 & 1.941 & 3.2\\
$-\infty$ & 5.4 & 3.9 & 1.05 & 3.125 & 1.921 & 3.8\\
$-\infty$ & 6.3 & 4.5 & 0.80 & 3.246 & 1.803 & 4.4\\
4.6 & 6.3 & 5.5 & 0.60 & 2.986 & 1.678 & 5.4\\
\hline
\end{tabular}
\end{table}
\subsubsection{Mass-luminosity relation of the stars}
\reft{5}, column 3, gives the mean absolute magnitude $\overline{M_V}$ of each group, computed from the luminosity function \citep{Sandage1957a}. We notice that the difference between this mean magnitude and the maximum magnitude is never big, which confirms what has been said above.
The values of $\mu$ and $\overline{M_V}$ allow us to plot the mass-luminosity relation of the cluster (\reff{20}); $\mu$ is proportional to the mass $m$ of the stars, according to (6.32). The bolometric correction has been uniformly taken equal to -0.1. In order to extend the relation as much as possible for the faintest stars, we made an exception and considered the group of stars whose magnitude is between 4.6 and 6.3 (in the case of the narrower group $5.4<M_V<6.3$, the dispersion of the points is too large and we cannot plot a curve). This group is compared to the closest theoretical curve in \reff{19}; the parameters are given in the last row of \reft{5}.
\begin{figure}
\includegraphics[width=\columnwidth]{f20}
\caption{Mass-luminosity relation in M3, from the spacial distribution of the stars (dots) and from the theory of stellar evolution (line).}
\label{fig:20}
\end{figure}
The points in \reff{20} show a well-defined relation\footnote{\citet[Figure~8]{vonHoerner1957} has obtained a similar curve, from the same observations, but with a different theoretical model. Our results indicate a larger variation of the mass as a function of the luminosity.}. We observe a strong bend at $M_V = 3$, which corresponds exactly to the bend in the Hertzsprung-Russell diagram, and which marks the beginning of the zone of rapid evolution of the stars; the stars with a magnitude below 3 almost have all the same mass. The slight increase of the points on the left-hand side is likely not real and comes from errors when computing $\mu$.
It is interesting to compare this mass-luminosity relation, deduced from the purely mechanical considerations, to those provided by the stellar evolution theory. \citet[Table~6]{Sandage1957b} has computed the relation between the present-day magnitude and the initial magnitude of the stars of M3, thanks to a semi-empirical method based on the observed HR diagram and on the theoretical evolutionary tracks. We get the mass from the initial magnitude, assuming that the initial mass-luminosity relation is that of the main sequence, i.e.
\begin{equation}
\frac{L}{\textrm{L}_\odot} = \left(\frac{m}{\U{M}_{\odot}}\right)^4.
\end{equation}
This way, we get the present-day mass-luminosity relation, the solid line on \reff{20}; the two relations have been vertically shifted so that their horizontal parts match.
The two curves are clearly in disagreement. The position of the bend is more or less the same; but on the low luminosities end, the ``dynamical'' masses decrease much faster than the mass found from the evolution.
It is hard to tell which curve is wrong. Neither is very reliable. The finding of the dynamical masses relies on an approximate calculation (Chapter~VI), which may lead to large errors when (as is the case for globular clusters) the mass spectrum is very broad. On the other hand, the hypothesis stating that the initial mass-luminosity relation of the cluster would match the main sequence, should be taken carefully, because of different chemical abundances.
We will explore the possibilities of progress in the computation of the dynamical masses in the last Chapter.
The adjustment of the vertical scales gives an interesting piece of information: when compared to (6.32), we find that
\begin{equation}
\frac{\overline{m^2}}{\overline{m}} = 0.80 \U{M}_{\odot}.
\end{equation}
This value can also be computed from the mass spectrum of the cluster. From the initial luminosity function $\psi(M_V)$ given by \citet[Table~2 and Figure~2]{Sandage1957a} and the mass-luminosity relation of \citet{Kuiper1942}, we obtain the initial mass spectrum. The present-day spectrum is derived by assuming that all the stars heavier than $1.44 \U{M}_{\odot}$ have been evolving down to this value through their transformation into white dwarves. The effect of escape can be neglected. We find
\begin{eqnarray}
\overline{m} &=& 0.353 \U{M}_{\odot},\\
\overline{m^2} &=& 0.265 \U{M}_{\odot},\nonumber\\
\overline{m^2}/\overline{m} &=& 0.75 \U{M}_{\odot},\nonumber
\end{eqnarray}
in very good agreement with the value (8.8).
This previous finding is quite sensitive to the mass chosen for the white dwarves, $m_b$ \tn{The subscript $b$ stands for \emph{naine blanche}: white dwarf.}. For $m_b = 1\U{M}_{\odot}$, we find $\overline{m^2}/\overline{m} = 0.61 \U{M}_{\odot}$; for $m_b = 2\U{M}_{\odot}$, we find $\overline{m^2}/\overline{m} = 0.98 \U{M}_{\odot}$. The comparison of the two findings seems to confirm that the white dwarves have the limit mass of Chandrasekhar, $m_b = 1.44 \U{M}_{\odot}$. However, here again, the uncertainties of the theory and of the determination of the mass spectrum question the validity of this conclusion.
\subsubsection{Projected density: brightness measurements (47~Tuc)}
\citet{Gascoigne1956} have obtained the curves of projected density of the clusters 47~Tucanae and $\omega$~Centauri from a photometric method. The disadvantage of this procedure is to provide only the total brightness, without distinguishing the stars of different magnitudes; however, it allows for a precise measurement of the density from the center of the cluster up to quite remote distance.
The case of $\omega$~Centauri will be studied later. \reff{21} plots the observed values for 47~Tuc = NGC~104 and the closest theoretical curve, which corresponds to:
\begin{eqnarray}
1' &=& 0.088 \textrm{ canonical units},\\
\mu &=& 1.33,\nonumber\\
C &=& 0.555 \textrm{ stars of the tenth magnitude per minute}.\nonumber
\end{eqnarray}
The external radius is then:
\begin{equation}
r_e = 53.4'.
\end{equation}
\begin{figure}
\includegraphics[width=\columnwidth]{f21}
\caption{47~Tuc: comparison between observations (dots) and theory (line).}
\label{fig:21}
\end{figure}
The value of $\mu$ found here can been seen as a particular ``mean mass'', obtained by weighting the stars proportionally to their luminosity. It is only slightly lighter than the mass of the heaviest stars ($\mu\simeq 1.45$ from the results obtained for M3) which contribute to the major part of the total brightness.
\reff{21} shows good agreement (the small systematic deviations of the points from the theoretical curve can be explained: the data of Gascoigne \& Burr does not directly come from observations, but has rather been read on a smoothed curve). Near the center, the theoretical curve is too high, for reasons already listed in relation to M3.
Gascoigne \& Burr have measured the total luminosity in a series of circles of small radius around the center of the cluster. For each circle, we can compute the mean value of $\rho_P r$, by using (8.3), with $r_1 = 0$; the crosses on \reff{21} mark the value obtained. We clearly see that $\rho_P r$ tends toward a well-defined central value. This seems to be a good confirmation of the shape predicted by the theory near the center, and in particular of the fact that the central density is infinite\footnote{However, \citet{King1961b} has observed the central region of several clusters in detail and concluded that a high but finite central density exists.}
\subsection{Mass-radius relation}
The relation (3.7) links the mass of a cluster with its radius and the mean value of the negative curvature of the galactic field along its orbit. Unfortunately, the exact motions of the globular clusters in the Galaxy are unknown; the radial velocities can be measured with a good precision \citep{Mayall1946, Kinman1959}, but the proper motions stand at the limit of the observational possibilities, and have been obtained for 9 clusters only \citep{Gamalej1948}, with a low precision. It seems even not possible to know wether the orbits of the clusters are in average rather radial or circular \citep{vonHoerner1955, Kurth1960}. That is why we will consider the curvature of the galactic field to be the same for all the clusters, for want of anything better.
The radius and the mass of the globular clusters should be linked, at least approximately, through a relation like (3.8). We are going to check this using observational data. To limit the effect of observational errors, we will only use a series of homogeneous observations, done by the same author, under the same circumstances.
The distances have been taken from \citet{Kinman1958}. In order of preference, we take the distances derived from the position of the main sequence, from the magnitude of the variable stars, or from the magnitude of the 25 brightest stars. If none of these three estimates exists, the cluster is not considered.
\citet{Christie1940} made very precise measurements of the apparent magnitudes of the clusters. From them, we derive the absolute photographic magnitudes $M_{pg}$; the possible effect of absorption is cancelled, because the distances themselves are derived from the apparent magnitudes of the stars.
Then, we assume that the ratio of the total mass to the total luminosity is the same for all the globular clusters; by taking the mass estimate of M3 by \citet{Sandage1957a} as reference, we find that the mass $\mc{M}_e$ of a cluster is given as a function of its absolute magnitude $M_{pg}$ by
\begin{equation}
\log{\mc{M}_e} = -0.4\ M_{pg} + 2.00.
\end{equation}
This is confirmed by the study of \citet{Zeliakh1957} who, after a detailed discussion on several clusters, arrived at a quite similar value for the constant: 1.94 instead of 2.00.
Finally, we take the apparent external radii measured by \citet{Shapley1935}; they appeared to be more precise that those published more recently by \citet{Mowbray1946}. To derive the real radii, we should, in principle, use the distances corrected for absorption. However, as indicated by \citet{Shapley1935, Shapley1949}, absorption induces two opposite effects: an overestimate of the distance and an underestimate of the apparent radius. A quantitative study shows that these two effects balance each other almost exactly \citep{Parenago1949, Lohmann1952}. Therefore, we can completely neglect the possible existence of absorption. For that matter, this is very fortunate because the absorption is not quite well known.
\reft{6} gives the list of the 35 clusters for which the three measurements exist, with their mass (in solar masses), and their radius (in parsecs). The mass-radius diagram is plotted in \reff{22}: we see that, despite the dispersion of the points, a defined relation arises. The main source of error comes from the estimation of the external radius, which is difficult to observe precisely; this error alone is enough to explain the dispersion of the points, and it is even remarkable that it does not induce a larger dispersion. (This shows the quality of the measurements of Shapley \& Sayer; if one uses the radii from \citealt{Mowbray1946}, the dispersion is much larger.)
\begin{table}
\caption{}
\label{tab:6}
\begin{tabular}{ccc}
NGC & $\log{(\mc{M}_e)}$ & $\log{(r_e)}$ \\
\hline
\hline
288 & 4.64 & 1.40\\
1904 & 5.12 & 1.37\\
2419 & 5.08 & 1.71\\
4147 & 4.44 & 1.03\\
5024 & 5.12 & 1.66\\
5139 & 5.80 & 1.90\\
5272 & 5.40 & 1.64\\
5897 & 4.60 & 1.50\\
5904 & 5.28 & 1.60\\
6093 & 5.04 & 1.52\\
6121 & 4.48 & 1.30\\
6171 & 4.36 & 1.34\\
6205 & 5.20 & 1.38\\
6218 & 4.80 & 1.48\\
6229 & 4.88 & 1.38\\
6254 & 5.04 & 1.53\\
6266 & 5.28 & 1.60\\
6273 & 5.04 & 1.50\\
6284 & 4.56 & 1.49\\
6293 & 4.84 & 1.39\\
6333 & 5.00 & 1.53\\
6356 & 5.16 & 1.70\\
6402 & 5.00 & 1.69\\
6626 & 4.84 & 1.46\\
6638 & 4.88 & 1.32\\
6656 & 5.08 & 1.54\\
6715 & 5.36 & 1.71\\
6723 & 5.00 & 1.30\\
6779 & 4.68 & 1.38\\
6809 & 4.84 & 1.46\\
6864 & 5.48 & 1.74\\
6981 & 4.64 & 1.42\\
7078 & 5.44 & 1.60\\
7089 & 5.56 & 1.63\\
7099 & 4.76 & 1.29\\
\hline
\end{tabular}
\end{table}
\begin{figure}
\includegraphics[width=\columnwidth]{f22}
\caption{Mass-radius relation of the globular clusters.}
\label{fig:22}
\end{figure}
In addition, the small dispersion justifies the hypothesis of constant curvature of the galactic field for all the clusters.
The largest error being on $r_e$, we shall compute the means in the horizontal direction. This way, we find that the data points are well fitted with the relation
\begin{equation}
\log{(r_e)} =\frac{1}{3} \log{(\mc{M}_e)} - 0.17\ (\pm 0.11).
\end{equation}
i.e.,
\begin{equation}
\frac{\mc{M}_e}{r_e^3} = 3.2 \U{M_\odot/pc^3},
\end{equation}
in agreement with the theoretical formula (3.8).
The isolated point in the bottom-left corner of the figure is NGC~4147. This cluster is easily distinguishable from the others; Shapley \& Sayer qualify it as ``abnormal''. The point on the top-right corner is $\omega$~Cen, also abnormal, as we will see later.
\subsubsection{Correction of the external radii}
The observed external radius is, in fact, systematically smaller than the real radius. Indeed, the projected density decreases very rapidly toward the outskirts and becomes very small long before reaching the boundary of the cluster. This clearly appears in the artificial cluster of \reff{15}: a direct estimate of the radius of this cluster, from its aspect, gives a value of the order of 0.6 times the real radius, and this without any background effect. \reft{1} indicates that only 3\% of the total mass of the cluster is situated (in projection) beyond half of the radius. That is, even without absorption, the radius may be underestimated by a ratio of about 1/2. A good confirmation of this value can be obtained in the case of the two clusters studied above: for M3 and 47~Tuc, the observed radii are 11' and 28.2', i.e. 0.49 and 0.53 times the real radii derived from the theory and the projected density profiles, and given in (8.6) and (8.11), respectively. Furthermore, there is no absorption for these clusters.
Thus, we will systematically make a correction, assuming that \emph{the real radii are twice as large as the observed radii}, given by Shapley \& Sayer. As a consequence, the relation (8.14) must be replaced with
\begin{equation}
\frac{\mc{M}_e}{r_e^3} = 0.40 \U{M_\odot/pc^3}.
\end{equation}
This relation between the mass and the real external radius is plotted in \reff{22} as a dashed line. All the observed radii stand on its left side, as expected.
\subsubsection{Curvature of the galactic field}
When comparing (3.7) and (8.15), we find the mean curvature of the galactic field for the clusters:
\begin{equation}
\overline{\frac{\partial^2 U_G}{\partial x^2}} = -5.4\times 10^{-16} \U{yr}^{-2} = -520 \U{km^2 s^{-2} kpc^{-2}}.
\end{equation}
For comparison, we can compute the curvature in the solar neighborhood; it derives from the Oort's constants $A$ and $B$ (the numerical values are taken from \citealt{Allen1955}):
\begin{equation}
\frac{\partial^2 U_G}{\partial x^2} = (B-A) (3A+B) = -1370 \U{km^2 s^{-2} kpc^{-2}}.
\end{equation}
The mean curvature of the galactic field for the globular clusters is thus about 3 times smaller than in the solar neighborhood. This result is plausible, because the clusters travel in average at a quite long distance from the center of the Galaxy, and not in its plane. A more detailed comparison of the value (8.16) with the known structure of the Galaxy would be interesting, but is out of the scope of this work.
\subsection{Evolution}
Knowing the mass $\mc{M}_e$ and the radius $r_e$ of a real cluster, we can compute the parameters $\beta$ and $\gamma$ of the homology, using the transformation relations (6.31b), (6.34a), (2.32c), (2.43a). We get
\begin{eqnarray}
\beta &=& G \left(\frac{\mc{M}_e}{\fb{M}_e}\right)\left(\frac{r_e}{\fb{R}_e}\right)^{-1},\\
\gamma &=& (16\pi^2\ \overline{m})^{-1}\ G^{-3/2}\ \left(\frac{\mc{M}_e}{\fb{M}_e}\right)^{-1/2} \left(\frac{r_e}{\fb{R}_e}\right)^{-3/2}.\nonumber
\end{eqnarray}
This allows us to go from any quantity of the homologous model to the corresponding physical quantity, using the transformation formulae (2.31), (2.32), (2.43), (2.48), (6.31) and (6.34). In particular, we obtain:
\begin{equation}
\frac{\ensuremath{\mathrm{d}} t}{\ensuremath{\mathrm{d}} \fb{T}} = G^{-1/2} \frac{\overline{m}}{\overline{m^2}}\ \frac{1}{\ln{(n)}} \left(\frac{\mc{M}_e}{\fb{M}_e}\right)^{1/2} \left(\frac{r_e}{\fb{R}_e}\right)^{3/2}.
\end{equation}
This relation gives the time scale of the evolution of the cluster, because $\fb{T}$ is, as we have seen (Chapter~V), a parameter that measures the level of evolution of the cluster, while $t$ is the real time. $\fb{M}_e$ and $\fb{R}_e$ are given in (5.5). We use the value $\overline{m^2}/\overline{m}$ found from the observation of M3 (Equation 8.8). The factor $\ln{(n)}$ does not vary much for the different globular clusters; we use the typical value $n=10^5$. Finally, the radius is linked to the mass through (8.15), found from the observations. Thus, we find the very simple relation:
\begin{equation}
\frac{\ensuremath{\mathrm{d}} t}{\ensuremath{\mathrm{d}}\fb{T}} = 1.76 \times 10^5\ \mc{M}_e,
\end{equation}
where $t$ is in years and $\mc{M}_e$ in solar masses.
(6.22) provides the escape rate of the stars of given mass, which also reads
\begin{equation}
\frac{1}{n_m}\frac{\ensuremath{\mathrm{d}} n_m}{\ensuremath{\mathrm{d}}\fb{T}} = -\theta.
\end{equation}
Recall that $n_m$ is the number of stars whose mass is between $m$ and $m+\ensuremath{\mathrm{d}} m$. $\theta$ is given in \reft{2} and can be fitted with the relation:
\begin{equation}
\theta \simeq 0.4078\ (3-2\mu).
\end{equation}
Then, we can compute the variation of the total mass, by using (6.25b), (6.32) and (6.26). We find
\begin{equation}
\frac{\ensuremath{\mathrm{d}}\mc{M}_e}{\ensuremath{\mathrm{d}}\fb{T}}=-0.4078\ \mc{M}_e.
\end{equation}
We note that this result does not depend on the mass function.
When comparing with (8.20), the mass $\mc{M}_e$ cancels, and we get
\begin{equation}
\frac{\ensuremath{\mathrm{d}}\mc{M}_e}{\ensuremath{\mathrm{d}} t}=-2.3 \times 10^{-6} \U{M_\odot/yr}.
\end{equation}
Thus: \emph{the mass of a globular cluster decreases by $2300 \U{M}_{\odot}$ per billion years}; this rate is almost the same for all the globular clusters, and it also remains constant with time. The simplicity of this result is to be noted.
In fact, the rate is not exactly constant. The term $\ln{(n)}$ changes with time, which would lead to slight slowdown of the evolution near the end. On the other hand, the term $m_0 = \overline{m^2}/\overline{m}$ increases because the lightest stars escape faster. This effect is opposite to the previous one and likely stronger. For example, in the cluster M3 which is still at the beginning of its evolution, the value observed is $m_0 = 0.80 \U{M}_{\odot}$; near the end of its evolution (in about $100 \U{Gyr}$), only the heaviest stars will remain and we will get $m_0 = 1.44 \U{M}_{\odot}$, thus a multiplication of the evolution rate by a factor 1.8.
\subsection{Mass function of the globular clusters}
The evolution law of the masses of the globular clusters partially sets the present-day mass distribution of these masses. We are going to see whether a confirmation of (8.24) can be obtained this way.
From the apparent magnitudes and the apparent distance moduli provided by \citet{Lohmann1952} for 94 clusters, we compute the absolute magnitudes and then the masses using (8.12).
The mass function we obtain is plotted in \reff{23}, using the dotted line on the left-hand side. The two most massive clusters are off the figure: M2 and $\omega$~Cen of masses $3.4\times 10^{5} \U{M}_{\odot}$ and $5.2\times 10^{5} \U{M}_{\odot}$. We must also investigate the effect of observational selection. By looking at the apparent magnitudes $m_{pg}$ of the clusters, we see a clear upper limit at $m_{pg} = 12.5$; we assume that this value marks the highest observable magnitude. As a consequence, the clusters of given absolute magnitude $M_{pg}$ are observed only if their apparent distance modulus is smaller than $12.5-M_{pg}$. That is, we can compute, for each value of $M_{pg}$, the observable fraction of the clusters, as soon as we know the distribution of clusters as a function of the apparent modulus. This distribution (that we assume to be independent of the absolute magnitude) is easily found thanks to the brightest clusters, for which observational selection does not play any role.
\begin{figure}
\includegraphics[width=\columnwidth]{f23}
\caption{Mass function of the globular clusters, observed (dotted line), corrected for observational selection (solid line), initial (dashed line). \tn{The upper axis reads \emph{$t$ en milliards d'ann\'ees}: $t$ in billion years.}}
\label{fig:23}
\end{figure}
The solid line in \reff{23} is the corrected mass function. It can be fitted with a smooth curve. In particular, we note that this curve takes a finite value for $\mc{M}_e = 0$. But if we assume that all the globular clusters formed in the past and that no more are born now; and if the evolution of the masses is as
\begin{equation}
\frac{\ensuremath{\mathrm{d}}\mc{M}_e}{\ensuremath{\mathrm{d}} t} \propto \mc{M}_e^\nu,
\end{equation}
we can easily find that the present-day mass function must be proportional to $\mc{M}_e^{-\nu}$ near the origin \tn{i.e. for $\mc{M}_e \to 0$}. Therefore, this mass function must be zero for $\mc{M}_e = 0$ if $\nu$ is negative, and become infinite if $\nu$ is positive; it can take a finite value only if $\nu=0$. Therefore, the observed mass function seems to confirm the evolution law (8.24).
The validity of this point is slightly weakened because it relies on the number of the least massive clusters, that are the least well observed; but we can argue that the correction for observational selection (\reff{23}) is not very important; the conclusion would have been the same without the correction.
Note also that (8.25) is not fanciful but in fact covers all the evolutionary models already studied by several authors. The formulae of \citet{Chandrasekhar1943b, Chandrasekhar1943c} assume $\nu = +1$; the model from \citet{King1958b} corresponds to $\nu = -2.5$ and the similar model of \citet{vonHoerner1958} to $\nu = -3.55$.
The evolution formula (8.24) translates into a simple shift of the mass function toward the left; or alternatively, into a shift of the zero-mass toward the right. The upper axis of \reff{23} indicates the position of this zero-mass at different times, past and future. In particular, the \tn{vertical} dashed line on the left-hand side marks the position of the zero-mass point at the time of birth of the clusters, assuming it occurred $25 \U{Gyr}$ ago (see the next Section). By extrapolating the observed curve a little, we obtain the initial mass function. We find that about 62 globular clusters have disappeared since the origin of the galaxy. Nowadays, one cluster disappears every $500 \U{Myr}$, in average. Every cluster that has survived has lost $57000 \U{M}_{\odot}$ as escaping stars since its birth. For all the clusters, the mass of the escapers since their birth is: $8 \times 10^6 \U{M}_{\odot}$.
In the future side, we find that half of today's clusters will be gone in $30 \U{Gyr}$; but the most massive ones will survive much longer: $\omega$~Centauri will reach $230 \U{Gyr}$.
Up to now, we have neglected the phenomenon of gas ejection by the massive stars. \citet{vonHoerner1958} showed that this phenomenon is only relevant at the beginning of the life of the cluster: 30\% of the initial mass is lost in the gaseous form within the first billion years, then only 6\% during the next $5 \U{Gyr}$. The results obtained above are therefore valid, as soon as we put aside the initial evolutionary phase, lasting about one billion years.
\subsection{Age of the globular clusters}
When a cluster reaches the homologous state, any sign of its past (in particular its initial structure and its age) is lost. That is, one should specially care about the clusters that still show significant differences with the homologous model. According to the theory (see Equation 8.20), the most massive clusters are those which evolve the slowest, therefore, those where the largest differences should remain. That is, we are interested in $\omega$~Centauri, the most massive of all globular clusters.
But precisely, $\omega$~Centauri yields a peculiar structure, very different from those of the other clusters. The difference already arises as its concentration class in Shapley's classification is VII, while all the other high luminosity clusters lie in classes I to V. The difference is even more visible in the work of \citet{Gascoigne1956} who observed $\omega$~Cen and 47~Tuc under the same circumstances. The Table~II and the Figure~2 from these authors perfectly show that the light from a small circle of radius $r$ surrounding the center of the cluster is proportional to $r^2$ in the case of $\omega$~Cen, and to $r$ for 47~Tuc. In other words: in 47~Tuc, the projected density goes as $1/r$ near the center; but in $\omega$~Cen, it is almost constant. \emph{The central density of $\omega$~Cen is finite}.
The theoretical interpretation comes immediately: 47~Tuc has already reached the stage in its evolution where the central density becomes infinite; $\omega$~Cen, being more massive, has not reached it yet. (It seems to be the only globular cluster in this case.)
This confirms the theory, and we now have a simple way of estimating the age of the clusters. When comparing (2.32e), (7.44), (7.45) and (8.20), we find that the time required for the infinite central density to appear is
\begin{equation}
t_2 = 0.713\times 10^5\ \mc{M}_{e_0} = 0.854\times 10^5\ \mc{M}_{e_2},
\end{equation}
where $t_2$ is in years, $\mc{M}_{e_0}$ is the initial mass of the cluster and $\mc{M}_{e_2}$ its mass at the time when the infinite central density appears. The present-day masses of 47~Tuc and $\omega$~Cen, obtained as in the previous Section, are: $2.8 \times 10^5 \U{M}_{\odot}$ and $5.2\times 10^5 \U{M}_{\odot}$. We conclude that the age of these clusters (assuming it is the same) must be between
\begin{equation}
24\times 10^9 \U{yr} \qquad \textrm{and}\qquad 44\times 10^9 \U{yr}.
\end{equation}
Furthermore, the most recent estimates of the age of the globular clusters from the stellar evolution theory \citep{Sandage1961} are
\begin{eqnarray}
22\times 10^9 \U{yr} && \textrm{for M13},\\
26\times 10^9 \U{yr} && \textrm{for M3 and M5}.\nonumber
\end{eqnarray}
That is, two completely independent methods lead to values in agreement for the age of the globular clusters. This result is very encouraging, both for the theory of the dynamical evolution of the clusters and for the theory of stellar evolution. It also emphasizes the discrepancy underlined by \citet{Sandage1961} between the ages of the globular clusters and the recent models of an expanding Universe.
According to theory, $\omega$~Cen should have more differences with respect to the homologous model; the first proper difference should still exist with a non negligible amplitude. We have seen (see \reff{13}) that this difference affects mostly the external radius. And we precisely note that that the radius of $\omega$~Cen is far too large, given its mass (\reff{22}).
Finally, the strong ellipticity of $\omega$~Cen is likely the remnant of an initial difference from spherical symmetry, that has not enough time to vanish yet.
\section{Conclusions}
We are going to summarize the major results obtained in this work, and then list the directions in which it would be good to develop the research, in order to improve the theory and extend the range of its applications.
\subsection{Results obtained}
By making two hypotheses: (1) isotropy of the velocities and (2) equal masses, we have obtained the system of equations (2.25), which, supplemented by the right boundary conditions, allows us to compute the evolution of a cluster from a given initial state. We have looked for an homologous solution, i.e. a model that remains self-similar, the evolution being limited to the scaling of the various physical quantities; we have found that this model exists and is unique. Its mass and its radius are finite. The projected density matches the observations as much as their precision allows. Near the center, the homologous model yields unexpected properties: the central density and the central potential are infinite; a continuous flux of negative energy goes toward the center where it is absorbed by the formation of multiple stars. Furthermore, the escape of stars through the boundary of the cluster leads to a linear decrease of the total mass with time.
Then, we have supposed that the main population of the cluster is mixed with a second population, much less numerous, of stars with a different mass, and we have computed the distribution of this secondary population and its escape rate. These results allow for an approximate solution in the case of any mass function. By comparing with the observations, we obtain in particular the approximate mass-luminosity relation of the stars in clusters.
Finally we have studied the evolution of clusters close to the homologous model. We found that if the central density is finite, it rapidly increases and becomes infinite after a certain time. Generally, any difference with the homologous model decreases with time; a cluster becomes almost identical to the homologous model after the first third of its lifetime. In particular, these results lead to an estimate of the age of the globular clusters.
\subsection{Desirable improvements and extensions}
\begin{enumerate}
\item The most debatable hypotheses of the present model is indubitably the isotropy of the velocities. We could, as a first step, make the distribution function more general by adding an anisotropic term, small compared to the main term; this would allow us to quantify the error made when using the isotropy hypothesis. In a more ambitious second step, we would eliminate any restriction for the distribution function $f(E,A)$; but in this case, the equation of local evolution would become extremely complex (see \citealt{Rosenbluth1957}, Equation~31).
\item In the study of the approach to the homologous model made in Chapter~VII, we assume that the cluster is already similar enough to the homologous model that we can linearize the equations of the differences. It would be very interesting to also study the evolution of clusters far from the homologous model; this would allow us to describe the beginning of the evolution of a cluster, from a given initial state. To do so, we would have to come back to the real \tn{physical} variables and directly solve (2.25), by computing step by step the successive forms taken by the cluster in time\footnote{A similar calculation has been done, in the much simpler case of a homogeneous plasma, by \citet{MacDonald1957}.}. The results could be applied to a detailed explanation of the structure of $\omega$~Cen and to a precise estimate of its age. They would also allow for a more rigorous demonstration of the statement that clusters tend toward the homologous model, whatever their initial state is.
\item The previous paragraph implies the knowledge of the initial state of the clusters, therefore the building of a more or less approximate theory for the formation of the clusters and the initial phase during which the steady state is established. We could assume, for example, that the protocluster is an homogeneous sphere with a density slightly higher than the critical density (therefore, favoring a contraction), with negligible internal velocities. The evolution equations for this initial phase would be completely different from those we have considered here; on the one hand, the steady state is not reached and thus the distribution function does not only depend on $E$ and $A$, and on the other hand, because this phase is very short, the effect of perturbations can be neglected. Fortunately this implies that there is no need for the knowledge of the time of formation of the stars.
\item The hypothesis of equal mass stars is likely to be far from reality. In Chapter~VI, we have obtained a more general solution but at the cost of a quite arbitrary approximation. Hence, we should extend the investigations to the case of any mass function. But then a big difficulty appears: the escape rates of the stars of different masses not being the same, their relative importance always changes and the structure of the cluster varies too; as a consequence \emph{a homologous solution cannot exist}. In other words, if one introduces the mass as an additional variable, one must also consider the time variable, and the complexity of the calculations is suddenly hugely increased.
There are two possibilities to overcome this. First, we could seek a better model than the homologous one, that would allow us to separate the time variable; this model would necessary count, in addition to the two dimensional parameters, one or more other parameters depending on time. Second, the mass function could be initially described as a sum of \tn{Dirac's} $\delta$ distributions. The simplest case of the sum of two $\delta$ distributions (i.e. a cluster made of the mix of stars of two different masses, in any proportion), would already be a great improvement and would allow us to estimate the actual effect of a spread of masses.
\item We have only considered the value $\lambda=1/3$ of the power of the radius-mass relation (3.11), with the perspective of globular clusters. The calculations could be re-done with different values of $\lambda$, and in particular, with $\lambda = \infty$, which correspond to the isolated cluster, and seems to better fit the cases of galaxies and galaxy clusters. The other values of $\lambda$ do not seem to correspond to existing objects, but could be useful to describe some intermediate states.
\item The hypothesis of spherical symmetry is valid for most of the clusters; but some are clearly elliptical. Furthermore, we could foresee the application of the models to other objects (see point 8 below). Thus, we should study the non-spherical models. The ellipticity can originate from a global rotation; a first approximation would consider that the rotation is small and only implies a correction term. We would likely find a model similar to the homologous one, but slightly flattened. It would be particularly interesting to find whether the escaping stars take away a little or a lot of angular momentum, and thus, if the flattening increases or decreases.
We should also study the case of an angular momentum distribution without spherical symmetry, but still with no global rotation. Such an asymmetry would probably decrease, with a relaxation time of the same order as those found in Chapter~VII for the radial differences.
Finally, it would be good to consider the effect of the asymmetry in the galactic field (see Chapter~III). But this effect is likely not so important for the applications, because it only affects the external regions, which are invisible because of their very low density \citep{King1961a}.
\item From the observational side, it seems that much remains to be done. Accurate counts per luminosity class, as those made by Sandage for the cluster M3, would be very useful for other clusters. Furthermore, errors could probably be reduced by repeating the counts for several photographs of the same cluster. (\citealt{Tayler1954} noted big differences between the plates obtained with two different telescopes.)
The comparisons of accurate observation with a sufficiently elaborated theory would particularly allow for:
\begin{itemize}
\item a precise physical determination of the mass-luminosity relation; to push this relation as far as possible toward the low luminosities, one should particularly study the closest clusters. In the future, space observatories would allow us to reach much higher magnitudes;
\item an estimate of the age of the clusters, and perhaps information about their formations, when analyzing the differences between the present-day structure of the clusters and their theoretical final structure. However, these differences are small for most of the clusters, so that this analysis would require a high precision for the observations. One should start with the most massive clusters, where the largest differences remain;
\item a study of the galactic field and of the orbits of the globular clusters.
\end{itemize}
\item Here, we have only focussed on globular clusters. The models obtained could be applied to other objects: galactic clusters \tn{star clusters in the disk of the galaxy, as opposed to the globulars that are further away in the halo}, elliptical galaxies and cores \tn{central, spherical regions} of spiral galaxies, galaxy clusters. But in every case, difficulties appear, making the application of the theory more doubtful or more complex. That is, we think that the theory should first focus on describing properly the relatively simple case of globular clusters (which it is still far from doing) before extending its ambitions to more involved problems.
The galactic clusters would be extremely interesting, for many reasons: being closer, less massive stars are visible there; their ages are not all the same and thus we can observe all the phases of the dynamical evolution; and it is possible to measure individual velocities. Unfortunately, there are two major difficulties: the number of stars is small, so that statistical fluctuations forbid an accurate estimate of the projected density and of the other quantities; and above all, as \citet{Spitzer1958a} showed, the evolution of the cluster is largely influenced by passages near interstellar clouds.
\item Finally, we note some minor issues: the enhancement of the computation of the perturbations between the stars; the study of the accumulation of central energy as multiple stars and the final evolution of the cluster (see the end of Chapter~V); the influence of the motion of the cluster within the galaxy on the escape of stars (this could be simplified as a three-body problem: galactic center, cluster, star).
\end{enumerate}
One sees that there is no shortage of work. Stellar dynamics is a rising science, where almost everything remains to be done. It looks today much more like a collection of isolated attempts than like a homogeneous doctrine. These attempts have only led to isolated results, with limited range; the major problems, far from being solved, have remained almost untouched.
As we have seen, the issues do not arise from a poor knowledge of the physical mechanisms involved, but only from the complexity of the calculations. But today, a new, extremely powerful, weapon exists to overcome this kind of issues: the electronic computers. That is, we think that stellar dynamics should now leave the relative state of neglect where it has been left and spark off the interest and the long-term efforts of a growing number of researchers. It seems worthless to underline the importance of the results that could be obtained: the understanding of the collective dynamical phenomena is directly linked to the solving of the major cosmogonical and cosmological problems, and thus to our understanding of the Universe. We hope that the present essay, although it has a very limited scope, will contribute to the illumination of the new perspectives and give to others the wish to explore them.
\section*{Acknowledgments}
My acknowledgements go to Prof.~A.~Danjon, director of the Institut d'Astrophysique; to Prof.~E.~Schatzman, my supervisor from whose teaching I got the initial idea of this work, and who always encouraged and helped me; to Messieurs L.~Malavard and J.F.~Denisse, who accepted to be my research sponsors; to J.~Arsac, who created and organized the Service de Calcul Num\'erique de l'Observatoire de Meudon and taught my colleagues and myself the computational techniques on an electronic device; to the staff of this department, as well as the Bureau de Calcul de l'Institut d'Astrophysique, to Mrs Hernandez, Mrs Lagorce, Messieurs Charbey and de Postel, who assisted me for the technical realization of this work.
\onecolumn
\begin{center}
List of the main notations\\
The number is those of the equation where the notation is used first.\\
\begin{tabular}{ll|ll|ll|ll|ll|ll}
\hline
\hline
$a$ & 2.9 & $F$ & 2.23 & $L_e$ & 5.5 & $q$ & 2.19 & $t$ & 2.1 & $\fb{Z}$ & 5.1 \\
$A$ & 2.2 & $\fb{F}$ & 2.31 & $\mc{L}$ & 2.39 & $Q$ & 2.24 & $T$ & 2.24 & & \\
$b$ & 2.36 & $F_2$ & 6.1 & $m$ & 2.1 & $\fb{Q}$ & 2.32 & $\fb{T}$ & 2.32 & $\alpha_1$ & 4.3 \\
$c$ & 2.36 & $\fb{F}_2$ & 6.5 & $M$ & 2.41 & $r$ & 2.7 & $U$ & 2.3 & $\alpha_2$ & 4.3 \\
$c_2$ & 6.6 & $g$ & 2.12 & $\fb{M}$ & 2.43 & $r_e$ & 2.8 & $\fb{U}$ & 2.32 & $\beta$ & 2.31 \\
$C$ & 6.10 & $H$ & 4.26 & $\fb{M}_e$ & 3.12 & $r_m$ & 2.13 & $\fb{U}_e$ & 3.5 & $\gamma$ & 2.31 \\
$D$ & 2.24 & $\fb{H}$ & 4.31 & $\fb{M}_P$ & 2.51 & $R$ & 2.24 & $U_G$ & 3.1 & $\gamma_2$ & 6.5 \\
$\fb{D}$ & 2.32 & $K$ & 4.33 & $\fb{M}_\fb{R}$ & 2.45 & $\fb{R}$ & 2.32 & $\fb{U}_0$ & 7.1 & $\theta$ & 6.22 \\
$D_P$ & 2.47 & $K_2$ & 7.10 & $\mc{M}$ & 2.38 & $\fb{R}_e$ & 3.12 & $\fb{U}_{\infty}$ & 5.7 & $\lambda$ & 3.11 \\
$\fb{D}_P$ & 2.48 & $K_D$ & 4.7 & $\mc{M}_e$ & 3.3 & $s$ & 7.23 & $v$ & 2.3 & $\mu$ & 6.4 \\
$E$ & 2.2 & $L$ & 2.41 & $n$ & I & $S$ & 2.27 & $x$ & 2.1 & $\rho$ & 2.5 \\
$\fb{E}$ & 2.31 & $\fb{L}$ & 2.43 & $n_m$ & 6.24 & $\fb{S}$ & 2.32 & $y$ & 2.1 & $\rho_p$ & 2.46 \\
$f$ & 2.2 & & & $p$ & 6.21 & & & $z$ & 2.1 & & \\
\hline
\end{tabular}\\
The symbol $'$ indicates the derivation with respect to $E$ or $\fb{E}$.\\
The subscript $e$ corresponds to the values taken by the quantities at the boundary of the cluster.\\
The symbols in capital letters ($F$, $D$, $R$, ...) generally represent the ``normalized variables''; the bold font ($\fb{F}$, $\fb{D}$, $\fb{R}$, ...) represent the ``canonical variables'' (see Chapter~II).
\end{center}
\begin{multicols}{2}
|
\section{Introduction}
The QCD Lagrangian is known and well verified by particle experiments over many orders of magnitude \cite{Cacciari:1998it,Jager:2002xm,deFlorian:2002az}. But just as the collective behavior of electrically charged objects---take for instance the phase diagram of water---is far from obvious given the QED Lagrangian, the bulk dynamics of QCD are not well understood \cite{LRP}. With the Relativistic Heavy Ion Collider (RHIC), the Large Hadron Collider (LHC), and the soon-to-be-running Facility for Antiproton and Ion Research (FAIR) facilities the world scientific community has a novel opportunity to experimentally explore the state of the universe a few microseconds after the Big Bang and the phase transition from the usual, confining hadronic matter to deconfined quarks and gluons by colliding heavy nuclei, such as gold and lead, at near the speed of light. In particular rare high momentum partons, those that produce jets of particles, provide the most direct probe of the fundamental degrees of freedom in the new phase of QCD matter created at a few times the transition temperature, $T_c$, at these cutting-edge facilities \cite{Gyulassy:2004zy,Wiedemann:2009sh,Majumder:2010qh}. \fig{fig:diagram} shows a cartoon of the energy loss of these \highpt probes as they propagate through a deconfined plasma of quarks and gluons in (a) completely perturbative---calculable via the methods of perturbative QCD (pQCD)---and (b) completely nonperturbative---calculable via the methods of AdS/CFT---pictures. The goal, then, is to compare theoretical predictions to data to determine the properties of the quark-gluon plasma created in heavy ion collisions.
\begin{figure}[!htbp]
\centering
$\begin{array}{cc}
\includegraphics[width=.48 \columnwidth]{pQCDElossCartoon.png} &
\includegraphics[width=.48 \columnwidth]{AdSCFTElossCartoon.png} \\[-10pt]
\hspace{-2in}\mathrm{(a)} & \hspace{-2in}\mathrm{(b)}
\end{array}$
\caption{\label{fig:diagram}
Cartoon of the production, in-medium energy loss, and fragmentation processes that may occur using (a) pQCD for in-medium energy loss or (b) AdS/CFT for in-medium energy loss for a \highpt parton produced in a heavy ion collision. Figures adapted from \cite{Horowitz:2009eb} and \cite{Friess:2006fk}.
}
\end{figure}
\section{Results}
The suppression of \highpt particles in heavy ion collisions is usefully expressed in terms of \raacomma: the ratio of measured yield in nucleus-nucleus (A + A) collisions divided by the yield in p + p collisions scaled by the expected number of p + p collisions in an A + A collision. Extracting information from \raacomma, and suppression observables in general, is made nontrivial by the many physics processes---some of which are interesting in their own right---that are effectively integrated out. Of greatest import for our discussion is the initial distribution of \highpt quarks and gluons in both coordinate and momentum space. At RHIC, reference spectra from p + p collisions at the same $\sqrt{s}$ as in Au + Au collisions \cite{Adams:2005ph,Adare:2011vy} allow for a precise measurement of \raa \cite{Adams:2005ph,Adare:2008cg} and constrain the initial quark and gluon spectra; there is also abundant evidence from the d + Au collision data that at midrapidity the initial \highpt partonic spectrum is not significantly modified from that in p + p collisions \cite{Adams:2003im,Adler:2006wg}, and direct photon measurements reassure us that the binary distribution from a Glauber model approximates well the number of hard proton-proton-like collisions in a heavy ion event \cite{Adler:2005ig}.
These measurements do not yet exist at LHC, critically limiting the precision of---and our ability to interpret---the recent \raa results \cite{Aamodt:2010jd}, shown in \fig{fig:RAA} (a). First, there is no reference p + p data, which leads to about a factor of 2 $p_T$-dependent systematic uncertainty in the normalization of the reported \raacomma. Second, there is no ``control measurement'' from a p + Pb run. This control measurement is crucial as there are predictions of a significant \emph{initial state} suppression of \highpt particles due to unitarizing effects in the initial gluon parton distribution function \cite{Albacete:2010bs}; see \fig{fig:CGC} (a). Although not shown in that work, these effects should be $x$ and $Q^2$ dependent (the effects should disappear by $\eqnpt\sim25-50$ GeV/c) naturally introducing a rise in \raa as a function of \ptcomma, perhaps similar to that seen currently in the data. The ATLAS measurement \cite{Collaboration:2010px} consistent with a lack of suppression in its 38 Z boson candidates is encouragingly suggestive that the binary scaling seen at RHIC also holds at LHC, although the result is currently inconclusive due to the large systematic uncertainty and lack of statistics.
\begin{figure}[!htbp]
\centering
$\begin{array}{ccc}
\includegraphics[width=.3 \columnwidth]{Raa_peri_5.png} & \hspace{1in} &
\includegraphics[width=.35 \columnwidth]{WHDGpredictionsLHC.png} \\[-10pt]
\hspace{-2in}\mathrm{(a)} & \hspace{1in} & \hspace{-2in}\mathrm{(b)}
\end{array}$
\caption{\label{fig:RAA}
(a) Measured $R_{AA}$ of charged hadrons at 2.76 TeV at LHC \cite{Aamodt:2010jd}. (b) Predictions from the WHDG energy loss model \cite{Wicks:2005gt} for the $R_{AA}$ of pions at 5.5 TeV \cite{Horowitz:2007nq}.
}
\end{figure}
There is currently debate over even a qualitative understanding of the energy loss mechanism(s) in a QGP \cite{Horowitz:2007su,Marquet:2009eq,Horowitz:2010yi,Jia:2011pi}. Current pictures assume that either all interactions are strongly-coupled, some are strong and some are weak, or that all interactions are weak (see \cite{Horowitz:2007su,Majumder:2010qh} and references therein). It is worth noting, though, that none of the purely perturbative approaches quantitatively describes \emph{simultaneously} any two single particle \highpt observables at RHIC.
Nevertheless, the new ALICE data \cite{Aamodt:2010jd} shows a feature that appears strikingly similar to what one would expect from pQCD-based energy loss: \raa rises significantly as a function of \ptcomma. One expects a rise as a function of momentum as the fractional energy loss of a \highpt parton goes as $\epsilon\sim\log(\eqnpt)/\eqnpt$ \cite{Gyulassy:2000fs}, where the final momentum, $\eqnpt^f$, is related to the initial momentum, $\eqnpt^i$, by $\eqnpt^f=(1-\epsilon)\eqnpt^i$. If particle production is well approximated by a power law, $dN/d\eqnpt\sim\eqnpt^{-n}$, then $\eqnraa \sim \langle (1-\epsilon)^{n-1} \rangle$ \cite{Horowitz:2010dm}. The suppression at RHIC is flat within the uncertainty of the experiment \cite{Adare:2008cg}. If the picture of perturbative energy loss is correct, this flatness is due to a coincidental cancellation between 1) the fraction of \highpt gluons to quarks, 2) the hardening of the production spectrum as a function of \ptcomma, and 3) the decrease in energy loss as a function of \ptcomma. At LHC, the production spectrum is much flatter than at RHIC while energy loss till decreases with \ptcomma; this then would lead to an \raa that increases with \ptcomma. One can see from \fig{fig:RAA} (b) that that is exactly what we find for the WHDG model of pQCD-based energy loss we study here.
Furthermore, ALICE reports \cite{Aamodt:2010pb} a factor $\sim2.2$ increase in the central charged particle multiplicity at LHC compared to RHIC. Assuming that the quark-gluon plasma medium density scales with charged particle multiplicity, it becomes a quantitative question whether an energy loss calculation can simultaneously describe the \raa normalization at both RHIC and LHC.
An exciting future measurement is of the ratio of suppression patterns of charm and bottom quarks, $R_{AA}^c/R_{AA}^b$, as a function of \ptcomma, shown in \fig{fig:CGC} (b). One readily notices how the interplay of mass and momentum scales, very different for perturbative and AdS/CFT energy loss calculations, manifests itself: in perturbative calculations the momentum loss per unit time scales as $dp_T/dt \sim -LT^3\log(p_T/M_Q)$ whereas in the AdS/CFT calculations $dp_T/dt \sim -(T^3/M_Q^2)p_T$, where $L$ is the pathlength traversed by the \highpt parton, $T$ is the temperature of the plasma, and $M_Q$ is the mass of the heavy quark.
\begin{figure}[!htbp]
\centering
$\begin{array}{ccc}
\includegraphics[width=.4 \columnwidth]{CGC.png} &
&
\includegraphics[width=.55 \columnwidth]{RcRb.png} \\[-10pt]
\hspace{-2in}\mathrm{(a)} &
& \hspace{-2in}\mathrm{(b)}
\end{array}$
\caption{\label{fig:CGC}
(a) Predictions of the suppression of gluons in top energy Pb + Pb LHC collisions from initial state effects \emph{only} \cite{Albacete:2010bs}. (b) The ratio of charm to bottom $R_{AA}$ should provide a clear indication of the dominant physics of \highpt energy loss \cite{Horowitz:2007su}.}
\end{figure}
\section{Conclusions and Outlook}
A new era in heavy ion physics started on November 8$^\mathrm{th}$, 2011 when LHC began colliding lead nuclei at $\sqrt{s}_{NN}$ = 2.76 TeV, an unprecedented center of mass energy. While the initial \highpt results have qualitative features that appear to reflect the dominance of perturbative physics, one must quantitatively compare theoretical predictions to data at both RHIC and LHC energies simultaneously. However, even before making this comparison, it is clear that even a qualitative scientific conclusion based on the measured suppression of charged hadrons will be difficult to reach due to the influence of possibly very large, currently experimentally unconstrained initial state effects and to the large systematic uncertainty in the initial spectrum of \highpt particles. Future measurements from p + p and p + A runs at the same $\sqrt{s}_{NN}$ will greatly improve the ability for the community to reach a scientific consensus; future measurements of heavy quark suppression at LHC will provide a novel qualitative tool for investigating the gross properties of the quark-gluon plasma, such as whether the medium is best described as a strongly coupled fluid or a weakly coupled plasma.
\providecommand{\href}[2]{#2}\begingroup\raggedright |
\section{Introduction}
\vspace{0.2 cm}
The refined analytic torsion was introduced by M. Braverman and T. Kappeler ([4], [5]) on an odd dimensional closed Riemannian manifold
with a flat bundle as an analytic analogue of the refined combinatorial torsion introduced by M. Farber and V. Turaev ([10], [11], [25], [26]).
Even though these two objects do not coincide exactly, they are closely related.
The refined analytic torsion is defined by using the graded zeta-determinant of the odd signature operator and is described as
an element of the determinant line of the cohomologies.
Specially, when the odd signature operator is defined by an acyclic Hermitian connection on a closed manifold, the refined analytic torsion is a complex number,
whose modulus is the Ray-Singer analytic torsion and the phase part is the $\rho$-invariant determined by the given odd signature operator
and the trivial odd signature operator acting on the trivial line bundle.
In the previous work ([14]) we introduced the well-posed boundary conditions ${\mathcal P}_{-, {\mathcal L}_{0}}$ and ${\mathcal P}_{+, {\mathcal L}_{1}}$
for the odd signature operator, which are complementary to each other and have similar properties as the relative and absolute boundary conditions.
We showed that the refined analytic torsion is well-defined under these boundary conditions on a compact oriented Riemannian manifold with boundary.
In this paper we discuss the gluing formula of the refined analytic torsion with respect to the boundary conditions
${\mathcal P}_{-, {\mathcal L}_{0}}$ and ${\mathcal P}_{+, {\mathcal L}_{1}}$
when the odd signature operator is given by an acyclic Hermitian connection.
In this case the refined analytic torsion consists of the Ray-Singer analytic torsion, the eta invariant
and the values of the zeta functions at zero. The gluing formula of the Ray-Singer analytic torsion with respect to the relative and absolute boundary conditions has been obtained by W. L\"uck ([21]), D. Burghelea, L. Friedlander and T. Kappeler in [9] (cf. [29]).
The gluing formula of the eta invariant
with respect to the Atiyah-Patodi-Singer (APS) boundary condition has
been studied by many authors, for instance, K. Wojciechowski ([32], [33]), U. Bunke ([7]), J. Br\"uning, M. Lesch ([6]), P. Kirk and M. Lesch ([17]).
To use these results we first compare the Ray-Singer analytic torsion subject to the boundary condition ${\mathcal P}_{-, {\mathcal L}_{0}}$ or
${\mathcal P}_{+, {\mathcal L}_{1}}$ with the Ray-Singer analytic torsion subject to the relative or the absolute boundary condition.
We next compare the eta invariant associated to the odd signature operator subject to ${\mathcal P}_{-, {\mathcal L}_{0}}$ or
${\mathcal P}_{+, {\mathcal L}_{1}}$ with the eta invariant subject to the APS boundary condition.
To compare the Ray-Singer analytic torsions we are going to use the BFK-gluing formula for zeta-determinants ([8], [18], [19]) and the adiabatic limit method.
To compare the eta invariants we are going to follow the method given in [6].
These comparison results together with the well known gluing formulas lead to our main result.
The boundary value problem and the gluing formula of the refined analytic torsion
have been already studied by B. Vertman ([27], [28]) but our method is completely different from what he presented.
\vspace{0.3 cm}
Let $(M, g^{M})$ be a compact oriented odd dimensional Riemannian
manifold with boundary $Y$, where $g^{M}$ is assumed to be a product metric near the boundary $Y$.
We denote the dimension of $M$ by $m = 2r - 1$. Suppose that $\rho : \pi_{1}(M) \rightarrow
GL(n, {\Bbb C})$ is a representation of the fundamental group and $E = {\widetilde M}
\times_{\rho} {\Bbb C}^{n}$ is the associated flat bundle, where ${\widetilde M}$
is a universal covering space of $M$. We choose a flat connection
$\nabla$ and extend it to a covariant differential
$$
\nabla : \Omega^{\bullet}(M, E) \rightarrow \Omega^{\bullet + 1}(M, E).
$$
\noindent
Using the Hodge star operator $\ast_{M}$, we define the involution
$\Gamma = \Gamma(g^{M}) : \Omega^{\bullet}(M, E) \rightarrow \Omega^{m - \bullet}(M, E)$ by
\begin{equation}\label{E:1.1}
\Gamma \omega := i^{r} (-1)^{\frac{q(q+1)}{2}} \ast_{M} \omega, \qquad \omega \in \Omega^{q}(M, E),
\end{equation}
\noindent
where $r$ is given as above by $r = \frac{m+1}{2}$.
It is straightforward to see that $\Gamma^{2} = \operatorname{Id}$.
We define the odd signature operator $\mathcal{B}$ by
\begin{equation}\label{E:1.2}
\mathcal{B}\ = \ \mathcal{B}(\nabla,g^M) \ := \ \Gamma\,\nabla \ + \ \nabla\,\Gamma:\,\Omega^\bullet(M,E)\ \longrightarrow \ \Omega^\bullet(M,E).
\end{equation}
\noindent
Then $\mathcal{B}$ is an elliptic differential operator of order $1$.
Let $N$ be a collar neighborhood of $Y$ which is isometric to $[0, 1) \times Y$.
Any $q$-form $\omega$ can be written , on $N$, by
$$
\omega = \omega_{\operatorname{tan}} + du \wedge \omega_{\operatorname{nor}},
$$
where $\omega_{\operatorname{tan}}$ and $\omega_{\operatorname{nor}}$ are the tangential and normal parts of $\omega$
and $du$ is the dual of the inward unit normal vector field $\partial u$ to the boundary $Y$ on $N$.
Then we have a natural isomorphism
\begin{equation}\label{E:1.3}
\Psi : \Omega^{p}(N, E|_{N}) \rightarrow C^{\infty}([0, 1), \Omega^{p}(Y, E|_{Y}) \oplus \Omega^{p-1}(Y, E|_{Y})), \quad
\Psi(\omega_{\operatorname{tan}} + du \wedge \omega_{\operatorname{nor}}) = (\omega_{\operatorname{tan}}, \hspace{0.1 cm} \omega_{\operatorname{nor}}).
\end{equation}
\vspace{0.2 cm}
\noindent
Using the product structure we can
induce a flat connection $\nabla^{Y} : \Omega^{\bullet}(Y, E|_{Y}) \rightarrow \Omega^{\bullet + 1}(Y, E|_{Y})$ from $\nabla$ and a Hodge
star operator $\ast_{Y} : \Omega^{\bullet}(Y, E|_{Y}) \rightarrow
\Omega^{m-1-\bullet}(Y, E|_{Y})$ from $\ast_{M}$.
We define two maps $\beta$, $\Gamma^{Y}$ by
\vspace{0.2 cm}
\begin{equation}\label{E:1.4}
\begin{aligned}
\beta & : \Omega^{p}(Y, E|_{Y}) \rightarrow \Omega^{p}(Y, E|_{Y}), \quad \beta(\omega) = (-1)^{p} \omega \\
\Gamma^{Y} & : \Omega^{p}(Y, E|_{Y}) \rightarrow \Omega^{m-1-p}(Y,
E|_{Y}),\quad \Gamma^{Y}(\omega) = i^{r-1} (-1)^{\frac{p(p+1)}{2}}
\ast_{Y} \omega.
\end{aligned}
\end{equation}
\vspace{0.2 cm}
\noindent
It is straightforward that
\begin{equation}\label{E:1.5}
\beta^{2} = \operatorname{Id}, \qquad \Gamma^{Y} \Gamma^{Y} = \operatorname{Id}.
\end{equation}
\noindent
Simple computation shows that
\begin{equation}\label{E:1.6}
\Gamma = i \beta \Gamma^{Y} \left( \begin{array}{clcr} 0 & -1 \\ 1 & 0 \end{array} \right), \qquad
\nabla = \left( \begin{array}{clcr} 0 & 0 \\ 1 & 0 \end{array} \right) \nabla_{\partial u} +
\left( \begin{array}{clcr} 1 & 0 \\ 0 & -1 \end{array} \right) \nabla^{Y}.
\end{equation}
\noindent
Hence the odd signature operator $\mathcal{B}$ is expressed, under the isomorphism (\ref{E:1.3}), by
\begin{equation}\label{E:1.8}
\mathcal{B} = - i \beta \Gamma^{Y} \left\{ \left( \begin{array}{clcr} 1 & 0 \\
0 & 1 \end{array} \right) \nabla_{\partial u} + \left( \begin{array}{clcr} 0 & -1
\\ -1 & 0 \end{array} \right) \left( \nabla^{Y} + \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \right) \right\}.
\end{equation}
\noindent
We denote
\begin{equation} \label{E:1.9}
{\mathcal \gamma} := - i \beta \Gamma^{Y} \left( \begin{array}{clcr} 1 & 0 \\ 0 & 1 \end{array} \right), \qquad
{\mathcal A} := \left( \begin{array}{clcr} 0 & -1 \\ -1 & 0 \end{array} \right) \left( \nabla^{Y} + \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \right)
\end{equation}
\noindent
so that $\mathcal{B}$ has the form of
\noindent
\begin{equation} \label{E:1.10}
\mathcal{B} = {\mathcal \gamma} \left( \partial_{u} + {\mathcal A} \right) \qquad \text{with} \quad {\mathcal \gamma}^{2} = - \operatorname{Id}, \quad
{\mathcal \gamma} {\mathcal A} = - {\mathcal A} {\mathcal \gamma}.
\end{equation}
\noindent
Since $\nabla_{\partial u} \nabla^{Y} = \nabla^{Y} \nabla_{\partial u}$, we have
\begin{equation}\label{E:1.11}
\mathcal{B}^{2} = - \left( \begin{array}{clcr} 1 & 0 \\ 0 & 1 \end{array} \right) \nabla_{\partial u}^{2} +
\left( \begin{array}{clcr} 1 & 0 \\ 0 & 1 \end{array} \right) \left( \nabla^{Y} + \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \right)^{2} =
\left( - \nabla_{\partial u}^{2} + \mathcal{B}_{Y}^{2} \right) \left( \begin{array}{clcr} 1 & 0 \\ 0 & 1 \end{array} \right),
\end{equation}
\noindent
where
$$
\mathcal{B}_{Y} = \Gamma^{Y} \nabla^{Y} + \nabla^{Y} \Gamma^{Y}.
$$
We next choose a Hermitian inner product $h^{E}$.
All through this paper we assume that $\nabla$ is a Hermitian connection with respect to $h^{E}$,
which means that $\nabla$ is compatible with $h^{E}$, {\it i.e.}
for any $\phi$, $\psi \in C^{\infty}(E)$,
$$
d h^{E}(\phi, \psi) = h^{E}(\nabla \phi, \psi) + h^{E}(\phi, \nabla \psi).
$$
\noindent
The Green formula for $\mathcal{B}$ is given as follows (cf. [14]).
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:1.1}
(1) For $\phi \in \Omega^{q}(M, E)$, $\psi \in \Omega^{m-q}(M,E)$,
$\hspace{0.2 cm} \langle \Gamma \phi, \hspace{0.1 cm} \psi \rangle_{M} \hspace{0.1 cm}
= \hspace{0.1 cm} \langle \phi, \hspace{0.1 cm} \Gamma \psi \rangle_{M}$. \newline
(2) For $\phi \in \Omega^{q}(M, E)$, $\psi \in \Omega^{q+1}(M,E)$,
$$
\langle \nabla \phi, \hspace{0.1 cm} \psi \rangle_{M} \hspace{0.1 cm} = \hspace{0.1 cm}
\langle \phi, \hspace{0.1 cm} \Gamma \nabla \Gamma \psi \rangle_{M} \hspace{0.1 cm} - \hspace{0.1 cm}
\langle \phi_{\operatorname{tan}}|_{Y}, \hspace{0.1 cm} \psi_{\operatorname{nor}}|_{Y} \rangle_{Y}.
$$
(3) For $\phi$, $\psi \in \Omega^{\operatorname{even}}(M, E)$ or $\Omega^{\operatorname{odd}}(M,E)$,
$$
\langle \mathcal{B} \phi, \hspace{0.1 cm} \psi \rangle_{M} \hspace{0.1 cm} - \hspace{0.1 cm} \langle \phi, \hspace{0.1 cm} \mathcal{B} \psi \rangle_{M} = \hspace{0.1 cm}
- \langle \phi_{\operatorname{tan}}|_{Y}, \hspace{0.1 cm} i \beta \Gamma^{Y} (\psi_{\operatorname{tan}}|_{Y}) \rangle_{Y} \hspace{0.1 cm} -
\langle \phi_{\operatorname{nor}}|_{Y}, \hspace{0.1 cm} i \beta \Gamma^{Y} (\psi_{\operatorname{nor}}|_{Y}) \rangle_{Y}
\hspace{0.1 cm} = \hspace{0.1 cm} \langle \phi|_{Y}, \hspace{0.1 cm} ( {\mathcal \gamma} \psi)|_{Y} \rangle_{Y}.
$$
\end{lemma}
\vspace{0.2 cm}
\noindent
{\it Remark} : In the assertions (2) and (3) the signs on the inner products on $Y$ are different from those in [14] because in [14]
$\partial u$ is an outward unit normal vector field.
\vspace{0.2 cm}
We note that $\mathcal{B}_{Y}$ is a self-adjoint elliptic operator on $Y$.
Putting ${\mathcal H}^{\bullet}(Y, E|_{Y}) := \operatorname{ker} \mathcal{B}^{2}_{Y}$, ${\mathcal H}^{\bullet}(Y, E|_{Y})$ is a finite dimensional vector space and we have
$$
\Omega^{\bullet}(Y, E|_{Y}) = \operatorname{Im} \nabla^{Y} \oplus \operatorname{Im} \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \oplus {\mathcal H}^{\bullet}(Y, E|_{Y}).
$$
\noindent
If $\nabla \phi = \Gamma \nabla \Gamma \phi = 0$ for $\phi \in \Omega^{\bullet}(M, E)$, simple computation shows that $\phi$ is expressed,
near the boundary $Y$, by
\begin{equation} \label{E:1.12}
\phi = \nabla^{Y} \phi_{\operatorname{tan}} + \phi_{\operatorname{tan}, h} + du \wedge ( \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \phi_{\operatorname{nor}} + \phi_{\operatorname{nor}, h}), \qquad
\phi_{\operatorname{tan}, h}, \hspace{0.1 cm} \phi_{\operatorname{nor}, h} \in {\mathcal H}^{\bullet}(Y, E|_{Y}).
\end{equation}
\noindent
We define ${\mathcal K}$ by
\begin{equation} \label{E:1.13}
{\mathcal K} := \{ \phi_{\operatorname{tan}, h} \in {\mathcal H}^{\bullet}(Y, E|_{Y}) \mid \nabla \phi = \Gamma \nabla \Gamma \phi = 0 \},
\end{equation}
\noindent
where $\phi$ has the form (\ref{E:1.12}).
If $\phi$ satisfies $\nabla \phi = \Gamma \nabla \Gamma \phi = 0$, so is $\Gamma \phi$ and hence
\begin{equation} \label{E:2.47}
\Gamma^{Y} {\mathcal K} = \{ \phi_{\operatorname{nor}, h} \in {\mathcal H}^{\bullet}(Y, E|_{Y}) \mid \nabla \phi = \Gamma \nabla \Gamma \phi = 0 \},
\end{equation}
\noindent
where $\phi$ has the form (\ref{E:1.12}).
The second assertion in Lemma \ref{Lemma:1.1} shows that
${\mathcal K}$ is perpendicular to $\Gamma^{Y} {\mathcal K}$.
We then have the following decomposition (cf. Corollary 8.4 in [17], Lemma 2.4 in [14]).
\begin{equation} \label{E:1.144}
{\mathcal K} \oplus \Gamma^{Y} {\mathcal K} = {\mathcal H}^{\bullet}(Y, E|_{Y}),
\end{equation}
\noindent
which shows that
$( {\mathcal H}^{\bullet}(Y, E|_{Y}), \hspace{0.1 cm} \langle \hspace{0.1 cm}, \hspace{0.1 cm} \rangle_{Y}, \hspace{0.1 cm} - i \beta \Gamma^{Y} )$
is a symplectic vector space with Lagrangian subspaces ${\mathcal K}$ and $\Gamma^{Y} {\mathcal K}$.
We denote by
\begin{equation} \label{E:1.1555}
{\mathcal L}_{0} = \left( \begin{array}{clcr} {\mathcal K} \\ {\mathcal K} \end{array} \right), \qquad
{\mathcal L}_{1} = \left( \begin{array}{clcr} \Gamma^{Y} {\mathcal K} \\ \Gamma^{Y} {\mathcal K} \end{array} \right).
\end{equation}
\vspace{0.2 cm}
We next define the orthogonal projections
${\mathcal P}_{-, {\mathcal L}_{0}}$, ${\mathcal P}_{+, {\mathcal L}_{1}} :
\Omega^{\bullet}(Y, E|_{Y}) \oplus \Omega^{\bullet}(Y, E|_{Y}) \rightarrow \Omega^{\bullet}(Y, E|_{Y}) \oplus \Omega^{\bullet}(Y, E|_{Y})$ by
\begin{equation} \label{E:1.112}
\operatorname{Im} {\mathcal P}_{-, {\mathcal L}_{0}} \hspace{0.1 cm} = \hspace{0.1 cm}
\left( \begin{array}{clcr} \operatorname{Im} \nabla^{Y} \oplus {\mathcal K} \\ \operatorname{Im} \nabla^{Y} \oplus {\mathcal K} \end{array} \right), \qquad
\operatorname{Im} {\mathcal P}_{+, {\mathcal L}_{1}} \hspace{0.1 cm} = \hspace{0.1 cm}
\left( \begin{array}{clcr} \operatorname{Im} \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \oplus \Gamma^{Y} {\mathcal K} \\
\operatorname{Im} \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \oplus \Gamma^{Y} {\mathcal K} \end{array} \right) .
\end{equation}
\noindent
Then ${\mathcal P}_{-, {\mathcal L}_{0}}$, ${\mathcal P}_{+, {\mathcal L}_{1}}$ are pseudodifferential operators and give well-posed
boundary conditions for $\mathcal{B}$ and the refined analytic torsion.
We denote by $\mathcal{B}_{{\mathcal P}_{-, {\mathcal L}_{0}}}$ and $\mathcal{B}^{2}_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}$ the realizations of $\mathcal{B}$ and $\mathcal{B}^{2}_{q}$
with respect to ${\mathcal P}_{-, {\mathcal L}_{0}}$, {\it i.e.}
\begin{eqnarray}\label{E:2.17}
\operatorname{Dom} \left( \mathcal{B}_{{\mathcal P}_{-, {\mathcal L}_{0}}} \right) & = &
\left\{ \psi \in \Omega^{\bullet}(M, E) \mid {\mathcal P}_{-, {\mathcal L}_{0}} \left( \psi|_{Y} \right) = 0 \right\}, \nonumber\\
\operatorname{Dom} \left( \mathcal{B}^{2}_{q, {\mathcal P}_{-, {\mathcal L}_{0}}} \right) & = &
\left\{ \psi \in \Omega^{q}(M, E) \mid {\mathcal P}_{-, {\mathcal L}_{0}} \left( \psi|_{Y} \right) = 0, \hspace{0.1 cm}
{\mathcal P}_{-, {\mathcal L}_{0}} \left( (\mathcal{B} \psi)|_{Y} \right) = 0 \right\}.
\end{eqnarray}
\noindent
We define $\mathcal{B}_{{\mathcal P}_{+, {\mathcal L}_{1}}}$, $\mathcal{B}^{2}_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}$, $\mathcal{B}^{2}_{q, \operatorname{abs}}$, $\mathcal{B}^{2}_{q, \operatorname{rel}}$
and $\mathcal{B}_{\Pi_{>}}$, $\mathcal{B}_{\Pi_{<}}$ (see Section 3) in the similar way.
The following result is straightforward (Lemma 2.11 in [14]).
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:1.2}
$$
\operatorname{ker} \mathcal{B}^{2}_{q, {\mathcal P}_{-, {\mathcal L}_{0}}} = \ker \mathcal{B}^{2}_{q, \operatorname{rel}} = H^{q}(M, Y ; E), \qquad
\operatorname{ker} \mathcal{B}^{2}_{q, {\mathcal P}_{+, {\mathcal L}_{1}}} = \ker \mathcal{B}^{2}_{q, \operatorname{abs}} = H^{q}(M ; E).
$$
\end{lemma}
\vspace{0.2 cm}
We choose an Agmon angle $\theta$ by $- \frac{\pi}{2} < \theta < 0$.
For ${\frak D} = {\mathcal P}_{-, {\mathcal L}_{0}}$ or ${\mathcal P}_{+, {\mathcal L}_{1}}$ we define the zeta function $\zeta_{\mathcal{B}^{2}_{q, {\frak D}}}(s)$
and eta function $\eta_{\mathcal{B}_{\operatorname{even}, {\frak D}}}(s)$ by
\begin{eqnarray*}
\zeta_{\mathcal{B}^{2}_{q, {\frak D}}}(s) & = & \frac{1}{\Gamma(s)} \int_{0}^{\infty} t^{s-1}
\left( \operatorname{Tr} e^{- t \mathcal{B}^{2}_{q, {\frak D}}} - \operatorname{dim} \operatorname{ker} \mathcal{B}^{2}_{q, {\frak D}} \right) dt \\
\eta_{\mathcal{B}_{\operatorname{even}, {\frak D}}}(s) & = & \frac{1}{\Gamma(\frac{s+1}{2})} \int_{0}^{\infty} t^{\frac{s-1}{2}}
\operatorname{Tr} \left( \mathcal{B} e^{- t \mathcal{B}^{2}_{\operatorname{even}, {\frak D}}} \right) dt.
\end{eqnarray*}
\noindent
It was shown in [14] that $\zeta_{\mathcal{B}^{2}_{q, {\frak D}}}(s)$ and $\eta_{\mathcal{B}_{\operatorname{even}, {\frak D}}}(s)$ have regular values at $s=0$.
We define the zeta-determinant and eta-invariant by
\begin{eqnarray}
\log \operatorname{Det}_{2\theta} \mathcal{B}^{2}_{q, {\frak D}} & := & - \zeta_{\mathcal{B}^{2}_{q, {\frak D}}}^{\prime}(0), \label{E:1.15} \\
\eta(\mathcal{B}_{\operatorname{even}, {\frak D}}) & := & \frac{1}{2} \left( \eta_{\mathcal{B}_{\operatorname{even}, {\frak D}}}(0) + \operatorname{dim} \operatorname{ker} \mathcal{B}_{\operatorname{even}, {\frak D}} \right). \label{E:1.16}
\end{eqnarray}
\vspace{0.2 cm}
\noindent
We denote
\begin{eqnarray} \label{E:1.188}
& & \Omega^{q}_{-}(M, E) = \operatorname{Im} \nabla \cap \Omega^{q}(M, E), \qquad
\Omega^{q}_{+}(M, E) = \operatorname{Im} \Gamma \nabla \Gamma \cap \Omega^{q}(M, E), \nonumber \\
& & \Omega^{\operatorname{even}}_{\pm}(M, E) = \sum_{q = \operatorname{even}} \Omega^{q}_{\pm}(M, E),
\end{eqnarray}
\noindent
and denote by $\mathcal{B}_{\operatorname{even}}^{\pm}$ the restriction of $\hspace{0.1 cm} \mathcal{B}_{\operatorname{even}} \hspace{0.1 cm}$ to $\hspace{0.1 cm} \Omega^{\operatorname{even}}_{\pm}(M, E)$.
The graded zeta-determinant $\operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\frak D}})$ of $\mathcal{B}_{\operatorname{even}}$ with respect to the boundary condition ${\frak D}$
is defined by
$$
\operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\frak D}}) = \frac{\operatorname{Det}_{\theta} \mathcal{B}^{+}_{\operatorname{even}, {\frak D}}}
{\operatorname{Det}_{\theta} \left( - \mathcal{B}^{-}_{\operatorname{even}, {\frak D}} \right)} .
$$
\vspace{0.2 cm}
We next define the projections ${\widetilde {\mathcal P}}_{0}$,
${\widetilde {\mathcal P}}_{1} : \Omega^{\bullet}(Y, E|_{Y}) \oplus \Omega^{\bullet}(Y, E|_{Y}) \rightarrow
\Omega^{\bullet}(Y, E|_{Y}) \oplus \Omega^{\bullet}(Y, E|_{Y})$ as follows.
For $\phi \in \Omega^{q}(M, E)$
\vspace{0.2 cm}
$$
{\widetilde {\mathcal P}}_{0} (\phi|_{Y}) = \begin{cases} {\mathcal P}_{-, {\mathcal L}_{0}} (\phi|_{Y}) \quad
\text{if} \quad q \quad \text{is} \quad \text{even} \\
{\mathcal P}_{+, {\mathcal L}_{1}} (\phi|_{Y}) \quad \text{if} \quad q \quad \text{is} \quad \text{odd} ,
\end{cases}
\qquad
{\widetilde {\mathcal P}}_{1} (\phi|_{Y}) = \begin{cases} {\mathcal P}_{+, {\mathcal L}_{1}} (\phi|_{Y}) \quad
\text{if} \quad q \quad \text{is} \quad \text{even} \\
{\mathcal P}_{-, {\mathcal L}_{0}} (\phi|_{Y}) \quad \text{if} \quad q \quad \text{is} \quad \text{odd} .
\end{cases}
$$
\vspace{0.2 cm}
\noindent
We denote by
\begin{equation} \label{E:1.199}
l_{q} := \operatorname{dim} \operatorname{ker} \mathcal{B}_{Y, q}^{2}, \qquad l_{q}^{+} := \operatorname{dim} {\mathcal K} \cap \operatorname{ker} \mathcal{B}_{Y, q}^{2}, \qquad \text{and} \qquad
l_{q}^{-} := \operatorname{dim} \Gamma^{Y} {\mathcal K} \cap \operatorname{ker} \mathcal{B}_{Y, q}^{2},
\end{equation}
\vspace{0.2 cm}
\noindent
so that $l_{q} = l_{q}^{+} + l_{q}^{-}$ and $l_{q}^{-} = l_{m-1-q}^{+}$.
Simple computation shows that $\log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}})$ and
$\log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\mathcal P}_{+, {\mathcal L}_{1}}})$ are described as follows ([14]).
\begin{eqnarray}
(1) \label{E:1.17}
\hspace{0.2 cm} \log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}}) & = &
\frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2\theta} \mathcal{B}^{2}_{q, {\widetilde {\mathcal P}_{0}}}
- i \pi \hspace{0.1 cm} \eta(\mathcal{B}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}}) \nonumber \\
& + & \frac{\pi i }{2} \left( \frac{1}{4} \sum_{q=0}^{m-1} \zeta_{\mathcal{B}_{Y, q}^{2}}(0) + \sum_{q=0}^{r-2}(r-1-q) (l_{q}^{+} - l_{q}^{-}) \right). \\
(2) \label{E:1.18}
\hspace{0.2 cm} \log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\mathcal P}_{+, {\mathcal L}_{1}}}) & = &
\frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2\theta} \mathcal{B}^{2}_{q, {\widetilde {\mathcal P}_{1}}}
- i \pi \hspace{0.1 cm} \eta(\mathcal{B}_{\operatorname{even}, {\mathcal P}_{+, {\mathcal L}_{1}}}) \nonumber \\
& - & \frac{\pi i }{2} \left( \frac{1}{4} \sum_{q=0}^{m-1} \zeta_{\mathcal{B}_{Y, q}^{2}}(0) + \sum_{q=0}^{r-2}(r-1-q) (l_{q}^{+} - l_{q}^{-}) \right).
\end{eqnarray}
\vspace{0.2 cm}
To define the refined analytic torsion we introduce the trivial connection $\nabla^{\operatorname{trivial}}$ acting on the trivial bundle $M \times {\Bbb C}$
and define the trivial odd signature operator $\mathcal{B}^{\operatorname{trivial}}_{\operatorname{even}} : \Omega^{\operatorname{even}}(M, {\Bbb C}) \rightarrow \Omega^{\operatorname{even}}(M, {\Bbb C})$ in the same way as
(\ref{E:1.2}). The eta invariant $\eta ( \mathcal{B}^{\operatorname{trivial}}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}} } )$
associated to $\mathcal{B}^{\operatorname{trivial}}_{\operatorname{even}}$
and subject to the boundary condition ${\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}}$ is defined in the same way
as in (\ref{E:1.16}) by simply replacing $\mathcal{B}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}} }$
by $\mathcal{B}^{\operatorname{trivial}}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}} }$.
When $\nabla$ is acyclic in the de Rham complex,
the refined analytic torsion subject to the boundary condition ${\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}}$ is defined
by
\begin{eqnarray}
\log T_{{\mathcal P}_{-, {\mathcal L}_{0}}} (g^{M}, \nabla) & = & \log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}}) +
\frac{\pi i}{2} (\operatorname{rank} E) \eta_{ \mathcal{B}^{\operatorname{trivial}}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}} } (0) \\
\log T_{{\mathcal P}_{+, {\mathcal L}_{1}}} (g^{M}, \nabla) & = & \log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}_{\operatorname{even}, {\mathcal P}_{+, {\mathcal L}_{1}}}) +
\frac{\pi i}{2} (\operatorname{rank} E) \eta_{ \mathcal{B}^{\operatorname{trivial}}_{\operatorname{even}, {\mathcal P}_{+, {\mathcal L}_{1}}} } (0)
\end{eqnarray}
\vspace{0.2 cm}
\noindent
The refined analytic torsion on a closed manifold is defined similarly.
In this paper we are going to discuss the gluing formula
of the refined analytic torsion with respect to the boundary conditions ${\mathcal P}_{-, {\mathcal L}_{0}}$ and ${\mathcal P}_{+, {\mathcal L}_{1}}$.
For this purpose in the next two sections we are going to compare the Ray-Singer analytic torsion and eta invariant subject to the boundary
condition ${\mathcal P}_{-, {\mathcal L}_{0}}$ (or ${\mathcal P}_{+, {\mathcal L}_{1}}$) with those subject to the relative and APS boundary conditions,
respectively.
\vspace{0.3 cm}
\section{Comparison of the Ray-Singer analytic torsions}
\vspace{0.2 cm}
In this section we are going to compare the Ray-Singer analytic torsion subject to the boundary condition ${\mathcal P}_{-, {\mathcal L}_{0}}$
with the Ray-Singer analytic torsion subject to the relative boundary condition. For this purpose we are going to use the BFK-gluing formula and
the method of the adiabatic limit for stretching the cylinder part.
We recall that $(M, g^{M})$ is a compact oriented Riemannian manifold with boundary $Y$
with a collar neighbrhood $N = [0, 1) \times Y$ and $g^{M}$ is assumed to be a product metric on $N$.
We denote by $M_{1, 1} = [0, 1] \times Y$ and $M_{2} = M - N$.
To use the adiabatic limit we stretch the cylinder part $M_{1, 1}$ to the cylinder of length $r$ and
denote $M_{1, r} = [0, r] \times Y$ with the product metric and
$$
M_{r} = M_{1, r} \cup_{Y_{r}} M_{2} \quad \text{with} \hspace{0.2 cm} Y_{r} = \{ r \} \times Y.
$$
\noindent
Then we can extend the bundle $E$ and
the odd signature operator $\mathcal{B}$ on $M$ to $M_{r}$ in the natural way and we denote these extensions
by $E_{r}$ and $\mathcal{B}(r)$ ($\mathcal{B} = \mathcal{B}(1)$). We denote the restriction of $\mathcal{B}(r)$ to $M_{1, r}$, $M_{2}$ by $\mathcal{B}_{M_{1, r}}$, $\mathcal{B}_{M_{2}}$.
It is well known (cf. [16], [2]) that the Dirichlet boundary value problem for $\mathcal{B}_{q}^{2}$ on $M_{2}$ has a unique
solution, {\it i.e.} for $f + du \wedge g \in \Omega^{q}(M_{2}, E|_{M_{2}})|_{Y_{r}}$,
there exists a unique $\psi \in \Omega^{q}(M_{2}, E|_{M_{2}})$
such that
$$
\mathcal{B}_{q}^{2}\psi = 0, \qquad \psi|_{Y_{r}} = f + du \wedge g.
$$
\noindent
Let ${\frak D}$ be one of the following boundary conditions :
${\mathcal P}_{-, {\mathcal L}_{0}}$, ${\mathcal P}_{+, {\mathcal L}_{1}}$,
the absolute boundary condition, the relative boundary condition or the Dirichlet boundary condition.
We define the Neumann jump operators
$$
Q_{q, 1, {\frak D}}(r), \hspace{0.1 cm} Q_{q, 2} : \Omega^{q}(Y_{r}, E|_{Y_{r}}) \oplus \Omega^{q-1}(Y_{r}, E|_{Y_{r}}) \rightarrow
\Omega^{q}(Y_{r}, E|_{Y_{r}}) \oplus \Omega^{q-1}(Y_{r}, E|_{Y_{r}})
$$
\vspace{0.2 cm}
\noindent
as follows.
For $f + du \wedge g \in \Omega^{q}(Y_{r}, E|_{Y_{r}}) \oplus \Omega^{q-1}(Y_{r}, E|_{Y_{r}})$, we choose
$\phi \in \Omega^{q}(M_{1, r}, E|_{M_{1, r}})$ and $\psi \in \Omega^{q}(M_{2}, E|_{M_{2}})$ such that
\begin{equation} \label{E:2.1}
\mathcal{B}_{q, M_{1, r}}^{2}\phi = 0, \qquad \mathcal{B}_{q, M_{2}}^{2} \psi = 0, \qquad \phi|_{Y_{r}} = \psi|_{Y_{r}} = f + du \wedge g,
\qquad {\frak D}(\phi|_{Y_{0}}) = 0.
\end{equation}
\noindent
Then we define
$$
Q_{q, 1, {\frak D}}(r) (f) \hspace{0.1 cm} = \hspace{0.1 cm} (\nabla_{\partial_{u}} \phi)|_{Y_{r}}, \qquad
Q_{q, 2}(f) \hspace{0.1 cm} = \hspace{0.1 cm} - \hspace{0.1 cm} (\nabla_{\partial_{u}} \psi)|_{Y_{r}},
$$
\noindent
where $\partial u$ is the inward unit normal vector field on $N \subset M$.
We next define the Dirichlet-to-Neumann operator $R_{q, {\frak D}}(r)$ as follows.
$$
R_{q, {\frak D}}(r) : \Omega^{q}(Y_{r}, E|_{Y_{r}}) \oplus \Omega^{q-1}(Y_{r}, E|_{Y_{r}}) \rightarrow
\Omega^{q}(Y_{r}, E|_{Y_{r}}) \oplus \Omega^{q-1}(Y_{r}, E|_{Y_{r}})
$$
\begin{equation} \label{E:2.2}
R_{q, {\frak D}} (r) = Q_{q, 1, {\frak D}}(r) + Q_{q, 2}.
\end{equation}
\vspace{0.1 cm}
\noindent
The following lemma is well known (cf. [18]).
\begin{lemma} \label{Lemma:2.1}
(1) $R_{q, \frak D} (r)$ is a non-negative elliptic pseudodifferential operator of order $1$ and has the form of
\begin{equation} \label{E:2.3}
R_{q, \frak D}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
\left( \begin{array}{clcr} 2 \sqrt{\mathcal{B}_{Y, q}^{2}} & 0 \\ 0 & 2 \sqrt{\mathcal{B}_{Y, q-1}^{2}} \end{array} \right) \hspace{0.1 cm} +
\hspace{0.2 cm} \text{ a smoothing operator}.
\end{equation}
\noindent
(2) $\ker R_{q, \frak D} = \{ \phi|_{Y_{r}} \mid \phi \in \operatorname{ker} \mathcal{B}_{q, {\frak D}}^{2} \}$.
\end{lemma}
\vspace{0.2 cm}
We denote by $\mathcal{B}^{2}_{q, M_{1, r}, {\frak D}, D}$ ($\mathcal{B}^{2}_{q, M_{2}, D}$)
the restriction of $\mathcal{B}^{2}_{q}(r)$ to $M_{1, r}$
($M_{2}$) subject to the boundary condition ${\frak D}$ on $Y_{0}$ and the Dirichlet boundary condition on $Y_{r}$
(the Dirichlet boundary condition on $Y_{r}$).
We denote by $\mathcal{B}^{2}_{q, {\frak D}}(r)$ the operator $\mathcal{B}_{q}^{2}(r)$ on $M_{r}$ subject to the boundary condition ${\frak D}$
on $Y_{0}$. Then Lemma \ref{Lemma:1.2} shows that $\operatorname{dim} \operatorname{ker} \mathcal{B}^{2}_{q, {\frak D}}(r)$ is a topological invariant.
Let $\operatorname{dim} \operatorname{ker} \mathcal{B}^{2}_{q, {\frak D}}(r) = k$ and
$\{\varphi_{1}, \cdots, \varphi_{k} \}$ be an orthonormal basis of $\operatorname{ker} \mathcal{B}^{2}_{q, {\frak D}}(r)$.
We define a positive definite $k \times k$ Hermitian matrix $A_{q, {\frak D}}(r)$ by
$$
A_{q, {\frak D}}(r) = (a_{ij}), \qquad a_{ij} = \langle \varphi_{i}|_{Y_{0}}, \varphi_{j}|_{Y_{0}} \rangle_{Y_{0}}.
$$
Then the BFK-gluing formula ([8], [18], [19]) is described as follows.
Setting $l_{q} = \operatorname{dim} \operatorname{ker} \mathcal{B}_{Y, q}^{2}$, we have
\begin{eqnarray}
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\frak D}}(r)
& = & \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{1, r}, {\frak D}, D} + \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{2}, D}
+ \log \operatorname{Det}_{2 \theta} R_{q, {\frak D}}(r) \nonumber \\
& & - \log 2 \cdot (\zeta_{\mathcal{B}^{2}_{Y, q}}(0) + \zeta_{\mathcal{B}^{2}_{Y, q-1}}(0) + l_{q} + l_{q-1}) - \log \operatorname{det} A_{q, {\frak D}}(r).
\end{eqnarray}
\vspace{0.2 cm}
\noindent
The above equality can be rewritten as follows.
\begin{corollary} \label{Corollary:2.3}
\begin{eqnarray*}
(1) & & \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{-, {\mathcal L}_{0}}, D} + \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{2}, D} +
\log \operatorname{Det}_{2 \theta} R_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}(r) \\
& & \hspace{4.5 cm} - \log 2 \cdot (\zeta_{\mathcal{B}^{2}_{Y, q}}(0) + \zeta_{\mathcal{B}^{2}_{Y, q-1}}(0) + l_{q} + l_{q-1}) - \log \operatorname{det} A_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}(r) \\
(2) & & \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{+, {\mathcal L}_{1}}, D} + \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{2}, D} +
\log \operatorname{Det}_{2 \theta} R_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}(r) \\
& & \hspace{4.5 cm} - \log 2 \cdot (\zeta_{\mathcal{B}^{2}_{Y, q}}(0) + \zeta_{\mathcal{B}^{2}_{Y, q-1}}(0) + l_{q} + l_{q-1}) - \log \operatorname{det} A_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}(r) \\
(3) & & \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{rel}}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} + \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{2}, D} +
\log \operatorname{Det}_{2 \theta} R_{q, \operatorname{rel}}(r) \\
& & \hspace{4.5 cm} - \log 2 \cdot (\zeta_{\mathcal{B}^{2}_{Y, q}}(0) + \zeta_{\mathcal{B}^{2}_{Y, q-1}}(0) + l_{q} + l_{q-1}) - \log \operatorname{det} A_{q, \operatorname{rel}}(r) \\
(4) & & \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{abs}}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{abs}, D} + \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, M_{2}, D} +
\log \operatorname{Det}_{2 \theta} R_{q, \operatorname{abs}}(r) \\
& & \hspace{4.5 cm} - \log 2 \cdot (\zeta_{\mathcal{B}^{2}_{Y, q}}(0) + \zeta_{\mathcal{B}^{2}_{Y, q-1}}(0) + l_{q} + l_{q-1}) - \log \operatorname{det} A_{q, \operatorname{abs}}(r)
\end{eqnarray*}
\end{corollary}
\vspace{0.2 cm}
\noindent
{\it Remark} : The BFK-gluing formula was proved originally on a closed manifold in [8]. But it can be extended
to a compact manifold with boundary with only minor modification when a cutting hypersurface does not intersect the boundary.
\vspace{0.2 cm}
We define $\Omega^{q}_{\pm}(Y, E|_{Y})$ similarly to (\ref {E:1.188})
and denote $\mathcal{B}_{Y, q}^{2, \pm} := \mathcal{B}^{2}_{Y, q}|_{\Omega^{q}_{\pm}(Y, E|_{Y})}$.
Simple computation leads to the following results.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:2.3}
The spectra of $\hspace{0.1 cm} \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{-, {\mathcal L}_{0}}, D} \hspace{0.1 cm}$,
$\hspace{0.1 cm} \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{+, {\mathcal L}_{1}}, D} \hspace{0.1 cm}$,
$\hspace{0.1 cm} \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} \hspace{0.1 cm}$ and $\hspace{0.1 cm} \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{abs}, D} \hspace{0.1 cm}$
are given as follows. Let $k = 1, 2, 3, \cdots$.
\begin{eqnarray*}
& (1) & Spec \left( \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{-, {\mathcal L}_{0}}, D} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\left\{ \lambda_{q-1, j} + (\frac{k \pi}{r})^{2} \right\} \hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q-2, j} + (\frac{k \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q, j} + (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\} \\
& & \hspace{1.5 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q-1, j} + (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{k \pi}{r})^{2} \right\}_{l_{q}^{+} + l_{q-1}^{+}}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}_{l_{q}^{-} + l_{q-1}^{-}}, \\
& (2) & Spec \left( \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{+, {\mathcal L}_{1}}, D} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\left\{ \lambda_{q-1, j} + (\frac{(k- \frac{1}{2}) \pi}{r})^{2} \right\} \hspace{0.1 cm} \cup \hspace{0.1 cm}
\left\{ \lambda_{q-2, j} + (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q, j} + (\frac{k \pi}{r})^{2} \right\} \\
& & \hspace{1.5 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q-1, j} + (\frac{k \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{k \pi}{r})^{2} \right\}_{l_{q}^{-} + l_{q-1}^{-}}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}_{l_{q}^{+} + l_{q-1}^{+}}, \\
& (3) & Spec \left( \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\left\{ \lambda_{q-1, j} + (\frac{k \pi}{r})^{2} \right\} \hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q, j} + (\frac{k \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q-2, j} + (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\} \\
& & \hspace{1.5 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q-1, j} + (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{k \pi}{r})^{2} \right\}_{l_{q}}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}_{l_{q-1}}, \\
& (4) & Spec \left( \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{abs}, D} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\left\{ \lambda_{q-1, j} + (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\} \hspace{0.1 cm} \cup \hspace{0.1 cm}
\left\{ \lambda_{q, j} + (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q-2, j} + (\frac{k \pi}{r})^{2} \right\} \\
& & \hspace{1.5 cm} \cup \hspace{0.1 cm} \left\{ \lambda_{q-1, j} + (\frac{k \pi}{r})^{2} \right\}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{k \pi}{r})^{2} \right\}_{l_{q-1}}
\hspace{0.1 cm} \cup \hspace{0.1 cm} \left\{ (\frac{(k - \frac{1}{2}) \pi}{r})^{2} \right\}_{l_{q}},
\end{eqnarray*}
where each $\lambda_{q, j}$ runs on $\operatorname{Spec} \left( \mathcal{B}_{Y, q}^{2, +} \right)$ and
$\left\{ (\frac{k \pi}{r})^{2} \right\}_{l_{q}^{+}}$ means that the multiplicity of each $(\frac{k \pi}{r})^{2}$ is $l_{q}^{+}$.
\end{lemma}
For each $q$ we define
\begin{eqnarray*}
\zeta_{\Delta_{q, N}} (s) & = & \sum_{\lambda_{q, j} \in \operatorname{Spec}(\mathcal{B}_{Y, q}^{2, +})} \sum_{k=1}^{\infty}
\left( \lambda_{q, j} + \left( \frac{(k - \frac{1}{2}) \pi}{r} \right)^{2} \right)^{-s}, \\
\zeta_{\Delta_{q, D}} (s) & = & \sum_{\lambda_{q, j} \in \operatorname{Spec}(\mathcal{B}_{Y, q}^{2, +})} \sum_{k=1}^{\infty}
\left( \lambda_{q, j} + \left( \frac{k \pi}{r} \right)^{2} \right)^{-s}.
\end{eqnarray*}
\noindent
The following result is well known (cf. [22]).
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:2.4}
We put $\xi_{Y, q}(s) = \frac{\Gamma(s - \frac{1}{2})}{2 \sqrt{\pi} \Gamma(s)} \zeta_{\mathcal{B}_{Y, q}^{2, +}}(s - \frac{1}{2})$. Then :
\begin{eqnarray*}
& (1) & - \zeta_{\Delta_{q, N}}^{\prime} (0) \hspace{0.1 cm} = \hspace{0.1 cm} - r \xi_{Y, q}^{\prime}(0) + \sum_{\lambda_{q, j} \in \operatorname{Spec}(\mathcal{B}_{Y, q}^{2, +})}
\log ( 1 + e^{- 2 r \sqrt{\lambda_{q, j}}}), \\
& (2) & - \zeta_{\Delta_{q, D}}^{\prime} (0) \hspace{0.1 cm} = \hspace{0.1 cm} - r \xi_{Y, q}^{\prime}(0) - \frac{1}{2} \log \operatorname{Det} (\mathcal{B}_{Y, q}^{2, +})
\hspace{0.1 cm} + \sum_{\lambda_{q, j} \in \operatorname{Spec}(\mathcal{B}_{Y, q}^{2, +})}\log ( 1 - e^{- 2 r \sqrt{\lambda_{q, j}}}).
\end{eqnarray*}
\end{lemma}
\begin{proof}
The computation of
$- \zeta_{\Delta_{q, D}}^{\prime} (0)$ was done in Proposition 5.1 of [22].
Using the Poisson summation formula, we have the following identity
$$
\sum_{k=1}^{\infty} e^{- \pi^{2} (k - \frac{1}{2})^{2} t} \hspace{0.1 cm} = \hspace{0.1 cm}
\frac{1}{\sqrt{\pi t}} \hspace{0.1 cm} \left( \frac{1}{2} \hspace{0.1 cm} + \hspace{0.1 cm} 2 \sum_{k=1}^{\infty} e^{-\frac{4 k^{2}}{t}}
\hspace{0.1 cm} - \hspace{0.1 cm} \sum_{k=1}^{\infty} e^{- \frac{k^{2}}{t}} \right),
$$
from which we can compute $- \zeta_{\Delta_{q, N}}^{\prime} (0)$.
\end{proof}
\vspace{0.2 cm}
\begin{corollary} \label{Corollary:2.5}
Putting $\hspace{0.1 cm} {\mathcal C}_{q}^{+}(r) = \prod_{\lambda_{q, j} \in \operatorname{Spec}(\mathcal{B}_{Y, q}^{2, +})}
\left( 1 + \frac{2 e^{- r \sqrt{\lambda_{q, j}}}}{e^{r \sqrt{\lambda_{q, j}}} - e^{- r \sqrt{\lambda_{q, j}}}} \right)$, we have
\begin{eqnarray*}
\left( - \zeta_{\Delta_{q, N}}^{\prime} (0) \right) - \left( - \zeta_{\Delta_{q, D}}^{\prime} (0) \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\frac{1}{2} \log \operatorname{Det} \mathcal{B}_{Y, q}^{2, +} \hspace{0.1 cm} + \log {\mathcal C}_{q}^{+}(r).
\end{eqnarray*}
\end{corollary}
\vspace{0.2 cm}
\noindent
If we denote the Riemann zeta function by $\zeta_{R}(s)$, it is well known that
$\zeta_{R}(0) = - \frac{1}{2}$ and $\zeta_{R}^{\prime}(0) = - \frac{1}{2} \log 2 \pi$, which leads to the following result.
\begin{lemma} \label{Lemma:2.6}
Setting
$ \hspace{0.1 cm}
\zeta_{1}(s) = \sum_{k=1}^{\infty} \left( \frac{k \pi}{r} \right)^{-2s}$ and
$\hspace{0.1 cm} \zeta_{2}(s) = \sum_{k=1}^{\infty} \left( \frac{(k - \frac{1}{2}) \pi}{r} \right)^{-2s}$,
we have
$ \hspace{0.1 cm} \zeta_{1}^{\prime}(0) = - \log 2r \hspace{0.1 cm}$ and $\hspace{0.1 cm} \zeta_{2}^{\prime}(0) = - \log 2$.
\end{lemma}
\noindent
Lemma \ref{Lemma:2.3} together with Corollary \ref{Corollary:2.5} and Lemma \ref{Lemma:2.6} yields the following result.
\begin{lemma} \label{Lemma:2.7}
\begin{eqnarray*}
(1) \hspace{0.2 cm} \log \operatorname{Det} \left( \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{-, {\mathcal L}_{0}}, D} \right) -
\log \operatorname{Det} \left( \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} \right) & = & \frac{1}{2} \left( \log \operatorname{Det} \mathcal{B}_{Y, q}^{2, +} - \log \operatorname{Det} \mathcal{B}_{Y, q-2}^{2, +} \right) \\
& + & \left( \log {\mathcal C}_{q}^{+}(r) - \log {\mathcal C}_{q-2}^{+}(r) \right) + \left( l_{q-1}^{+} - l_{q}^{-} \right) \log r \\
(2) \hspace{0.2 cm} \log \operatorname{Det} \left( \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{+, {\mathcal L}_{1}}, D} \right) -
\log \operatorname{Det} \left( \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} \right) & = & \left( l_{q-1}^{-} - l_{q}^{+} \right) \log r
\end{eqnarray*}
\begin{eqnarray*}
\noindent
(3) \hspace{0.2 cm} \sum_{q=0}^{m} (-1)^{q+1} q \left( \log \operatorname{Det} \mathcal{B}^{2}_{q, M_{1, r}, {\widetilde {\mathcal P}}_{0}, D}
- \log \operatorname{Det} \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} \right) & = & \sum_{q=\operatorname{even}} \log \operatorname{Det} \mathcal{B}_{Y, q}^{2, +} \hspace{0.1 cm} + \hspace{0.1 cm}
2 \sum_{q=\operatorname{even}} \log {\mathcal C}_{q}^{+}(r) \\
+ \hspace{0.1 cm} \left( \sum_{q = \operatorname{even}} (2q+1) l_{q}^{-} - \sum_{q = \operatorname{odd}} (2q+1) l_{q}^{+} \right) \log r & &
\end{eqnarray*}
\begin{eqnarray*}
\noindent
(4) \hspace{0.2 cm} \sum_{q=0}^{m} (-1)^{q+1} q \left( \log \operatorname{Det} \mathcal{B}^{2}_{q, M_{1, r}, {\widetilde {\mathcal P}}_{1}, D}
- \log \operatorname{Det} \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} \right) & = & - \sum_{q=\operatorname{odd}} \log \operatorname{Det} \mathcal{B}_{Y, q}^{2, +} \hspace{0.1 cm} - \hspace{0.1 cm}
2 \sum_{q=\operatorname{odd}} \log {\mathcal C}_{q}^{+}(r) \\
+ \hspace{0.1 cm} \left( \sum_{q = \operatorname{even}} (2q+1) l_{q}^{+} - \sum_{q = \operatorname{odd}} (2q+1) l_{q}^{-} \right) \log r & &
\end{eqnarray*}
\begin{eqnarray*}
\noindent
(5) \hspace{0.2 cm} \sum_{q=0}^{m} (-1)^{q+1} q \left( \log \operatorname{Det} \mathcal{B}^{2}_{q, M_{1, r}, {\mathcal P}_{-, {\mathcal L}_{0}}, D}
- \log \operatorname{Det} \mathcal{B}^{2}_{q, M_{1, r}, \operatorname{rel}, D} \right) & = & \frac{m}{2} \hspace{0.1 cm} \chi(Y, E|_{Y}) \hspace{0.1 cm} \log r,
\end{eqnarray*}
where $\hspace{0.1 cm} \chi(Y, E|_{Y}) := \sum_{q=0}^{m-1} (-1)^{q} \cdot l_{q} \hspace{0.1 cm}$ the Euler characteristic of $Y$
with respect to $H^{\bullet}(Y, E|_{Y})$.
\end{lemma}
\vspace{0.2 cm}
We finally discuss the Dirichlet-to-Neumann operator $R_{q, {\frak D}}(r)$ defined by
$R_{q, {\frak D}}(r) = Q_{q, 1, {\frak D}}(r) + Q_{q, 2}$, where ${\frak D}$ is one of
${\mathcal P}_{-, {\mathcal L}_{0}}$, ${\mathcal P}_{+, {\mathcal L}_{1}}$, the absolute or the relative boundary condition.
The following lemma is straightforward.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:2.8}
$R_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}(r)$, $R_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}(r)$ and $R_{q, \operatorname{rel}}(r)$ are described as follows.
\begin{eqnarray*}
R_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}(r) & = & Q_{q, 2} \hspace{0.1 cm} + \hspace{0.1 cm}
\left( \begin{array}{clcr} \sqrt{\mathcal{B}_{Y, q}^{2}} & 0 \\ 0 & \sqrt{\mathcal{B}_{Y, q-1}^{2}} \end{array} \right) \hspace{0.1 cm} + \hspace{0.1 cm}
\begin{cases}
\frac{ 2 \sqrt{\mathcal{B}_{Y}^{2}} e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
{e^{\sqrt{\mathcal{B}_{Y}^{2}} r} - e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
\hspace{0.2 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{-, {\mathcal L}_{0}} \cap (\operatorname{ker} \mathcal{B}_{Y}^{2})^{\perp}\\
\frac{1}{r} \hspace{2.7 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{-, {\mathcal L}_{0}} \cap \operatorname{ker} \mathcal{B}_{Y}^{2}\\
- \frac{ 2 \sqrt{\mathcal{B}_{Y}^{2}} e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
{e^{\sqrt{\mathcal{B}_{Y}^{2}} r} + e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
\hspace{0.2 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{+, {\mathcal L}_{1}} \cap (\operatorname{ker} \mathcal{B}_{Y}^{2})^{\perp}\\
0 \hspace{2.8 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{+, {\mathcal L}_{1}} \cap \operatorname{ker} \mathcal{B}_{Y}^{2}
\end{cases} \\
R_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}(r) & = & Q_{q, 2} \hspace{0.1 cm} + \hspace{0.1 cm}
\left( \begin{array}{clcr} \sqrt{\mathcal{B}_{Y, q}^{2}} & 0 \\ 0 & \sqrt{\mathcal{B}_{Y, q-1}^{2}} \end{array} \right) \hspace{0.1 cm} + \hspace{0.1 cm}
\begin{cases}
- \frac{ 2 \sqrt{\mathcal{B}_{Y}^{2}} e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
{e^{\sqrt{\mathcal{B}_{Y}^{2}} r} + e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
\hspace{0.2 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{-, {\mathcal L}_{0}} \cap (\operatorname{ker} \mathcal{B}_{Y}^{2})^{\perp}\\
0 \hspace{2.7 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{-, {\mathcal L}_{0}} \cap \operatorname{ker} \mathcal{B}_{Y}^{2}\\
\frac{ 2 \sqrt{\mathcal{B}_{Y}^{2}} e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
{e^{\sqrt{\mathcal{B}_{Y}^{2}} r} - e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
\hspace{0.2 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{+, {\mathcal L}_{1}} \cap (\operatorname{ker} \mathcal{B}_{Y}^{2})^{\perp}\\
\frac{1}{r} \hspace{2.8 cm} \text{on} \hspace{0.2 cm}
\operatorname{Im} {\mathcal P}_{+, {\mathcal L}_{1}} \cap \operatorname{ker} \mathcal{B}_{Y}^{2}
\end{cases} \\
R_{q, \operatorname{rel}}(r) & = & Q_{q, 2} \hspace{0.1 cm} + \hspace{0.1 cm}
\left( \begin{array}{clcr} \sqrt{\mathcal{B}_{Y, q}^{2}} & 0 \\ 0 & \sqrt{\mathcal{B}_{Y, q-1}^{2}} \end{array} \right) \hspace{0.1 cm} + \hspace{0.1 cm}
\begin{cases}
\frac{ 2 \sqrt{\mathcal{B}_{Y}^{2}} e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}{e^{\sqrt{\mathcal{B}_{Y}^{2}} r} - e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
\hspace{0.2 cm} \text{on} \hspace{0.2 cm} \Omega^{q}_{-}(Y, E|_{Y}) \oplus \Omega^{q}_{+}(Y, E|_{Y}) \\
\frac{1}{r} \hspace{2.7 cm} \text{on} \hspace{0.2 cm} \operatorname{ker} \mathcal{B}_{Y}^{2} \cap \Omega^{q}(Y, E|_{Y})\\
- \frac{ 2 \sqrt{\mathcal{B}_{Y}^{2}} e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}{e^{\sqrt{\mathcal{B}_{Y}^{2}} r} + e^{- \sqrt{\mathcal{B}_{Y}^{2}} r}}
\hspace{0.2 cm} \text{on} \hspace{0.2 cm} \Omega^{q-1}_{-}(Y, E|_{Y}) \oplus \Omega^{q-1}_{+}(Y, E|_{Y}) \\
0 \hspace{2.8 cm} \text{on} \hspace{0.2 cm} \operatorname{ker} \mathcal{B}_{Y}^{2} \cap \Omega^{q-1}(Y, E|_{Y})
\end{cases}
\end{eqnarray*}
\end{lemma}
\vspace{0.2 cm}
We next discuss
$\lim_{r \rightarrow \infty} \left(
\log \operatorname{Det} R_{q, {\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}}}(r) - \log \operatorname{Det} R_{q, \operatorname{rel}}(r) \right)$
when $H^{q}(M, Y ; E) = \{ 0 \}$ for each $0 \leq q \leq m$.
The Poincar\'e duality and long exact sequence imply that $ H^{q}(M ; E) = H^{q}(Y ; E|_{Y}) = 0$ for each $0 \leq q \leq m$.
Then Lemma \ref{Lemma:1.2} and Lemma \ref{Lemma:2.1} show that $R_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}(r)$,
$R_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}(r)$ and $R_{q, \operatorname{rel}}(r)$ are invertible operators and
\begin{eqnarray*}
\lim_{r \rightarrow \infty} R_{q, {\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}}}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
\lim_{r \rightarrow \infty} R_{q, \operatorname{rel}}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
Q_{q, 2} \hspace{0.1 cm} + \hspace{0.1 cm}
\left( \begin{array}{clcr} \sqrt{\mathcal{B}_{Y, q}^{2}} & 0 \\ 0 & \sqrt{\mathcal{B}_{Y, q-1}^{2}} \end{array} \right)
\hspace{0.1 cm} = \hspace{0.1 cm} Q_{q, 2} \hspace{0.1 cm} + \hspace{0.1 cm} | {\mathcal A} | .
\end{eqnarray*}
\vspace{0.2 cm}
\noindent
The kernel of $\hspace{0.1 cm} Q_{q, 2} \hspace{0.1 cm} + \hspace{0.1 cm} | {\mathcal A} | \hspace{0.1 cm}$ is described as follows.
For $f \in \Omega^{q}(M_{2}, E)|_{Y}$, choose $\psi \in \Omega^{q}(M_{2}, E)$ such that
$ \mathcal{B}^{2} \psi = 0$ and $\psi|_{Y} = f$.
Then,
\begin{eqnarray} \label{E:2.55}
0 \hspace{0.1 cm} = \hspace{0.1 cm} \langle \mathcal{B}^{2} \psi, \hspace{0.1 cm} \psi \rangle & = &
\langle \mathcal{B} \psi, \hspace{0.1 cm} \mathcal{B} \psi \rangle \hspace{0.1 cm} + \hspace{0.1 cm}
\langle (\mathcal{B} \psi)|_{Y}, \hspace{0.1 cm} ({\mathcal \gamma} \psi)|_{Y} \rangle_{Y} \nonumber \\
& = & \langle \mathcal{B} \psi, \hspace{0.1 cm} \mathcal{B} \psi \rangle \hspace{0.1 cm} + \hspace{0.1 cm}
\langle (\nabla_{\partial_{u}} \psi + {\mathcal A} \psi)|_{Y}, \hspace{0.1 cm} f \rangle_{Y} \nonumber \\
& = & \parallel \mathcal{B} \psi \parallel^{2} \hspace{0.1 cm} - \hspace{0.1 cm}
\langle Q_{q, 2} f, \hspace{0.1 cm} f \rangle_{Y} \hspace{0.1 cm} + \hspace{0.1 cm}
\langle {\mathcal A} f, \hspace{0.1 cm} f \rangle_{Y} ,
\end{eqnarray}
\noindent
which leads to
\begin{eqnarray} \label{E:2.56}
\langle (Q_{q, 2} + |{\mathcal A}|)f, \hspace{0.1 cm} f \rangle_{Y} & = &
\parallel \mathcal{B} \psi \parallel^{2} \hspace{0.1 cm} + \hspace{0.1 cm}
\langle (|{\mathcal A}| + {\mathcal A})f, \hspace{0.1 cm} f \rangle_{Y}.
\end{eqnarray}
\noindent
Hence, $f \in \operatorname{ker} (Q_{q, 2} + |{\mathcal A}|)$ if and only if $\mathcal{B} \psi = 0$ and $(|{\mathcal A}| + {\mathcal A})f = 0$.
From the assumption $H^{\bullet}(Y, E|_{Y}) = 0$ ${\mathcal A}$ is an invertible operator,
which shows that $\psi$ is expressed, on a collar neighborhood of $Y$, by
\begin{equation} \label{E:2.77}
\psi = \sum_{\lambda_{j} \in \operatorname{Spec} ({\mathcal A}) \atop \lambda_{j} < 0} a_{j} e^{- \lambda_{j} u} \phi_{j}, \qquad \text{where}
\quad {\mathcal A} \phi_{j} = \lambda_{j} \phi_{j}.
\end{equation}
Let $M_{\infty} := \left( (- \infty, 0] \times Y \right) \cup_{Y} M_{2}$.
We can extend $E$ and $\mathcal{B}$ canonically to $M_{\infty}$, which we denote by
$E_{\infty}$ and $\mathcal{B}_{\infty}$. Then $\psi$ in (\ref{E:2.77}) can be extended to $M_{\infty}$ as an $L^{2}$-solution of $\mathcal{B}_{\infty}$.
Hence,
$$
\operatorname{ker} (Q_{q, 2} + |{\mathcal A}|) = \{ \psi|_{Y} \mid \psi \hspace{0.2 cm} \text{is} \hspace{0.2 cm} \text{an} \hspace{0.2 cm}
L^{2}\text{-solution} \hspace{0.2 cm} \text{of} \hspace{0.2 cm} \mathcal{B}_{\infty} \hspace{0.2 cm} \text{in}
\hspace{0.2 cm} \Omega^{q}(M_{\infty}, E_{\infty}) \hspace{0.1 cm} \}.
$$
\vspace{0.2 cm}
\noindent
It is a well known fact (Proposition 4.9 in [1]) that the space of $L^{2}$-solutions of $\mathcal{B}_{\infty}$ is isomorphic to the image of
$H^{\bullet}(M, Y ; E) \rightarrow H^{\bullet}(M ; E)$, which is zero under our assumption. This shows that $(Q_{q, 2} + |{\mathcal A}|)$
is injective and hence invertible, which leads to the following result.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:2.99}
We assume that for each $0 \leq q \leq m$, $H^{q}(M ; E) = H^{q}(M, Y ; E) = \{ 0 \}$.
Then
\begin{eqnarray*}
\lim_{r \rightarrow \infty} \log \operatorname{Det} R_{q, {\mathcal P}_{-, {\mathcal L}_{0}}/{\mathcal P}_{+, {\mathcal L}_{1}}}(r)
\hspace{0.1 cm} = \hspace{0.1 cm}
\log \operatorname{Det} \lim_{r \rightarrow \infty} R_{q, \operatorname{rel}}(r) \hspace{0.1 cm} = \hspace{0.1 cm}
\log \operatorname{Det} \left( Q_{q, 2} \hspace{0.1 cm} + \hspace{0.1 cm} | {\mathcal A}| \right) .
\end{eqnarray*}
\end{lemma}
\vspace{0.2 cm}
\noindent
Corollary \ref{Corollary:2.3} and Lemma \ref{Lemma:2.7} together with Lemma \ref{Lemma:2.99} lead to the following result.
\vspace{0.2 cm}
\begin{corollary} \label{Corollary:2.10}
We assume that for each $0 \leq q \leq m$, $H^{q}(M ; E) = H^{q}(M, Y ; E) = \{ 0 \}$.
Then :
\begin{eqnarray*}
& (1) & \lim_{r \rightarrow \infty}
\sum_{q=0}^{m} (-1)^{q+1} q \cdot \left( \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\widetilde {\mathcal P}}_{0}}(r) -
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{rel}}(r) \right) = \frac{1}{4} \sum_{q = 0}^{m-1} \log \operatorname{Det}_{2 \theta} \mathcal{B}_{Y, q}^{2}. \\
& (2) & \lim_{r \rightarrow \infty}
\sum_{q=0}^{m} (-1)^{q+1} q \cdot \left( \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\widetilde {\mathcal P}}_{1}}(r) -
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{rel}}(r) \right) = - \frac{1}{4} \sum_{q = 0}^{m-1} \log \operatorname{Det}_{2 \theta} \mathcal{B}_{Y, q}^{2}. \\
& (3) & \lim_{r \rightarrow \infty}
\sum_{q=0}^{m} (-1)^{q+1} q \cdot \left( \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}(r) -
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{rel}}(r) \right) \\
& & \hspace{1.0 cm} = \hspace{0.1 cm} \lim_{r \rightarrow \infty}
\sum_{q=0}^{m} (-1)^{q+1} q \cdot \left( \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}(r) -
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{abs}}(r) \right) \hspace{0.1 cm} = \hspace{0.1 cm} 0 .
\end{eqnarray*}
\end{corollary}
\vspace{0.2 cm}
The following lemma is well known (cf. [4], [20]).
\begin{lemma} \label{Lemma:2.11}
Let $M$ be a compact manifold with boundary $Y$ and $N$ be a collar neighborhood of $Y$.
We suppose that $\{ g_{t}^{M} \mid - \delta_{0} < t < \delta_{0} \}$ is a family of metrics such that each $g^{M}_{t}$ is a product metric and
does not vary on $N$.
Let ${\frak D}$ be one of the following boundary conditions :
${\widetilde {\mathcal P}}_{0}$, ${\widetilde {\mathcal P}}_{1}$, the absolute or the relative boundary condition.
We denote by $\mathcal{B}^{2}_{q, {\frak D}}(t)$ the square of the odd signature operator acting on $q$-forms subject to ${\frak D}$ with respect to the metric $g^{M}_{t}$.
If $H^{q}(M ; E) = H^{q}(M, Y ; E) = \{ 0 \}$ for each $0 \leq q \leq m$, then we have
$$
\frac{d}{dt} \left( \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det} \mathcal{B}^{2}_{q, {\frak D}}(t) \right) \hspace{0.1 cm} = \hspace{0.1 cm} 0.
$$
\end{lemma}
\vspace{0.3 cm}
We fix $\delta_{0} > 0$ sufficiently small and
choose a smooth function $\hspace{0.1 cm} f(r, u) : [0, \infty) \times [0, 1] \rightarrow [0, \infty), \hspace{0.1 cm}$ ($r \geq 1$) such that
$$
\operatorname{supp}_{u}f(r, u) \subset [\delta_{0}, 1 - \delta_{0}], \quad \int_{0}^{1} f(r, u) du = r - 1 , \quad
\text{and} \quad f(1, u) \equiv 0.
$$
Setting $F(r, u) = u + \int_{0}^{u} f(r, t) dt$,
$\quad F_{r} := F(r, \cdot) : [0, 1] \rightarrow [0, r] \hspace{0.1 cm}$ is a diffeomorphism satisfying
\begin{eqnarray*}
F_{r}(u) = \begin{cases} u \hspace{1.2 cm} \text{for} \hspace{0.2 cm} 0 \leq u \leq \delta_{0} \\
u + r -1 \hspace{0.6 cm} \text{for} \hspace{0.2 cm} 1 - \delta_{0} \leq u \leq 1.
\end{cases}
\end{eqnarray*}
\noindent
Let $g^{M}_{r}$ be a metric on $M_{r} : = \left( [0, r] \times Y \right) \cup_{\{ r \} \times Y} M_{2}$
which is a product one on $[0, r] \times Y$.
Then $F_{r}^{\ast}g^{M}_{r}$ is a metric on $M$, which is
$\left( \begin{array}{clcr} F^{\prime}(u)^{2} & 0 \\ 0 & g_{Y} \end{array} \right)$ on $[0, 1] \times Y$.
Hence, $F_{r}^{\ast}g^{M}_{r}$ is a metric on $M$ which is a product one near $Y$. Furthermore,
$(M, F_{r}^{\ast}g^{M}_{r})$ and $(M_{r}, g^{M}_{r})$ are isometric.
Let ${\tilde \mathcal{B}}(r)$ and $\mathcal{B}(r)$ be the odd signature operators defined on $M$ and $M_{r}$ associated to the metrics $F_{r}^{\ast}g^{M}_{r}$
and $g^{M}_{r}$, respectively.
We now assume that for each $0 \leq q \leq m$, $H^{q}(M ; E) = H^{q}(M, Y ; E) = \{ 0 \}$.
Then ${\widetilde \mathcal{B}}^{2}_{q, {\frak D}}(r)$ and $\mathcal{B}^{2}_{q, {\frak D}}(r)$ are invertible operators.
Lemma \ref{Lemma:2.11} leads to the following equalities.
\begin{eqnarray*}
& & \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot
\left( \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, {\widetilde {\mathcal P}}_{0}}^{2} - \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, \operatorname{rel}}^{2} \right) \\
& = & \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot
\left( \log \operatorname{Det}_{2 \theta} {\tilde \mathcal{B}(r)^{2}_{q, {\widetilde {\mathcal P}}_{0}}} -
\log \operatorname{Det}_{2 \theta} {\tilde \mathcal{B}(r)^{2}_{q, \operatorname{rel}}} \right) \\
& = & \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot
\left( \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\widetilde {\mathcal P}}_{0}}(r) - \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{rel}}(r) \right) \\
& = & \lim_{r \rightarrow \infty} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot
\left( \log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, {\widetilde {\mathcal P}}_{0}}(r) -
\log \operatorname{Det}_{2 \theta} \mathcal{B}^{2}_{q, \operatorname{rel}}(r) \right)
\hspace{0.1 cm} = \hspace{0.1 cm} \frac{1}{4} \sum_{q = 0}^{m-1} \log \operatorname{Det}_{2 \theta} \mathcal{B}_{Y, q}^{2}.
\end{eqnarray*}
\noindent
Similarly, we have
$$
\sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot
\left( \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, {\widetilde {\mathcal P}}_{1}}^{2} - \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, \operatorname{rel}}^{2} \right)
\hspace{0.1 cm} = \hspace{0.1 cm} - \frac{1}{4} \sum_{q = 0}^{m-1} \log \operatorname{Det}_{2 \theta} \mathcal{B}_{Y, q}^{2}.
$$
\noindent
Corollary \ref{Corollary:2.3}, Corollary \ref{Corollary:2.10}, the Poincar\'e duality and the above equality
lead to the following theorem, which is the main result of this section.
\vspace{0.2 cm}
\begin{theorem} \label{Theorem:2.11}
Let $(M, g^{M})$ be a compact Riemannian manifold with boundary $Y$ and $g^{M}$ be a product metric near $Y$.
We assume that for each $0 \leq q \leq m$, $H^{q}(M ; E) = H^{q}(M, Y ; E) = \{ 0 \}$. Then :
\begin{eqnarray*}
& (1) & \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, {\widetilde {\mathcal P}}_{0}}^{2} \hspace{0.1 cm} = \hspace{0.1 cm}
\sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, \operatorname{rel}}^{2}
\hspace{0.1 cm} + \hspace{0.1 cm} \frac{1}{4} \sum_{q = 0}^{m-1} \log \operatorname{Det}_{2 \theta} \mathcal{B}_{Y, q}^{2} \\
& (2) & \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, {\widetilde {\mathcal P}}_{1}}^{2} \hspace{0.1 cm} = \hspace{0.1 cm}
\sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, \operatorname{rel}}^{2}
\hspace{0.1 cm} - \hspace{0.1 cm} \frac{1}{4} \sum_{q = 0}^{m-1} \log \operatorname{Det}_{2 \theta} \mathcal{B}_{Y, q}^{2} \\
& (3) & \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, {\mathcal P}_{-, {\mathcal L}_{0}}}^{2} \hspace{0.1 cm} = \hspace{0.1 cm}
\sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, {\mathcal P}_{+, {\mathcal L}_{1}}}^{2} \\
& & \hspace{0.2 cm} = \hspace{0.1 cm} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, \operatorname{rel}}^{2}
\hspace{0.1 cm} = \hspace{0.1 cm} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2 \theta} \mathcal{B}_{q, \operatorname{abs}}^{2}
\end{eqnarray*}
\end{theorem}
\vspace{0.3 cm}
\section{Comparison of the eta invariants}
\vspace{0.2 cm}
In this section we are going to compare the eta-invariant $\eta(\mathcal{B}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}})$
with $\eta(\mathcal{B}_{\operatorname{even}, \Pi_{>, {\mathcal L}_{0}}})$, the eta-invariant of $\mathcal{B}_{\operatorname{even}}$ subject to ${\mathcal P}_{-, {\mathcal L}_{0}}$
and the generalized APS boundary condition $\Pi_{>, {\mathcal L}_{0}}$,
where $\Pi_{>} : \Omega^{\operatorname{even}}(M, E)|_{Y} \rightarrow \Omega^{\operatorname{even}}(M, E)|_{Y}$ is the orthogonal projection onto the space spanned by
the positive eigenspaces of ${\mathcal A}$ (cf. (\ref{E:1.9})).
For this purpose we are going to follow the arguments in [6] strongly.
Throughout this section we write the odd signature operator acting on $\Omega^{\operatorname{even}}(M, E)$ by $\mathcal{B}$
rather than $\mathcal{B}_{\operatorname{even}}$ for simplicity.
We begin with the descriptions of $\operatorname{Im} \Pi_{>}$ and $\operatorname{Im} {\mathcal P}_{-}$ as graphs of some unitary operators.
We denote by $\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}$ the orthogonal complement of
$\left( \begin{array}{clcr} {\mathcal H}^{\operatorname{even}}(Y, E|_{Y}) \\ {\mathcal H}^{\operatorname{odd}}(Y, E|_{Y}) \end{array} \right)$ in
$\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)$.
Then the action of the unitary operator ${\mathcal \gamma}$ splits according to the following decomposition.
$$
{\mathcal \gamma} :
\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast} \oplus
\left( \begin{array}{clcr} {\mathcal H}^{\operatorname{even}}(Y, E|_{Y}) \\ {\mathcal H}^{\operatorname{odd}}(Y, E|_{Y}) \end{array} \right)
\hspace{0.1 cm} \rightarrow \hspace{0.1 cm} \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast} \oplus
\left( \begin{array}{clcr} {\mathcal H}^{\operatorname{even}}(Y, E|_{Y}) \\ {\mathcal H}^{\operatorname{odd}}(Y, E|_{Y}) \end{array} \right)
$$
\vspace{0.2 cm}
\noindent
Since $\hspace{0.1 cm} {\mathcal \gamma}^{2} = - \operatorname{Id} \hspace{0.1 cm}$, we denote
the $\pm i$-eigenspace of ${\mathcal \gamma}$ in $\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}$ by
$\hspace{0.1 cm} \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{\pm i} \hspace{0.1 cm}$, which are
\begin{eqnarray} \label{E:3.1}
\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{\pm i} & = & \frac{I \mp i {\mathcal \gamma}}{2}
\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast} .
\end{eqnarray}
\vspace{0.2 cm}
\noindent
It is a well known fact that $\operatorname{Im} \Pi_{>}$ and $\operatorname{Im} {\mathcal P}_{-}$ are expressed by the graphs of some unitary operators from
$\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+i}$ to $\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{-i}$.
When restricted to $\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}$, $B_{Y}^{2}$ is an invertible operator and we denote its inverse by
$\left( B_{Y}^{2} \right)^{-1}$.
In view of (\ref{E:1.3}) we define $U_{\Pi_{>}}$, $U_{{\mathcal P}_{-}}$ as follows.
\begin{equation} \label{E:3.51}
U_{\Pi_{>}}, \hspace{0.2 cm} U_{{\mathcal P}_{-}} : \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+i}
\rightarrow \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{-i}
\end{equation}
\begin{eqnarray} \label{E:3.52}
U_{\Pi_{>}} & = & (\mathcal{B}_{Y}^{2})^{-\frac{1}{2}} \left( \nabla^{Y} + \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \right) \left( \begin{array}{clcr} 0 & -1 \\
-1 & 0 \end{array} \right) \nonumber \\
U_{{\mathcal P}_{-}} & = & (\mathcal{B}_{Y}^{2})^{- 1} \left( (\mathcal{B}_{Y}^{2})^{-} - (\mathcal{B}_{Y}^{2})^{+} \right) \left( \begin{array}{clcr} 1 & 0 \\
0 & 1 \end{array} \right) ,
\end{eqnarray}
\vspace{0.2 cm}
\noindent
where
$(\mathcal{B}_{Y}^{2})^{-} := \nabla^{Y} \Gamma^{Y} \nabla^{Y} \Gamma^{Y} : \Omega^{\bullet}_{-}(Y, E|_{Y}) \rightarrow \Omega^{\bullet}_{-}(Y, E|_{Y})$ and
$(\mathcal{B}_{Y}^{2})^{+} := \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \nabla^{Y} : \Omega^{\bullet}_{+}(Y, E|_{Y}) \rightarrow \Omega^{\bullet}_{+}(Y, E|_{Y})$.
Then $U_{\Pi_{>}}$ and $U_{{\mathcal P}_{-}}$ are well defined $\Psi$DO's and their adjoints are given by
\begin{equation} \label{E:3.53}
U_{\Pi_{>}}^{\ast}, \hspace{0.2 cm} U_{{\mathcal P}_{-}}^{\ast} : \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{-i}
\rightarrow \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+i}
\end{equation}
\begin{eqnarray} \label{E:3.54}
U_{\Pi_{>}}^{\ast} & = & (\mathcal{B}_{Y}^{2})^{-\frac{1}{2}} \left( \nabla^{Y} + \Gamma^{Y} \nabla^{Y} \Gamma^{Y} \right) \left( \begin{array}{clcr} 0 & -1 \\
-1 & 0 \end{array} \right) \nonumber \\
U_{{\mathcal P}_{-}}^{\ast} & = & (\mathcal{B}_{Y}^{2})^{- 1} \left( (\mathcal{B}_{Y}^{2})^{-} - (\mathcal{B}_{Y}^{2})^{+} \right) \left( \begin{array}{clcr} 1 & 0 \\
0 & 1 \end{array} \right) .
\end{eqnarray}
\vspace{0.2 cm}
\noindent
The following lemma is straightforward.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.1}
(1) Both $U_{\Pi_{>}}$ and $U_{{\mathcal P}_{-}}$ are unitary operators satisfying
$$
U_{\Pi_{>}}^{\ast} \hspace{0.1 cm} U_{\Pi_{>}} \hspace{0.1 cm} = \hspace{0.1 cm} U_{{\mathcal P}_{-}}^{\ast} \hspace{0.1 cm} U_{{\mathcal P}_{-}}
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Id}, \qquad
{\mathcal \gamma} \hspace{0.1 cm} U_{\Pi_{>}} = - \hspace{0.1 cm} U_{\Pi_{>}} \hspace{0.1 cm} {\mathcal \gamma}, \qquad
{\mathcal \gamma} \hspace{0.1 cm} U_{{\mathcal P}_{-}} = - \hspace{0.1 cm} U_{{\mathcal P}_{-}} \hspace{0.1 cm} {\mathcal \gamma}.
$$
\noindent
(2) $\hspace{0.2 cm} \operatorname{Im} \Pi_{>}$ $(\hspace{0.1 cm} \operatorname{Im} \Pi_{<} \hspace{0.1 cm})$ and $\hspace{0.1 cm} \operatorname{Im} {\mathcal P}_{-}$
$(\hspace{0.1 cm} \operatorname{Im} {\mathcal P}_{+} \hspace{0.1 cm})$ are graphs of
$\hspace{0.1 cm} U_{\Pi_{>}}$ $(\hspace{0.1 cm} - \hspace{0.1 cm} U_{\Pi_{>}} \hspace{0.1 cm})$ and
$\hspace{0.1 cm} U_{{\mathcal P}_{-}}$ $( \hspace{0.1 cm} - \hspace{0.1 cm} U_{{\mathcal P}_{-}} \hspace{0.1 cm})$,
respectively, {\it i.e.}
\begin{eqnarray*}
\operatorname{Im} \Pi_{>} = \{ x + U_{\Pi_{>}} x \mid x \in \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+ i} \}, & &
\operatorname{Im} \Pi_{<} = \{ x - U_{\Pi_{>}} x \mid x \in \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+ i} \}, \\
\operatorname{Im} {\mathcal P}_{-} = \{ x + U_{{\mathcal P}_{-}} x \mid x \in \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+ i} \}, & &
\operatorname{Im} {\mathcal P}_{+} = \{ x - U_{{\mathcal P}_{-}} x \mid x \in \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+ i} \}.
\end{eqnarray*}
(3) $\hspace{0.2 cm} U_{\Pi_{>}}$ anticommutes with $U_{{\mathcal P}_{-}}$ in the following sense, {\it i.e.}
$$ U_{\Pi_{>}}^{\ast} \hspace{0.1 cm} U_{{\mathcal P}_{-}} = - \hspace{0.1 cm} U_{{\mathcal P}_{-}}^{\ast} \hspace{0.1 cm} U_{\Pi_{>}}, \qquad
U_{\Pi_{>}} \hspace{0.1 cm} U_{{\mathcal P}_{-}}^{\ast} = - \hspace{0.1 cm} U_{{\mathcal P}_{-}} \hspace{0.1 cm} U_{\Pi_{>}}^{\ast}.
$$
\end{lemma}
\vspace{0.4 cm}
We define $P(\theta) : \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{i} \rightarrow \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{-i}$ by
\begin{equation}
P(\theta) := U_{\Pi_{>}} cos \theta + U_{{\mathcal P}_{-}} sin \theta, \qquad (0 \leq \theta \leq \frac{\pi}{2}).
\end{equation}
\vspace{0.2 cm}
\noindent
Then $P(\theta)$ is a unitary operator satisfying the property (1) in Lemma \ref{Lemma:3.1} and a smooth path connecting $U_{\Pi_{>}}$ and
$U_{{\mathcal P}_{-}}$.
We here note that the orthogonal projections $\hspace{0.1 cm} \Pi_{>}$, ${\mathcal P}_{-} :
\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+i} \oplus \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{-i}
\rightarrow \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{+i} \oplus \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}_{-i} \hspace{0.1 cm}$
are expressed as follows.
\vspace{0.2 cm}
$$
\Pi_{>} = \frac{1}{2} \left( \begin{array}{clcr} \operatorname{Id} & U_{\Pi_{>}}^{\ast} \\ U_{\Pi_{>}} & \operatorname{Id} \end{array} \right) {\mathcal P}_{\ast} , \qquad
{\mathcal P}_{-} = \frac{1}{2} \left( \begin{array}{clcr} \operatorname{Id} & U_{{\mathcal P}_{-}}^{\ast} \\ U_{{\mathcal P}_{-}} & \operatorname{Id}
\end{array} \right) {\mathcal P}_{\ast} ,
$$
\vspace{0.2 cm}
\noindent
where ${\mathcal P}_{\ast}$ is the orthogonal projection onto $\left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)^{\ast}$.
Let ${\mathcal L}_{0} = \left( \begin{array}{clcr} {\mathcal K} \\ {\mathcal K} \end{array} \right) \cap \hspace{0.1 cm} \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)$
and ${\mathcal L}_{1} = \left( \begin{array}{clcr} \Gamma^{Y} {\mathcal K} \\
\Gamma^{Y} {\mathcal K} \end{array} \right) \cap \hspace{0.1 cm} \left( \Omega^{\operatorname{even}}(M, E)|_{Y} \right)$
so that ${\mathcal L}_{0} \oplus {\mathcal L}_{1} = \left( \begin{array}{clcr} {\mathcal H}^{\operatorname{even}}(Y, E|_{Y}) \\ {\mathcal H}^{\operatorname{odd}}(Y, E|_{Y})
\end{array} \right)$.
We denote by ${\mathcal P}_{{\mathcal L}_{0}}$ and ${\mathcal P}_{{\mathcal L}_{1}}$ the orthogonal projections onto ${\mathcal L}_{0}$ and ${\mathcal L}_{1}$.
We define the orthogonal projections ${\mathcal P}_{-, {\mathcal L}_{0}}$ and $\Pi_{>, {\mathcal L}_{0}}$ on $\Omega^{\operatorname{even}}(M, E)|_{Y}$ as follows.
\vspace{0.2 cm}
\begin{equation} \label{E:3.77}
{\mathcal P}_{-, {\mathcal L}_{0}} := {\mathcal P}_{-} + {\mathcal P}_{{\mathcal L}_{0}}, \qquad
\Pi_{>, {\mathcal L}_{0}} := \Pi_{>} + {\mathcal P}_{{\mathcal L}_{0}}
\end{equation}
\vspace{0.2 cm}
\noindent
We define ${\mathcal P}_{+, {\mathcal L}_{1}}$ and $\Pi_{<, {\mathcal L}_{1}}$ in the same way.
Similarly, we define the orthogonal projection ${\widetilde P}(\theta)$ by
\vspace{0.2 cm}
\begin{equation}
{\widetilde P}(\theta) := \frac{1}{2} \left( \begin{array}{clcr} \operatorname{Id} & P(\theta)^{\ast} \\ P(\theta) & \operatorname{Id} \end{array} \right) {\mathcal P}_{\ast}
+ {\mathcal P}_{{\mathcal L}_{0}}
= \Pi_{>} cos \theta + {\mathcal P}_{-} sin \theta + \frac{1}{2} ( 1 - cos \theta - sin \theta ) \hspace{0.1 cm} {\mathcal P}_{\ast}
+ {\mathcal P}_{{\mathcal L}_{0}}.
\end{equation}
\vspace{0.2 cm}
\noindent
${\widetilde P}(\theta)$ satisfies the following properties.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.2}
(1) ${\mathcal \gamma} \hspace{0.1 cm} {\widetilde P}(\theta) \hspace{0.1 cm} = \hspace{0.1 cm} (I - {\widetilde P}(\theta))
\hspace{0.1 cm} {\mathcal \gamma}, \hspace{0.2 cm}$ and
$\hspace{0.2 cm} {\widetilde P}(\theta) \hspace{0.1 cm} \mathcal{B}_{Y}^{2} \hspace{0.1 cm} = \hspace{0.1 cm} \mathcal{B}_{Y}^{2} {\widetilde P}(\theta)$. \newline
(2) ${\widetilde P}(\theta) \hspace{0.1 cm} {\mathcal A} \hspace{0.1 cm} {\widetilde P}(\theta)
\hspace{0.1 cm} = \hspace{0.1 cm} cos \theta \hspace{0.1 cm} \vert {\mathcal A} \vert \hspace{0.1 cm} {\widetilde P}(\theta)
\hspace{0.1 cm} = \hspace{0.1 cm} cos \theta \hspace{0.1 cm} \sqrt{(\mathcal{B}_{Y}^{2})} \hspace{0.1 cm} {\widetilde P}(\theta)$.
\end{lemma}
\vspace{0.2 cm}
\begin{proof} : The proofs are straightforward. For the second statement we may need the following identities.
$$
\Pi_{>} {\mathcal P}_{-} \Pi_{>} \hspace{0.1 cm} = \hspace{0.1 cm} \frac{1}{2} \Pi_{>}, \qquad
{\mathcal P}_{-} \Pi_{>} \hspace{0.1 cm} + \hspace{0.1 cm} \Pi_{>} {\mathcal P}_{-} \hspace{0.1 cm} = \hspace{0.1 cm}
\left( {\mathcal P}_{-} \hspace{0.1 cm} + \hspace{0.1 cm} \Pi_{>} \hspace{0.1 cm} - \hspace{0.1 cm} \frac{1}{2} \operatorname{Id} \right) {\mathcal P}_{\ast}.
$$
\end{proof}
\begin{lemma} \label{Lemma:3.33}
Let $\mathcal{B}_{{\widetilde P}(\theta)}$ be the realization of $\mathcal{B}$ with respect to ${\widetilde P}(\theta)$, {\it i.e.} \newline
$\operatorname{Dom} \left( \mathcal{B}_{{\widetilde P}(\theta)} \right) = \{ \phi \in H^{1} \left( \Omega^{\operatorname{even}}(M, E) \right) \mid {\widetilde P}(\theta) (\phi|_{Y}) = 0 \}$.
Then $\mathcal{B}_{{\widetilde P}(\theta)}$ is essentially self-adjoint.
\end{lemma}
\noindent
\begin{proof} :
It was shown in [24] (cf. [12]) that the adjoint $\left( \mathcal{B}_{{\widetilde P}(\theta)} \right)^{\ast}$ is the realization of $\mathcal{B}^{\ast} = \mathcal{B}$ with
respect to the boundary condition $\left( I - {\widetilde P}(\theta) \right) \gamma^{\ast}$, {\it i.e.}
$$
\operatorname{Dom} \left( \mathcal{B}_{{\widetilde P}(\theta)} \right)^{\ast} \hspace{0.1 cm} = \hspace{0.1 cm}
\{ \phi \in H^{1} \left( \Omega^{\operatorname{even}}(M, E) \right) \mid \left( I - {\widetilde P}(\theta) \right) \gamma^{\ast} (\phi|_{Y}) = 0 \}
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Dom} \left( \mathcal{B}_{{\widetilde P}(\theta)} \right).
$$
\noindent
Hence, it's enough to show that $\mathcal{B}_{{\widetilde P}(\theta)}$ is a symmetric operator.
For $\phi$, $\psi \in \operatorname{Dom} \left( \mathcal{B}_{{\widetilde P}(\theta)} \right)$,
\begin{eqnarray*}
\langle \mathcal{B} \phi, \hspace{0.1 cm} \psi \rangle_{M} - \langle \phi, \hspace{0.1 cm} \mathcal{B} \psi \rangle_{M} & = &
\hspace{0.1 cm} \langle \phi|_{Y}, \hspace{0.1 cm} {\mathcal \gamma} \left( \psi|_{Y} \right) \rangle_{Y} \nonumber \\
& = & \hspace{0.1 cm} \langle (I - {\widetilde P}(\theta)) \phi|_{Y},
\hspace{0.1 cm} {\mathcal \gamma} (I - {\widetilde P}(\theta)) \left( \psi|_{Y} \right) \rangle_{Y}
\hspace{0.1 cm} = \hspace{0.1 cm} \hspace{0.1 cm} \langle (I - {\widetilde P}(\theta)) \phi|_{Y},
\hspace{0.1 cm} {\widetilde P}(\theta) {\mathcal \gamma} \left( \psi|_{Y} \right) \rangle_{Y} \nonumber \\
& = & \hspace{0.1 cm} \langle {\widetilde P}(\theta) (I - {\widetilde P}(\theta)) \phi|_{Y},
\hspace{0.1 cm} {\mathcal \gamma} \left( \psi|_{Y} \right) \rangle_{Y} \hspace{0.1 cm} = \hspace{0.1 cm} 0,
\end{eqnarray*}
\noindent
which completes the proof of the lemma.
\end{proof}
\vspace{0.2 cm}
Setting
$$
U(\theta) \hspace{0.1 cm} = \hspace{0.1 cm}
\left( \begin{array}{clcr} P(\theta)^{\ast} \hspace{0.1 cm} U_{\Pi_{>}} & 0 \\ 0 & \operatorname{Id} \end{array} \right) {\mathcal P}_{\ast}
+ \left( \operatorname{Id} - {\mathcal P}_{\ast} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\left( \begin{array}{clcr} cos \theta + U_{{\mathcal P}_{-}}^{\ast} U_{\Pi_{>}} sin \theta & 0 \\ 0 & \operatorname{Id} \end{array} \right) {\mathcal P}_{\ast}
+ \left( \operatorname{Id} - {\mathcal P}_{\ast} \right),
$$
\noindent
it is straightforward that
\begin{equation}
U(\theta) {\widetilde P}(0) U(\theta)^{\ast} = {\widetilde P}(\theta).
\end{equation}
\noindent
Moreover, setting
\begin{equation} \label{E:3.111}
\hspace{0.1 cm} T(\theta) = -i \hspace{0.1 cm} \theta \left( \begin{array}{clcr} U_{{\mathcal P}_{-}}^{\ast} U_{\Pi_{>}} & 0 \\ 0 & 0 \end{array} \right) {\mathcal P}_{\ast},
\end{equation}
\noindent
$T(\theta)$ is a self-adjoint operator and we have
\begin{equation}
exp\{ i T(\theta) \} = U(\theta).
\end{equation}
\noindent
$T(\theta)$ satisfies the following property.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.22}
$T(\theta)$ commutes with ${\mathcal \gamma}$ and $\mathcal{B}_{Y}^{2}$, {\it i.e.},
\begin{equation} \label{E:3.7}
{\mathcal \gamma} \hspace{0.1 cm} T(\theta) \hspace{0.1 cm} = \hspace{0.1 cm} T(\theta) \hspace{0.1 cm} {\mathcal \gamma}, \qquad
\mathcal{B}_{Y}^{2} \hspace{0.1 cm} T(\theta) \hspace{0.1 cm} = \hspace{0.1 cm} T(\theta) \hspace{0.1 cm} \mathcal{B}_{Y}^{2}.
\end{equation}
\end{lemma}
\vspace{0.2 cm}
\noindent
{\it Remark} : Contrary to the case of [6], $T(\theta)$ does not anticommute with ${\mathcal A}$.
\vspace{0.2 cm}
Let $\phi : [0, 1] \rightarrow [0, 1]$ be a decreasing smooth function such that $\phi = 1$ on a small neighborhood of $0$ and
$\phi = 0$ on a small neighborhood of $1$. We use this cut-off function to extend $T(\theta)$ defined on $\Omega^{\operatorname{even}}(M, E)|_{Y}$
to an operator defined on $\Omega^{\operatorname{even}}(M, E)$.
We define
$\Psi_{\theta} : \Omega^{\operatorname{even}}(M, E) \rightarrow \Omega^{\operatorname{even}}(M, E)$ by
\begin{equation} \label{E:3.8}
\Psi_{\theta} (\omega) (x) = e^{i \phi(x) T(\theta)} \omega (x),
\end{equation}
\noindent
where the support of $\phi(x) T(\theta)$ is contained in $N$, the collar neighborhood of $Y$.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.3}
$\Psi_{\theta}$ is a unitary operator mapping from $\operatorname{Dom} \left(\mathcal{B}_{{\widetilde P}(0)} \right)$ onto $\operatorname{Dom} \left(\mathcal{B}_{{\widetilde P}(\theta)} \right)$.
\end{lemma}
\begin{proof} : Clearly $\Psi_{\theta}$ is a unitary operator.
Let ${\widetilde P}(0) \omega(0) = 0$. Then
\begin{eqnarray*}
{\widetilde P}(\theta) (\Psi_{\theta} \omega) (0) & = & U(\theta) {\widetilde P}(0) U(\theta)^{\ast} \left( e^{i \phi(x) T(\theta)} \omega \right)|_{x=0} \\
& = & U(\theta) {\widetilde P}(0) e^{- i T(\theta)} \left( e^{i \phi(x) T(\theta)} \omega \right)|_{x=0} \hspace{0.1 cm} = \hspace{0.1 cm}
U(\theta) \left( {\widetilde P}(0) \omega(0) \right) \hspace{0.1 cm} = \hspace{0.1 cm} 0,
\end{eqnarray*}
which completes the proof of the lemma.
\end{proof}
\vspace{0.2 cm}
\noindent
We now consider the following diagram.
\vspace{0.3 cm}
$$\begin{CD}
\operatorname{Dom} \left( \mathcal{B}_{{\widetilde P}(0)} \right) & @> \mathcal{B}_{{\widetilde P}(0)} >> & \Omega^{\operatorname{even}}(M, E) \\
@V \Psi_{\theta} VV & & @V{\Psi_{\theta}}VV \\
\operatorname{Dom} \left( \mathcal{B}_{{\widetilde P}(\theta)} \right) & @> \mathcal{B}_{{\widetilde P}(\theta)} >> & \Omega^{\operatorname{even}}(M, E)
\end{CD}
$$
\vspace{0.3 cm}
\noindent
Setting $\mathcal{B}(\theta) := \Psi_{\theta}^{\ast} \mathcal{B}_{{\widetilde P}(\theta)} \Psi_{\theta}$,
$$
\mathcal{B}(\theta) : \operatorname{Dom} \left(\mathcal{B}_{{\widetilde P}(0)} \right) \rightarrow \Omega^{\operatorname{even}}(M, E)
$$
is an elliptic $\Psi$DO of order $1$
with a fixed domain $\operatorname{Dom} \left(\mathcal{B}_{{\widetilde P}(0)} \right)$ and have the same spectrum as $\mathcal{B}_{{\widetilde P}(\theta)}$.
\vspace{0.2 cm}
We next discuss one parameter family of eta functions $\eta_{\mathcal{B}(\theta)}(s)$ defined by
\vspace{0.2 cm}
\begin{equation} \label{E:3.9}
\eta_{\mathcal{B}(\theta)}(s) = \frac{1}{\Gamma(\frac{s+1}{2})} \int_{0}^{\infty} t^{\frac{s-1}{2}} \operatorname{Tr} \left( \mathcal{B}(\theta) e^{-t \mathcal{B}(\theta)^{2}} \right) dt.
\end{equation}
\vspace{0.2 cm}
\noindent
If $\eta_{\mathcal{B}(\theta)}(s)$ has a regular value at $s=0$, we define the eta invariant $\eta (\mathcal{B}(\theta))$ by
\begin{equation} \label{E:3.99}
\eta (\mathcal{B}(\theta)) = \frac{1}{2} \left( \eta_{\mathcal{B}(\theta)}(0) + \operatorname{dim} \operatorname{ker} \mathcal{B}(\theta) \right).
\end{equation}
\vspace{0.2 cm}
\noindent
For $0 \leq \theta_{0} \leq \frac{\pi}{2}$, there exist $c(\theta_{0}) > 0$ and $\delta > 0$ such that
$c(\theta_{0}) \notin \operatorname{Spec} \left( \mathcal{B}_{\theta} \right)$ for $\theta_{0} - \delta < \theta < \theta_{0} + \delta$.
We denote by $Q(\theta)$ the orthogonal projection onto the space spanned by eigensections of $\mathcal{B}(\theta)$ whose eigenvalues are
less than $c(\theta)$ for $\theta_{0} - \delta < \theta < \theta_{0} + \delta$.
We define
\begin{eqnarray*}
\eta_{\mathcal{B}(\theta)} \left( s \hspace{0.1 cm} ; \hspace{0.1 cm} c(\theta) \right) \hspace{0.1 cm} =
\sum_{| \lambda | > c(\theta)} sign (\lambda) \vert \lambda \vert^{-s}
\hspace{0.1 cm} = \hspace{0.1 cm} \frac{1}{\Gamma(\frac{s+1}{2})} \int^{\infty}_{0} t^{\frac{s-1}{2}}
\operatorname{Tr} \left\{ \left( I - Q(\theta) \right) \mathcal{B}(\theta) e^{- t \mathcal{B}(\theta)^{2}} \right\} dt.
\end{eqnarray*}
\noindent
Then $\eta_{\mathcal{B}(\theta)} (s) - \eta_{\mathcal{B}(\theta)} \left( s \hspace{0.1 cm} ; c(\theta) \right)$ is an entire function and
\vspace{0.2 cm}
\begin{equation} \label{E:3.155}
\left\{ \frac{1}{2} \left( \eta_{\mathcal{B}(\theta)} (s) + \operatorname{dim} \operatorname{ker} \mathcal{B}(\theta) \right) - \frac{1}{2} \eta_{\mathcal{B}(\theta)} \left( s \hspace{0.1 cm} ; c(\theta) \right) \right\}_{s=0}
\end{equation}
\vspace{0.2 cm}
\noindent
does not depend on $\theta$ for $\theta_{0} - \delta < \theta < \theta_{0} + \delta$ up to $\operatorname{mod} {\Bbb Z}$.
Simple computation shows that
\begin{eqnarray} \label{E:3.144}
& & \frac{\partial}{\partial \theta} \eta_{\mathcal{B}(\theta)} \left( s \hspace{0.1 cm} ; c(\theta) \right) \\
& = & \frac{1}{\Gamma(\frac{s+1}{2})} \int^{\infty}_{0} t^{\frac{s-1}{2}} \operatorname{Tr} \left( - {\dot Q}(\theta) \mathcal{B}(\theta) e^{-t \mathcal{B}(\theta)^{2}} +
\left( I - Q(\theta) \right) \frac{\partial}{\partial \theta} \left( \mathcal{B}(\theta) e^{- t \mathcal{B}(\theta)^{2}} \right) \right) dt \nonumber \\
& = & \frac{1}{\Gamma(\frac{s+1}{2})} \int^{\infty}_{0} t^{\frac{s-1}{2}} \operatorname{Tr} \left( - {\dot Q}(\theta) \mathcal{B}(\theta) e^{-t \mathcal{B}(\theta)^{2}} \right) dt -
\frac{s}{\Gamma(\frac{s+1}{2})} \int^{\infty}_{0} t^{\frac{s-1}{2}}
\operatorname{Tr} \left\{ \left( I - Q(\theta) \right) \left( {\dot \mathcal{B}}(\theta) e^{- t \mathcal{B}(\theta)^{2}} \right) \right\} dt \nonumber,
\end{eqnarray}
\vspace{0.2 cm}
\noindent
where ${\dot Q(\theta)}$ and ${\dot \mathcal{B}(\theta)}$ are derivatives of $Q(\theta)$ and $\mathcal{B}(\theta)$ with respect to $\theta$.
Furthermore, we have (cf. [14])
$$
\operatorname{Tr} \left( - {\dot Q}(\theta) \mathcal{B}(\theta) e^{-t \mathcal{B}(\theta)^{2}} \right) = 0, \qquad
\left\{ \frac{s}{\Gamma(\frac{s+1}{2})} \int^{\infty}_{0} t^{\frac{s-1}{2}}
\operatorname{Tr} \left( Q(\theta) {\dot \mathcal{B}}(\theta) e^{- t \mathcal{B}(\theta)^{2}} \right) dt \right\}_{s=0} = 0.
$$
\noindent
These equalities imply that
\vspace{0.2 cm}
\begin{eqnarray} \label{E:3.156}
\frac{\partial}{\partial \theta} \eta_{\mathcal{B}(\theta)} \left( s \hspace{0.1 cm} ; c(\theta) \right) & = &
- \hspace{0.1 cm} \frac{s}{\Gamma(\frac{s+1}{2})} \int^{\infty}_{0} t^{\frac{s-1}{2}}
\operatorname{Tr} \left( {\dot \mathcal{B}}(\theta) e^{- t \mathcal{B}(\theta)^{2}} \right) dt + F(s) ,
\end{eqnarray}
\vspace{0.2 cm}
\noindent
where $F(s)$ is an analytic function at least for $\operatorname{Re} s > -1$ with $F(0) = 0$.
\vspace{0.3 cm}
Recall that
\begin{equation} \label{E:3.11}
\mathcal{B}(\theta) \hspace{0.1 cm} = \hspace{0.1 cm} \Psi_{\theta}^{\ast} \mathcal{B}_{{\widetilde P}(\theta)} \Psi_{\theta}
\hspace{0.1 cm} = \hspace{0.1 cm} e^{- i \phi(x) T(\theta)} {\mathcal \gamma} (\partial_{x} + {\mathcal A}) e^{i \phi(x) T(\theta)}.
\end{equation}
Using the fact that $T(\theta) T^{\prime}(\theta) = T^{\prime}(\theta) T(\theta)$ and Lemma \ref{Lemma:3.22}, we have
\begin{equation} \label{E:3.12}
{\dot \mathcal{B}}(\theta) \hspace{0.1 cm} = \hspace{0.1 cm}
e^{- i \phi(x) T(\theta)} \left( i \phi^{\prime}(x) {\mathcal \gamma} T^{\prime}(\theta) -
i \phi(x) {\mathcal \gamma} T^{\prime}(\theta) {\mathcal A} + i \phi(x) {\mathcal \gamma} {\mathcal A} T^{\prime}(\theta)
\right) e^{i \phi(x) T(\theta)},
\end{equation}
\noindent
which leads to
\begin{equation} \label{E:3.13}
\operatorname{Tr} \left( {\dot \mathcal{B}(\theta)} e^{-t \mathcal{B}(\theta)^{2}} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Tr} \left\{ \left( i \phi^{\prime}(x) {\mathcal \gamma} T^{\prime}(\theta) -
i \phi(x) {\mathcal \gamma} [ T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A} ] \hspace{0.1 cm} \right) e^{-t \mathcal{B}^{2}_{{\widetilde P}(\theta)}} \right\}.
\end{equation}
\vspace{0.2 cm}
\noindent
Since the support of $\phi$ is in $[0, 1]$, the support of ${\dot \mathcal{B}(\theta)}$ is in $[0, 1] \times Y$.
Let $\mathcal{B}^{\operatorname{cyl}}$ be the odd signature operator defined by (\ref{E:1.8}) on $[0, \infty) \times Y$.
The heat kernel of $\left( \mathcal{B}^{\operatorname{cyl}}_{{\widetilde P}(\theta)} \right)^{2}$ was computed in [6] as follows.
\vspace{0.2 cm}
\begin{eqnarray} \label{E:3.14}
e^{-t \left( \mathcal{B}^{\operatorname{cyl}}_{{\widetilde P}(\theta)} \right)^{2}} (x, y) & = & (4 \pi t)^{- \frac{1}{2}} \left( e^{- \frac{(x-y)^{2}}{4t}} +
(I - 2 {\widetilde P} (\theta)) e^{- \frac{(x+y)^{2}}{4t}} \right) e^{-t {\mathcal A}^{2}} \nonumber \\
& & + (\pi t)^{- \frac{1}{2}} \left( I - {\widetilde P} (\theta) \right) \int_{0}^{\infty} e^{- \frac{(x+y+z)^{2}}{4t}} {\widetilde {\mathcal A}}(\theta)
e^{{\widetilde {\mathcal A}}(\theta)z - t {\mathcal A}^{2}} dz,
\end{eqnarray}
\vspace{0.2 cm}
\noindent
where
${\widetilde {\mathcal A}}(\theta) := (I - {\widetilde P}(\theta) ) {\mathcal A} (I - {\widetilde P } (\theta) )$.
The standard theory for heat kernel ([1], [3]) implies
that the asymptotic expansions of $\operatorname{Tr} \left( {\dot \mathcal{B}(\theta)} e^{-t \mathcal{B}(\theta)^{2}} \right)$ is equal
to that of $\operatorname{Tr} \left( {\dot \mathcal{B}^{\operatorname{cyl}}(\theta)} e^{-t \left( \mathcal{B}^{\operatorname{cyl}} (\theta) \right)^{2}} \right)$ up to
$\left( e^{- \frac{c}{t}} \right)$ for some $c > 0$.
With a little abuse of notation we write $\mathcal{B}^{\operatorname{cyl}}$ by $\mathcal{B}$ again.
Equation (\ref{E:3.13}) leads to the following equality.
\begin{eqnarray} \label{E:3.15}
& & \operatorname{Tr} \left( i \phi^{\prime}(x) {\mathcal \gamma} T^{\prime}(\theta) e^{-t \mathcal{B}^{2}_{{\widetilde P}(\theta)}} \right) \\
& = & \frac{i}{\sqrt{4 \pi t}} \int_{0}^{\infty} \phi^{\prime}(x) dx \hspace{0.1 cm}
\operatorname{Tr} \left( {\mathcal \gamma} T^{\prime}(\theta) e^{-t {\mathcal A}^{2}} \right)
\nonumber \\
& + & \frac{i}{\sqrt{4 \pi t}} \int_{0}^{\infty} \phi^{\prime}(x) e^{- \frac{x^{2}}{t}} dx \hspace{0.1 cm}
\operatorname{Tr} \left( {\mathcal \gamma} T^{\prime}(\theta) (I - 2 {\widetilde P}(\theta)) e^{-t {\mathcal A}^{2}} \right) \nonumber \\
& + & \frac{i}{\sqrt{\pi t}} \int_{0}^{\infty} \int_{0}^{\infty} \phi^{\prime}(x) e^{- \frac{(2x + z)^{2}}{4t}}
\operatorname{Tr} \left\{ {\mathcal \gamma} T^{\prime}(\theta) (I - {\widetilde P}(\theta)) {\widetilde {\mathcal A}}(\theta)
e^{{\widetilde {\mathcal A}}(\theta)z - t {\mathcal A}^{2}} \right\} dz dx . \nonumber
\end{eqnarray}
\vspace{0.2 cm}
\noindent
Lemma \ref{Lemma:3.2} and Lemma \ref{Lemma:3.22} imply that
$\operatorname{Tr} \left( {\mathcal \gamma} T^{\prime}(\theta) (I - 2 {\widetilde P}(\theta)) e^{-t {\mathcal A}^{2}} \right) = 0$.
Since $\phi (x) = 1$ near $x=0$, the third integral decays exponentially as $t \rightarrow 0^{+}$.
Hence,
\begin{equation} \label{E:3.16}
\operatorname{Tr} \left( i \phi^{\prime}(x) {\mathcal \gamma} T^{\prime}(\theta) e^{-t \mathcal{B}^{2}_{{\widetilde P}(\theta)}} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\frac{-i}{\sqrt{4 \pi t}} \operatorname{Tr} \left( {\mathcal \gamma} T^{\prime}(\theta) e^{-t {\mathcal A}^{2}} \right) + O(e^{- \frac{c}{t}}).
\end{equation}
\noindent
We refer to p.456 in [6] for the proof of the following equality.
\begin{equation} \label{E:3.31}
{\widetilde {\mathcal A}}(\theta) e^{{\widetilde {\mathcal A}}(\theta)z - t {\mathcal A}^{2}} \hspace{0.1 cm} = \hspace{0.1 cm}
(- cos\theta) | {\mathcal A}| (I - {\widetilde P}(\theta)) e^{- cos\theta |{\mathcal A}| z - t {\mathcal A}^{2} }.
\end{equation}
\vspace{0.2 cm}
\noindent
Then, we have
\begin{eqnarray} \label{E:3.17}
& & \operatorname{Tr} \left( - i \phi (x) {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] e^{-t \mathcal{B}^{2}_{{\widetilde P}(\theta)}} \right) \\
& = & \frac{- i}{\sqrt{4 \pi t}} \int_{0}^{\infty} \phi(x) dx \hspace{0.1 cm} \operatorname{Tr} \left( {\mathcal \gamma}
[T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] e^{-t {\mathcal A}^{2}} \right) \nonumber \\
& + & \frac{- i}{\sqrt{4 \pi t}} \int_{0}^{\infty} \phi(x) e^{- \frac{x^{2}}{t}} dx
\hspace{0.1 cm} \operatorname{Tr} \left( {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}]
(I - 2 {\widetilde P}(\theta)) e^{-t {\mathcal A}^{2}} \right) \nonumber \\
& + & \frac{- i}{\sqrt{\pi t}} \int_{0}^{\infty} \int_{0}^{\infty} \phi(x) e^{- \frac{(2x + z)^{2}}{4t}}
\hspace{0.1 cm} \operatorname{Tr} \left\{ {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}]
(I - {\widetilde P}(\theta)) (- cos\theta) |{\mathcal A}| (I - {\widetilde P}(\theta))
e^{- cos\theta |{\mathcal A}| z - t {\mathcal A}^{2} } \right\} dz dx \nonumber \\
& = : & (I) + (II) + (III). \nonumber
\end{eqnarray}
\vspace{0.2 cm}
\noindent
Lemma \ref{Lemma:3.22} shows that
$\operatorname{Tr} \left( {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] e^{-t {\mathcal A}^{2}} \right) = 0$, which yields
\begin{equation} \label{E:3.18}
(I) \hspace{0.1 cm} = \hspace{0.1 cm} 0, \qquad
(II) \hspace{0.1 cm} = \hspace{0.1 cm}
\frac{i}{2} \operatorname{Tr} \left( {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P}(\theta) e^{-t {\mathcal A}^{2}} \right)
+ O(e^{- \frac{c}{t}}).
\end{equation}
\vspace{0.2 cm}
\noindent
Change of variables, Lemma \ref{Lemma:3.2} and Lemma \ref{Lemma:3.22} show that
\begin{equation} \label{E:3.19}
(III) \hspace{0.1 cm} = \hspace{0.1 cm}
\frac{ - 2 i cos\theta}{\sqrt{\pi}} \sqrt{t} \int_{0}^{\infty} \int_{0}^{\infty} \phi(\sqrt{t}x) e^{- (x + z)^{2}} \operatorname{Tr} \left( {\mathcal \gamma}
[T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P}(\theta) |{\mathcal A}|
e^{-2 \sqrt{t} cos\theta |{\mathcal A}| z - t {\mathcal A}^{2}} \right) dx dz .
\end{equation}
\noindent
Since $|{\mathcal A}|$ commutes with ${\mathcal A}$, $T^{\prime}(\theta)$ and ${\widetilde P}(\theta)$, we denote
\begin{equation} \label{E:3.20}
d(\lambda) := \operatorname{Tr}_{\operatorname{ker} ( |{\mathcal A}| - \lambda )} \left( {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P} (\theta) \right).
\end{equation}
\begin{eqnarray} \label{E:3.21}
(III) & = & - i cos\theta \sum_{0 \neq \lambda \in \operatorname{Spec}(|{\mathcal A}|)} d(\lambda) \int_{0}^{\infty} \int_{0}^{\infty} \phi(\sqrt{t}x)
\frac{2}{\sqrt{\pi}} \sqrt{t} \lambda e^{-(x+z)^{2}} e^{-2 cos\theta \sqrt{t} \lambda z - t \lambda^{2}} dx dz \nonumber \\
& = & - i cos\theta \sum_{0 \neq \lambda \in \operatorname{Spec}(|{\mathcal A}|)} d(\lambda) \int_{0}^{\infty}
\frac{2}{\sqrt{\pi}} \int_{0}^{\infty} e^{-(x+z)^{2}} dx \sqrt{t} \lambda e^{-2 cos\theta \sqrt{t} \lambda z - t \lambda^{2}}dz
+ O(e^{- \frac{c}{t}}) \nonumber \\
& = & - i cos\theta \sum_{0 \neq \lambda \in \operatorname{Spec}(|{\mathcal A}|)} d(\lambda) \int_{0}^{\infty} erfc(z)
\sqrt{t} \lambda e^{-2 cos\theta \sqrt{t} \lambda z - t \lambda^{2}}dz + O(e^{- \frac{c}{t}}) ,
\end{eqnarray}
\noindent
where $erfc(x) := \frac{2}{\sqrt{\pi}} \int_{x}^{\infty} e^{-y^{2}} dy$. To compute $(III)$ more precisely we introduce the following concepts.
\vspace{0.3 cm}
Let $A$, $B$ be classical pseudodifferential operators and $A$ be an elliptic operator of positive order on a compact manifold.
Then $\operatorname{Tr} \left( Be^{-t A^{2}} \right)$ has an asymptotic expansion of the following type for $t \rightarrow 0^{+}$.
\begin{equation} \label{E:3.26}
\operatorname{Tr} \left( Be^{-t A^{2}} \right) \sim \sum_{\operatorname{Re} \alpha \rightarrow \infty} a_{\alpha, k}(A, B) t^{\alpha} (\log t)^{k}.
\end{equation}
\noindent
When $B$ commutes with $A^{2}$ and vanishes on $\operatorname{ker} A^{2}$,
we define the eta function $\eta(A, B \hspace{0.1 cm} ; \hspace{0.1 cm} s)$ by
\begin{eqnarray} \label{E:3.27}
\eta(A, B \hspace{0.1 cm} ; \hspace{0.1 cm} s) & := &
\frac{1}{\Gamma(\frac{s+1}{2})} \int_{0}^{\infty} t^{\frac{s-1}{2}} \operatorname{Tr} \left( Be^{-t A^{2}} \right) dt \nonumber \\
& = & \sum_{|\lambda| \in \operatorname{Spec}(|A|) - \{ 0 \}} \left( \operatorname{Tr}_{\operatorname{ker}(|A| - |\lambda|)} B \right) | \lambda |^{-s-1}.
\end{eqnarray}
\noindent
Then the noncommutative residue $\operatorname{res}$ is defined as follows ([30], [31], [15]).
\begin{equation} \label{E:3.28}
\operatorname{res} (B) := -2 \operatorname{ord} (A) \hspace{0.1 cm} a_{0,1}(A, B) \hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{ord}(A) \hspace{0.1 cm} \operatorname{Res}_{s=-1}
\eta(A, B \hspace{0.1 cm} ; \hspace{0.1 cm} s).
\end{equation}
\noindent
The following is well known ([30], [6]).
\begin{lemma} \label{Lemma:3.7}
If $B$ is a classical pseudodifferential operator on a compact manifold with $B^{2} = B$, then $\operatorname{res} (B) = 0$.
\end{lemma}
We now go back to (\ref{E:3.21}).
We define a function $F_{\theta}(x)$ and its Mellin transform ${\mathcal M}F_{\theta}(s)$
( see [6] for details ) by
\begin{equation} \label{E:3.22}
F_{\theta}(x) = x \int_{0}^{\infty} erfc(z) e^{-2 cos\theta x z - x^{2}} dz, \qquad
{\mathcal M}F_{\theta}(s) = \int_{0}^{\infty} x^{s-1} F_{\theta}(x) dx.
\end{equation}
\noindent
Using the inverse Mellin transform, we have
\begin{eqnarray} \label{E:3.23}
(III) & = & - i cos\theta \sum_{0 \neq \lambda \in \operatorname{Spec}(|{\mathcal A}|)} d(\lambda) \hspace{0.1 cm} F_{\theta}(\sqrt{t} \lambda) \hspace{0.1 cm}
+ \hspace{0.1 cm} O(e^{- \frac{c}{t}}) \nonumber \\
& = & - i cos\theta \sum_{0 \neq \lambda \in \operatorname{Spec}(|{\mathcal A}|)} d(\lambda) \hspace{0.1 cm} \frac{1}{2 \pi i} \int_{\operatorname{Re} w = c \gg 0} (\sqrt{t} \lambda)^{- w}
\hspace{0.1 cm} {\mathcal M} F_{\theta}(w) \hspace{0.1 cm} dw \hspace{0.1 cm} + \hspace{0.1 cm} O(e^{- \frac{c}{t}}) \nonumber \\
& = & \frac{- i cos\theta}{2 \pi i} \int_{\operatorname{Re} w = c \gg 0} t^{- \frac{w}{2}} \sum_{0 \neq \lambda \in \operatorname{Spec}(|{\mathcal A}|)} d(\lambda)
\hspace{0.1 cm}
\lambda^{- w} \hspace{0.1 cm} {\mathcal M} F_{\theta}(w) \hspace{0.1 cm} dw \hspace{0.1 cm} + \hspace{0.1 cm} O(e^{- \frac{c}{t}}) \nonumber \\
& = & \frac{- i cos\theta}{2 \pi i} \int_{\operatorname{Re} w = c \gg 0} t^{- \frac{w}{2}} \hspace{0.1 cm}
\eta \left( {\mathcal A}, {\mathcal \gamma} [ T^{\prime}(\theta), {\mathcal A} ] {\widetilde P} (\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} w-1 \right)
\hspace{0.1 cm} {\mathcal M} F_{\theta}(w) \hspace{0.1 cm} dw \hspace{0.1 cm} + \hspace{0.1 cm} O(e^{- \frac{c}{t}}) \nonumber \\
& = & - i cos\theta \cdot \operatorname{Res}_{\operatorname{Re} w < c} \left( t^{- \frac{w}{2}} \hspace{0.1 cm}
\eta \left( {\mathcal A}, {\mathcal \gamma} [ T^{\prime}(\theta), {\mathcal A} ] {\widetilde P} (\theta)
\hspace{0.1 cm} ; \hspace{0.1 cm} w-1 \right) \hspace{0.1 cm} {\mathcal M} F_{\theta}(w) \right)
\hspace{0.1 cm} + \hspace{0.1 cm} O(e^{- \frac{c}{t}}) .
\end{eqnarray}
\noindent
Equations (\ref{E:3.17}), (\ref{E:3.18}) and (\ref{E:3.23}) show that
\begin{eqnarray} \label{E:3.24}
& & \operatorname{Tr} \left( - i \phi (x) {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] e^{-t \mathcal{B}^{2}_{{\widetilde P}(\theta)}} \right) =
\frac{i}{2} \operatorname{Tr} \left( {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}]
{\widetilde P}(\theta) e^{-t {\mathcal A}^{2}} \right) \nonumber \\
& - & i cos\theta \cdot \operatorname{Res}_{\operatorname{Re} w < c} \left( t^{- \frac{w}{2}}
\eta \left( {\mathcal A}, {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm}
{\mathcal A}] {\widetilde P} (\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} w-1 \right) {\mathcal M} F_{\theta}(w) \right) +
O(e^{- \frac{c}{t}}) .
\end{eqnarray}
\noindent
Equations (\ref{E:3.13}), (\ref{E:3.16}) and (\ref{E:3.24}) lead to the following lemma.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.4}
\begin{eqnarray*}
\operatorname{Tr} \left( {\dot \mathcal{B}(\theta)} e^{-t \mathcal{B}(\theta)^{2}} \right) & = &
\frac{-i}{\sqrt{4 \pi t}} \operatorname{Tr} \left( {\mathcal \gamma} T^{\prime}(\theta) e^{-t {\mathcal A}^{2}} \right)
+ \frac{i}{2} \operatorname{Tr} \left( {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P}(\theta) e^{-t {\mathcal A}^{2}} \right) \nonumber \\
& - & i cos\theta \cdot \operatorname{Res}_{\operatorname{Re} w < c} \left( t^{- \frac{w}{2}}
\eta \left({\mathcal A}, {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P} (\theta) ; w-1 \right)
{\mathcal M} F_{\theta}(w) \right) + O(e^{- \frac{c}{t}}) .
\end{eqnarray*}
\end{lemma}
\vspace{0.2 cm}
It is known that (\ref{E:3.9}) has at most a simple pole at $s=0$ (Theorem 3.4 in [6]) and has regular values at
$s=0$ for $\theta = 0$ and $\frac{\pi}{2}$ (for the case of $\theta = \frac{\pi}{2}$, see [14]).
Moreover, ${\mathcal M} F_{\theta}(w)$ has only simple poles at negative integers (Lemma 3.3 in [6]).
The following lemma is due to [13] (cf. [6]).
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.5}
Let $A$ and $B$ be classical pseudodifferential operators of order $a$ and $b$, respectively, on a compact manifold $M$ with $\operatorname{dim} M = m$.
If $A$ is a self-adjoint elliptic operator of positive order, then for $t \rightarrow 0^{+}$,
$$
\operatorname{Tr} \left( B e^{-t A^{2}} \right) \sim \sum_{j=0}^{\infty} a_{j}(A, B) t^{\frac{j-m-b}{2a}} + \sum_{j=0}^{\infty} \left( b_{j}(A, B) \log t +
c_{j}(A, B) \right) t^{j}.
$$
\end{lemma}
\vspace{0.2 cm}
\noindent
The equation (\ref{E:3.156}) with
Lemma \ref{Lemma:3.4} and Lemma \ref{Lemma:3.5} (cf. Theorem 3.4 and 3.5 in [6]) implies that
\begin{equation} \label{E:3.25}
\frac{d}{d \theta} \operatorname{Res}_{s=0} \eta_{\mathcal{B}(\theta)}(s) \hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Res}_{s=0} \left( \frac{d}{d \theta} \eta_{\mathcal{B}(\theta)}(s) \right)
\hspace{0.1 cm} = \hspace{0.1 cm}
\frac{4}{\sqrt{\pi}} \hspace{0.1 cm} a_{- \frac{1}{2}, 1} (\mathcal{B}(\theta), {\dot B}(\theta))
\hspace{0.1 cm} = \hspace{0.1 cm}
\frac{1}{\pi} \operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) \right),
\end{equation}
\vspace{0.1 cm}
\noindent
where $a_{- \frac{1}{2}, 1} (\mathcal{B}(\theta), {\dot B}(\theta))$ is the coefficient of $t^{- \frac{1}{2}} \log t$ in the asymptotic expansion of
$\operatorname{Tr} \left( {\dot \mathcal{B}(\theta)} e^{-t \mathcal{B}(\theta)^{2}} \right) $ for $t \rightarrow 0^{+}$.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.6}
\begin{eqnarray*}
\operatorname{Tr} \left( i \hspace{0.1 cm} {\mathcal \gamma} T^{\prime}(\theta) \hspace{0.1 cm} e^{-t {\mathcal A}^{2}} \right) \hspace{0.1 cm} = \hspace{0.1 cm} 0.
\end{eqnarray*}
\noindent
Hence,
$\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) \right) = 0 $
and $\eta_{\mathcal{B}(\theta)}(s)$ has a regular value at $s=0$ for each $0 \leq \theta \leq \frac{\pi}{2}$.
\end{lemma}
\begin{proof}
We recall that $T^{\prime}(\theta) = - i \left( \begin{array}{clcr} U_{{\mathcal P}_{-}}^{\ast} U_{\Pi_{>}} & 0 \\ 0 & 0 \end{array} \right)
{\mathcal P}_{\ast}$.
Using (3) in Lemma \ref{Lemma:3.1}, we have
$$
\operatorname{Tr} \left( i \hspace{0.1 cm} {\mathcal \gamma} T^{\prime}(\theta) \hspace{0.1 cm} e^{-t {\mathcal A}^{2}} \right) =
\operatorname{Tr} \left( {\mathcal \gamma} \left( \begin{array}{clcr} U_{{\mathcal P}_{-}}^{\ast} U_{\Pi_{>}} & 0 \\ 0 & 0 \end{array} \right)
{\mathcal P}_{\ast} e^{- t {\mathcal A}^{2}}
\right)
\hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Tr} \left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & U_{{\mathcal P}_{-}}^{\ast} \\ 0 & 0 \end{array} \right)
\left( \begin{array}{clcr} 0 & 0 \\ U_{\Pi_{>}} & 0 \end{array} \right) {\mathcal P}_{\ast} e^{- t {\mathcal A}^{2}} \right)
$$
$$
=
\operatorname{Tr} \left( - {\mathcal \gamma} \left( \begin{array}{clcr} 0 & 0 \\ U_{\Pi_{>}} & 0 \end{array} \right)
\left( \begin{array}{clcr} 0 & U_{{\mathcal P}_{-}}^{\ast} \\ 0 & 0 \end{array} \right)
{\mathcal P}_{\ast} e^{- t {\mathcal A}^{2}} \right)
\hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Tr} \left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & 0 \\ 0 & U_{{\mathcal P}_{-}} U_{\Pi_{>}}^{\ast} \end{array} \right)
{\mathcal P}_{\ast} e^{- t {\mathcal A}^{2}}
\right) .
$$
\vspace{0.2 cm}
\noindent
Since $\Gamma^{Y}$ anticommutes with $\left( \left( \mathcal{B}_{Y}^{2} \right)^{-} - \left( \mathcal{B}_{Y}^{2} \right)^{+} \right)$,
we have, by (\ref{E:1.9}) and (\ref{E:3.54}),
\begin{eqnarray*}
& & \operatorname{Tr} \left( i \hspace{0.1 cm} {\mathcal \gamma} T^{\prime}(\theta) \hspace{0.1 cm} e^{-t {\mathcal A}^{2}} \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\frac{1}{2} \operatorname{Tr} \left( {\mathcal \gamma}
\left( \begin{array}{clcr} U_{{\mathcal P}_{-}}^{\ast} U_{\Pi_{>}} & 0 \\ 0 & U_{{\mathcal P}_{-}} U_{\Pi_{>}}^{\ast} \end{array} \right)
{\mathcal P}_{\ast} e^{- t {\mathcal A}^{2}} \right) \\
& = &
\frac{1}{2} \operatorname{Tr} \left\{ i \beta \left( \left( \mathcal{B}_{Y}^{2} \right)^{-1} \left( \left( \mathcal{B}_{Y}^{2} \right)^{-} - \left( \mathcal{B}_{Y}^{2} \right)^{+} \right) \right)
\left( \mathcal{B}_{Y}^{2} \right)^{-\frac{1}{2}} \left( \nabla^{Y} \Gamma^{Y} + \Gamma^{Y} \nabla^{Y} \right) \left( \begin{array}{clcr} 0 & -1 \\ -1 & 0 \end{array}
\right) {\mathcal P}_{\ast} e^{- t {\mathcal A}^{2}} \right\}
\hspace{0.1 cm} = \hspace{0.1 cm} 0 ,
\end{eqnarray*}
\noindent
which completes the proof of the lemma.
\end{proof}
\vspace{0.2 cm}
\noindent
Since ${\mathcal M} F_{\theta}(w)$ has a regular value at $w=1$,
Lemma \ref{Lemma:3.4} and (\ref{E:3.156}) imply that
\begin{eqnarray} \label{E:3.29}
& & \frac{d}{d \theta} \eta_{\mathcal{B} (\theta)}(0 \hspace{0.1 cm} ; c(\theta) ) \hspace{0.1 cm} = \hspace{0.1 cm}
- \frac{2}{\sqrt{\pi}} \hspace{0.1 cm} a_{- \frac{1}{2}, 0}(\mathcal{B}(\theta), \hspace{0.1 cm} {\dot \mathcal{B}}(\theta)) \\
& = & - \frac{1}{\sqrt{\pi}} \hspace{0.1 cm} a_{- \frac{1}{2}, 0}
\left( {\mathcal A}, \hspace{0.1 cm} i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P}(\theta) \right)
+ \hspace{0.1 cm} \frac{2}{\sqrt{\pi}} \hspace{0.1 cm} cos\theta \cdot {\mathcal M}F_{\theta}(1) \operatorname{Res}_{w=1}
\left( \eta \left( {\mathcal A}, \hspace{0.1 cm}
i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P}(\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} w-1 \right) \right). \nonumber
\end{eqnarray}
\noindent
We note that
\begin{eqnarray} \label{E:3.30}
a_{- \frac{1}{2}, 0} \left( {\mathcal A}, \hspace{0.1 cm} i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P}(\theta) \right)
& = & \frac{\sqrt{\pi}}{2} \operatorname{Res}_{s=0}
\left( \eta \left( {\mathcal A}, \hspace{0.1 cm} i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}]
{\widetilde P}(\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} s \right) \right) \nonumber \\
& = & \frac{\sqrt{\pi}}{2} \operatorname{Res}_{s=0}
\left( \eta \left( {\mathcal A}, \hspace{0.1 cm} i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} ( sign {\mathcal A} )]
{\widetilde P}(\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} s-1 \right) \right) \nonumber \\
& = & \frac{\sqrt{\pi}}{2} \operatorname{res} \left( i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} ( sign {\mathcal A} )] {\widetilde P}(\theta) \right).
\end{eqnarray}
\noindent
Similarly,
\begin{eqnarray} \label{E:3.31}
\operatorname{Res}_{w=1} \left( \eta({\mathcal A}, \hspace{0.1 cm}
i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} {\mathcal A}] {\widetilde P}(\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} w-1 ) \right)
& = & \operatorname{Res}_{w=1}
\left( \eta \left( {\mathcal A}, \hspace{0.1 cm} i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} ( sign {\mathcal A} )]
{\widetilde P}(\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} w-2 \right) \right) \nonumber \\
& = & \operatorname{Res}_{w=-1}
\left( \eta \left( {\mathcal A}, \hspace{0.1 cm} i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} ( sign {\mathcal A} )]
{\widetilde P}(\theta) \hspace{0.1 cm} ; \hspace{0.1 cm} w \right) \right) \nonumber \\
& = & \operatorname{res} \left( i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} ( sign {\mathcal A} )] {\widetilde P}(\theta) \right).
\end{eqnarray}
\vspace{0.2 cm}
\noindent
The following lemma is straightforward.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.10}
\begin{eqnarray*}
T^{\prime}(\theta) {\widetilde P}(\theta) \hspace{0.1 cm} = \hspace{0.1 cm}
\left( I - {\widetilde P}(\theta) \right) T^{\prime}(\theta) -
\frac{i}{2} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast},
\end{eqnarray*}
where ${\mathcal W} := - U_{\Pi_{>}} sin\theta + U_{{\mathcal P}_{-}} cos\theta$.
\end{lemma}
\vspace{0.2 cm}
\noindent
Equations (\ref{E:3.29}), (\ref{E:3.30}), (\ref{E:3.31}) and Lemma \ref{Lemma:3.10} lead to the following result.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:3.9}
\begin{eqnarray*}
\frac{d}{d \theta} \eta_{\mathcal{B} (\theta)} (0 \hspace{0.1 cm} ; c(\theta) ) & = &
\left( - \frac{1}{2} + \frac{2 cos\theta}{\sqrt{\pi}} \cdot {\mathcal M}F_{\theta}(1) \right)
\operatorname{res} \left( i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} ( sign {\mathcal A} )] {\widetilde P}(\theta) \right)
\hspace{0.1 cm} = \hspace{0.1 cm} 0.
\end{eqnarray*}
\end{lemma}
\begin{proof}
We note that
$$
\operatorname{res} \left( i {\mathcal \gamma} [T^{\prime}(\theta), \hspace{0.1 cm} (sign {\mathcal A})] {\widetilde P}(\theta) \right) =
\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) (sign {\mathcal A}) {\widetilde P}(\theta) \right) -
\operatorname{res} \left( i {\mathcal \gamma} (sign {\mathcal A}) T^{\prime}(\theta) {\widetilde P}(\theta) \right).
$$
We are going to show that $\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) (sign {\mathcal A}) {\widetilde P}(\theta) \right) = 0$ and
$\operatorname{res} \left( i {\mathcal \gamma} (sign {\mathcal A}) T^{\prime}(\theta) {\widetilde P}(\theta) \right) = 0$ can be shown in the same way.
Since $\operatorname{res}$ is a trace,
\begin{eqnarray} \label{E:3.32}
& & \operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) (sign {\mathcal A}) {\widetilde P}(\theta) \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{res} \left( i {\widetilde P}(\theta) {\mathcal \gamma} T^{\prime}(\theta) (sign {\mathcal A}) {\widetilde P}(\theta) \right) \nonumber \\
& = & \operatorname{res} \left( i {\mathcal \gamma} (I - {\widetilde P}(\theta))
T^{\prime}(\theta) (sign {\mathcal A}) {\widetilde P}(\theta) \right) \nonumber \\
& = & \operatorname{res} \left( i {\mathcal \gamma} \left( T^{\prime}(\theta) {\widetilde P}(\theta) +
\frac{i}{2} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast} \right)
(sign {\mathcal A}) {\widetilde P}(\theta) \right) \nonumber \\
& = & \operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) {\widetilde P}(\theta)
(sign {\mathcal A}) {\widetilde P}(\theta) \right) - \frac{1}{2} \operatorname{res}
\left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast}
(sign {\mathcal A}) {\widetilde P}(\theta) \right) \nonumber \\
& = & cos\theta \operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) {\widetilde P}(\theta) \right) - \frac{1}{2} \operatorname{res}
\left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast}
(sign {\mathcal A}) {\widetilde P}(\theta) \right). \nonumber
\end{eqnarray}
\noindent
We note that
\begin{eqnarray*}
\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) {\widetilde P}(\theta) \right) =
\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) ( I - {\widetilde P}(\theta)) \right) =
\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) \right) -
\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) {\widetilde P}(\theta) \right),
\end{eqnarray*}
\noindent
which together with Lemma \ref{Lemma:3.6} shows that
\begin{equation} \label{E:3.34}
\operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) {\widetilde P}(\theta) \right) =
\frac{1}{2} \operatorname{res} \left( i {\mathcal \gamma} T^{\prime}(\theta) \right) = 0.
\end{equation}
\vspace{0.2 cm}
We note that
$(sign {\mathcal A}) \hspace{0.1 cm} = \hspace{0.1 cm} \left( \begin{array}{clcr} 0 & U_{\Pi_{>}}^{\ast} \\ U_{\Pi_{>}} & 0 \end{array} \right)
{\mathcal P}_{\ast}$
and ${\mathcal \gamma}$ anticommutes with $\hspace{0.1 cm} (sign {\mathcal A}) \hspace{0.1 cm} $ and
$ \hspace{0.1 cm} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) \hspace{0.1 cm}$.
Hence, we have
\begin{eqnarray*}
& & \operatorname{res} \left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast}
(sign {\mathcal A}) {\widetilde P}(\theta) \right) \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{res} \left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast}
(sign {\mathcal A}) ( I - {\widetilde P}(\theta)) \right),
\end{eqnarray*}
\noindent
which shows that
\begin{eqnarray} \label{E:3.52}
& & \operatorname{res} \left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast}
(sign {\mathcal A}) {\widetilde P}(\theta) \right) \nonumber \\
& = & \frac{1}{2} \operatorname{res} \left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right)
{\mathcal P}_{\ast} (sign {\mathcal A}) \right)
\hspace{0.1 cm} = \hspace{0.1 cm}
\frac{1}{2} \operatorname{res} \left( \left( \begin{array}{clcr} i {\mathcal W}^{\ast} U_{\Pi_{>}} & 0 \\ 0 & i {\mathcal W} U_{\Pi_{>}}^{\ast} \end{array} \right)
{\mathcal P}_{\ast} \right) \nonumber \\
& = & - \frac{1}{2} \operatorname{res} \left( {\mathcal \gamma} (sign {\mathcal A})
\left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast} \right)
\hspace{0.1 cm} = \hspace{0.1 cm}
\frac{1}{2} \operatorname{res} \left( \left( \begin{array}{clcr} i U_{\Pi_{>}}^{\ast} {\mathcal W} & 0 \\ 0 & i U_{\Pi_{>}} {\mathcal W}^{\ast} \end{array} \right)
{\mathcal P}_{\ast} \right).
\end{eqnarray}
\noindent
The above equality with Lemma \ref{Lemma:3.1} and Lemma \ref{Lemma:3.7} shows that
\begin{eqnarray} \label{E:3.53}
\operatorname{res} \left( {\mathcal \gamma} \left( \begin{array}{clcr} 0 & {\mathcal W}^{\ast} \\ - {\mathcal W} & 0 \end{array} \right) {\mathcal P}_{\ast}
(sign {\mathcal A}) {\widetilde P}(\theta) \right)
& = & \frac{i}{4} \operatorname{res} \left( \left( \begin{array}{clcr} {\mathcal W}^{\ast} U_{\Pi_{>}} + U_{\Pi_{>}}^{\ast} {\mathcal W} & 0 \\
0 & {\mathcal W} U_{\Pi_{>}}^{\ast} + U_{\Pi_{>}} {\mathcal W}^{\ast} \end{array} \right) {\mathcal P}_{\ast} \right) \nonumber \\
& = & - \frac{i sin \theta}{2} \operatorname{res} \left( {\mathcal P}_{\ast} \right) \hspace{0.1 cm} = \hspace{0.1 cm} 0,
\end{eqnarray}
\noindent
which completes the proof of the lemma.
\end{proof}
\vspace{0.2 cm}
For one parameter family of essentially self-adjoint Dirac operators $\mathcal{B}_{{\widetilde P}(\theta)}$ ($ 0 \leq \theta \leq \frac{\pi}{2}$)
we define the spectral flow $\operatorname{SF} (\mathcal{B}_{{\widetilde P}(\theta)})_{\theta \in [0, \frac{\pi}{2}]}$ by
$$
\operatorname{SF} (\mathcal{B}_{{\widetilde P}(\theta)})_{\theta \in [0, \frac{\pi}{2}]} \hspace{0.1 cm} := \hspace{0.1 cm} m^{+} - m^{-},
$$
where $m^{+}$ ($m^{-}$) is the number of eigenvalues which start negative (non-negative) and end non-negative (negative).
The following formula is well known (cf. Lemma 3.4 in [17]).
\vspace{0.2 cm}
\begin{equation} \label{E:3.4557}
\eta( \mathcal{B}_{{\mathcal P}_{-, {\mathcal L}_{0}}} ) - \eta ( \mathcal{B}_{\Pi_{>, {\mathcal L}_{0}}} ) \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{SF} (\mathcal{B}_{{\widetilde P}(\theta)})_{\theta \in [0, \frac{\pi}{2}]} \hspace{0.1 cm} + \hspace{0.1 cm}
\int_{0}^{\frac{\pi}{2}} \frac{d}{d \theta} \eta_{\mathcal{B}_{{\widetilde P}(\theta)}}(0) \hspace{0.1 cm} d \theta.
\end{equation}
\vspace{0.2 cm}
\noindent
Lemma \ref{Lemma:3.9} and the result of Nicolaescu (Theorem 7.5 in [17], [23]) show that
\vspace{0.2 cm}
\begin{equation} \label{E:3.4558}
\eta( \mathcal{B}_{{\mathcal P}_{-, {\mathcal L}_{0}}} ) - \eta ( \mathcal{B}_{\Pi_{>, {\mathcal L}_{0}}} ) \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{SF} (\mathcal{B}_{{\widetilde P}(\theta)})_{\theta \in [0, \frac{\pi}{2}]} \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M})_{\theta \in [0, \frac{\pi}{2}]} ,
\end{equation}
\vspace{0.2 cm}
\noindent
where ${\mathcal C}_{M}$ is the Calder\'on projector for $\mathcal{B}$ on $M$
and $\operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M})_{\theta \in [0, \frac{\pi}{2}]}$
is the Maslov index for the path ${\widetilde P}(\theta)$ and the constant path ${\mathcal C}_{M}$.
We refer to [17] and [23] for the definitions of the Maslov index and Calder\'on projector.
\vspace{0.2 cm}
The unitary operators corresponding to the projection ${\mathcal P}_{+}$ is $- \hspace{0.1 cm} U_{{\mathcal P}_{-}}$,
which shows that for $0 \leq \theta \leq \frac{\pi}{2}$
$$
{\widetilde P}(- \theta) \hspace{0.1 cm} = \hspace{0.1 cm}
\frac{1}{2} \left( \begin{array}{clcr} \operatorname{Id} & P(- \theta)^{\ast} \\ P(- \theta) & \operatorname{Id} \end{array} \right) {\mathcal P}_{\ast}
\hspace{0.1 cm} + \hspace{0.1 cm} {\mathcal P}_{{\mathcal L}_{0}}
$$
\vspace{0.2 cm}
\noindent
is a smooth path connecting $\Pi_{>, {\mathcal L}_{0}}$ and ${\mathcal P}_{+, {\mathcal L}_{0}}$.
Similar computation shows that
\begin{equation} \label{E:3.4559}
\eta( \mathcal{B}_{{\mathcal P}_{+, {\mathcal L}_{0}}} ) - \eta ( \mathcal{B}_{\Pi_{>, {\mathcal L}_{0}}} ) \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{SF} (\mathcal{B}_{{\widetilde P}(- \theta)})_{\theta \in [0, \frac{\pi}{2}]} \hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Mas} ( {\widetilde P}(- \theta), \hspace{0.1 cm} {\mathcal C}_{M})_{\theta \in [0, \frac{\pi}{2}]} .
\end{equation}
\noindent
Summarizing the above arguments we have the following theorem, which is the main result of this section.
\vspace{0.2 cm}
\begin{theorem} \label{Theorem:3.13}
Let $(M, g^{M})$ be a compact Riemannian manifold with boundary $Y$ and $g^{M}$ be a product metric near $Y$.
Then :
\vspace{0.2 cm}
\noindent
(1) $\hspace{0.2 cm} \eta( \mathcal{B}_{{\mathcal P}_{-, {\mathcal L}_{0}}} ) - \eta ( \mathcal{B}_{\Pi_{>, {\mathcal L}_{0}}} )
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{SF} (\mathcal{B}_{{\widetilde P}(\theta)})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M})_{\theta \in [0, \frac{\pi}{2}]} $. \newline
(2) $\hspace{0.2 cm} \eta( \mathcal{B}_{{\mathcal P}_{+, {\mathcal L}_{0}}} ) - \eta ( \mathcal{B}_{\Pi_{>, {\mathcal L}_{0}}} )
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{SF} (\mathcal{B}_{{\widetilde P}(- \theta)})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Mas} ( {\widetilde P}(- \theta), \hspace{0.1 cm} {\mathcal C}_{M})_{\theta \in [0, \frac{\pi}{2}]} $. \newline
\end{theorem}
\vspace{0.3 cm}
\section{Gluing formula of the refined analytic torsion}
\vspace{0.2 cm}
The gluing formula of the analytic torsion with respect to the relative and absolute boundary conditions ([9], [21], [29]) and the gluing formula of the eta invariant with respect to the APS boundary condition ([6], [7], [17], [32], [33]) are well known.
In this section we are going to use Theorem \ref{Theorem:2.11} and Theorem \ref{Theorem:3.13} together with results in [9], [6] and [17] to
obtain the gluing formula of the refined analytic torsion when $\nabla$ is an acyclic Hermitian connection.
Let $({\widehat M}, g^{\widehat M})$ be a closed Riemannian manifold of dimension $m = 2r -1$ and
${\widehat E} \rightarrow {\widehat M}$ be a flat vector bundle with a flat connection $\nabla$.
We denote by $Y$ a hypersurface of ${\widehat M}$ such that ${\widehat M}-Y$ has two components
whose closures are denoted by $M_{1}$ and $M_{2}$, {\it i.e.}
${\widehat M} = M_{1} \cup_{Y} M_{2}$.
We assume that $g^{\widehat M}$ is a product metric near $Y$ and that $\nabla$ is a Hermitian connection.
Let $\partial u$ be the unit normal vector field on a collar neighborhood of $Y$ such that $\partial u$ is outward on $M_{1}$ and inward on $M_{2}$.
We denote by $\mathcal{B}^{{\widehat M}}$ the odd signature operator on ${\widehat M}$ and
denote by $\mathcal{B}^{M_{1}}$, $\mathcal{B}^{M_{2}}$ ($E_{1}$, $E_{2}$, $g^{M_{1}}$, $g^{M_{2}}$) the restriction of
$\mathcal{B}^{{\widehat M}}$ (${\widehat E}$, $g^{{\widehat M}}$) to $M_{1}$, $M_{2}$.
We impose the boundary condition ${\mathcal P}_{+, {\mathcal L}_{1}}$ on $M_{1}$ and ${\mathcal P}_{-, {\mathcal L}_{0}}$ on $M_{2}$.
Then (\ref{E:1.17}) and (\ref{E:1.18}) show that
\begin{eqnarray} \label{E:4.1}
& & \log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}^{M_{1}}_{\operatorname{even}, {\mathcal P}_{+, {\mathcal L}_{1}}}) +
\log \operatorname{Det}_{\operatorname{gr}, \theta} (\mathcal{B}^{M_{2}}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}}) \\
& = & \frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \left( \log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{1}}_{q, {\widetilde {\mathcal P}_{1}}})^{2} +
\log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{2}}_{q, {\widetilde {\mathcal P}_{0}}})^{2} \right)
- i \pi \hspace{0.1 cm} \left( \eta(\mathcal{B}^{M_{1}}_{\operatorname{even}, {\mathcal P}_{+, {\mathcal L}_{1}}}) +
\hspace{0.1 cm} \eta(\mathcal{B}^{M_{2}}_{\operatorname{even}, {\mathcal P}_{-, {\mathcal L}_{0}}}) \right). \nonumber
\end{eqnarray}
\vspace{0.2 cm}
\noindent
Theorem \ref{Theorem:2.11} together with Theorem 4.3 in [9] (p.36 in [9], cf. [21], [29]) leads to the following result.
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:4.1}
We assume that for each $0 \leq q \leq m$, $i = 1, 2$, $H^{q}({\widehat M}, {\widehat E}) = H^{q}(M_{i}, Y; E_{i}) = H^{q}(M_{i};E_{i}) = 0$. Then,
\begin{eqnarray*}
& & \frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \left( \log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{1}}_{q, {\widetilde {\mathcal P}_{1}}})^{2} +
\log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{2}}_{q, {\widetilde {\mathcal P}_{0}}})^{2} \right) \\
& = & \frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \left( \log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{1}}_{q, \operatorname{abs}})^{2} +
\log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{2}}_{q, \operatorname{rel}})^{2} \right)
= \frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2\theta} (\mathcal{B}^{{\widehat M}}_{q})^{2}.
\end{eqnarray*}
\end{lemma}
\vspace{0.3 cm}
Under the assumption in Lemma \ref{Lemma:4.1}
Theorem \ref{Theorem:3.13} shows that
\begin{equation} \label{E:4.2}
\eta ({\mathcal{B}^{M_{2}}_{{\mathcal P}_{-}}}) \hspace{0.1 cm} - \hspace{0.1 cm} \eta ({\mathcal{B}^{M_{2}}_{\Pi_{>}}})
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2}})_{\theta \in [0, \frac{\pi}{2}]}.
\end{equation}
\vspace{0.2 cm}
\noindent
We next consider $\eta ({\mathcal{B}^{M_{1}}_{{\mathcal P}_{+}}}) - \eta ({\mathcal{B}^{M_{1}}_{\Pi_{<}}})$.
Since $\partial u$ is the outward normal derivative on a collar neighborhood of $Y$ on $M_{1}$,
to use Theorem \ref{Theorem:3.13} we rewrite the odd signature operator $\mathcal{B}^{M_{1}}$ near the boundary by
$$
\mathcal{B}^{M_{1}} = {\mathcal \gamma} (\partial u + {\mathcal A}) = - {\mathcal \gamma} ( - \partial u - {\mathcal A}).
$$
\noindent
Here $- \partial u$ is the inward normal derivative to $Y$ on $M_{1}$.
Since $\Pi_{<}({\mathcal A}) = \Pi_{>}(- {\mathcal A})$ and $I - {\widetilde P}(\theta)$ is a path connecting
$\Pi_{<}({\mathcal A})$ and ${\mathcal P}_{+}$, Theorem \ref{Theorem:3.13} shows that
\begin{eqnarray} \label{E:4.3}
\eta ({\mathcal{B}^{M_{1}}_{{\mathcal P}_{+}}}) \hspace{0.1 cm} - \hspace{0.1 cm} \eta ({\mathcal{B}^{M_{1}}_{\Pi_{<}({\mathcal A})}})
& = & \eta ({\mathcal{B}^{M_{1}}_{{\mathcal P}_{+}}}) \hspace{0.1 cm} - \hspace{0.1 cm} \eta ({\mathcal{B}^{M_{1}}_{\Pi_{>}(- {\mathcal A})}})
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1}})_{\theta \in [0, \frac{\pi}{2}]}.
\end{eqnarray}
\noindent
Equations (\ref{E:4.2}) and (\ref{E:4.3}) together with Theorem 8.8 in [17] show that
\begin{eqnarray} \label{E:4.4}
& & \eta ({\mathcal{B}^{M_{1}}_{{\mathcal P}_{-}}}) \hspace{0.1 cm} + \hspace{0.1 cm} \eta ({\mathcal{B}^{M_{2}}_{{\mathcal P}_{+}}}) \nonumber \\
& = & \eta ({\mathcal{B}^{M_{1}}_{\Pi_{<}}}) \hspace{0.1 cm} + \hspace{0.1 cm} \eta ({\mathcal{B}^{M_{2}}_{\Pi_{>}}})
\hspace{0.1 cm} + \hspace{0.1 cm} \operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} + \hspace{0.1 cm} \operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1}})_{\theta \in [0, \frac{\pi}{2}]} \nonumber \\
& = & \eta ({\mathcal{B}^{{\widehat M}}})
\hspace{0.1 cm} + \hspace{0.1 cm} \operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} + \hspace{0.1 cm} \operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1}})_{\theta \in [0, \frac{\pi}{2}]} .
\end{eqnarray}
\vspace{0.2 cm}
\begin{lemma} \label{Lemma:4.2}
Under the assumption of Lemma \ref{Lemma:4.1} we have :
\begin{eqnarray*} \label{E:4.444}
\operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1}})_{\theta \in [0, \frac{\pi}{2}]} .
\end{eqnarray*}
\noindent
In particular,
$\hspace{0.1 cm} \eta ({\mathcal{B}^{M_{1}}_{{\mathcal P}_{-}}}) \hspace{0.1 cm} + \hspace{0.1 cm} \eta ({\mathcal{B}^{M_{2}}_{{\mathcal P}_{+}}})
\hspace{0.1 cm} \equiv \hspace{0.1 cm}
\eta ({\mathcal{B}^{{\widehat M}}}) \qquad (\operatorname{mod} 2 \hspace{0.1 cm} {\Bbb Z}). $
\end{lemma}
\begin{proof}
We put $M_{1, r} = M_{1} \cup_{Y} \left( [0, r] \times Y \right)$, $M_{2, r} = M_{2} \cup_{Y} \left( [-r, 0] \times Y \right)$ and
$M_{1, \infty} = M_{1} \cup_{Y} \left( [0, \infty) \times Y \right)$, $M_{2, \infty} = M_{2} \cup_{Y} \left( (- \infty, 0] \times Y \right)$.
We denote the extensions of $\mathcal{B}$ to $M_{i, r}$, $M_{i, \infty}$
by $\mathcal{B}_{M_{i, r}}$, $\mathcal{B}_{M_{i, \infty}}$ and
denote the corresponding Calder\'on projectors by ${\mathcal C}_{M_{i, r}}$, and
$\operatorname{Im} {\mathcal C}_{M_{i, r}} := L_{M_{i, r}}$, and $\lim_{r \rightarrow \infty} L_{M_{i, r}} := L_{M_{i, \infty}}$ for $i = 1, 2$.
We also denote the orthogonal projection to $L_{M_{i, \infty}}$ by ${\mathcal C}_{M_{i, \infty}}$.
Under the assumption of Lemma \ref{Lemma:4.1} it is shown in [17] (p.610 in [17]) that
$L_{M_{1, \infty}}$ and $L_{M_{2, \infty}}$ are Lagrangian subspaces
and $L_{M_{2, \infty}} = {\mathcal \gamma} L_{M_{1, \infty}}$.
Hence ${\mathcal C}_{M_{2, \infty}} = - {\mathcal \gamma} {\mathcal C}_{M_{1, \infty}} {\mathcal \gamma}$.
We define a homotopy $(F(\theta, s), \hspace{0.1 cm} G(\theta, s))$ on $M_{2}$ as follows.
$$
F(\theta, s) = {\widetilde P}(\theta), \qquad G(\theta, s) = {\mathcal C}_{M_{2, s}}, \qquad
(0 \leq \theta \leq \frac{\pi}{2}, \quad 0 \leq s \leq \infty).
$$
Then, $(F(\theta, 0), \hspace{0.1 cm} G(\theta, 0)) = ({\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2}})$ and
$(F(\theta, \infty), \hspace{0.1 cm} G(\theta, \infty)) = ({\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2, \infty}})$.
Since $\operatorname{ker} \mathcal{B}_{\Pi_{>}}$ and $\operatorname{ker} \mathcal{B}_{{\mathcal P}_{-}}$ are topological invariants (cf. Lemma \ref{Lemma:1.2} and Proposition 4.9 in [1]),
the assumption implies that
\begin{eqnarray*}
\operatorname{dim} \left( \operatorname{ker} F(0, s) \cap \operatorname{Im} G(0, s) \right) & = &
\operatorname{dim} \left( \operatorname{ker} \Pi_{>}({\mathcal A}) \cap \operatorname{Im} {\mathcal C}_{M_{2, s}} \right) \hspace{0.1 cm} = \hspace{0.1 cm} 0 , \\
\operatorname{dim} \left( \operatorname{ker} F(\frac{\pi}{2}, s) \cap \operatorname{Im} G(\frac{\pi}{2}, s) \right) & = &
\operatorname{dim} \left( \operatorname{ker} {\mathcal P}_{-} \cap \operatorname{Im} {\mathcal C}_{M_{2, s}} \right)
\hspace{0.1 cm} = \hspace{0.1 cm} 0 ,
\end{eqnarray*}
\noindent
which shows (cf. p.587 in [17]) that
$$
\operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2, \infty}})_{\theta \in [0, \frac{\pi}{2}]}.
$$
\noindent
Similarly, we have
$$
\operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm} \operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1, \infty}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Mas} ( - {\mathcal \gamma} {\widetilde P}(\theta) {\mathcal \gamma},
\hspace{0.1 cm} - {\mathcal \gamma} {\mathcal C}_{M_{2, \infty}} {\mathcal \gamma} )_{\theta \in [0, \frac{\pi}{2}]}.
$$
\noindent
Hence, we have (cf. p.586 in [17])
\begin{eqnarray*}
\operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2}})_{\theta \in [0, \frac{\pi}{2}]}
& = & \operatorname{Mas} ( {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{2, \infty}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Mas} ( - {\mathcal \gamma} {\widetilde P}(\theta) {\mathcal \gamma},
\hspace{0.1 cm} - {\mathcal \gamma} {\mathcal C}_{M_{2, \infty}} {\mathcal \gamma})_{\theta \in [0, \frac{\pi}{2}]} \\
& = & \operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1, \infty}})_{\theta \in [0, \frac{\pi}{2}]}
\hspace{0.1 cm} = \hspace{0.1 cm}
\operatorname{Mas} ( I - {\widetilde P}(\theta), \hspace{0.1 cm} {\mathcal C}_{M_{1}})_{\theta \in [0, \frac{\pi}{2}]} ,
\end{eqnarray*}
\noindent
which completes the proof of the lemma.
\end{proof}
\vspace{0.3 cm}
Under the assumption of Lemma \ref{Lemma:4.1} the refined analytic torsion $T_{{\widehat M}}(g^{{\widehat M}}, \nabla)$ on
${\widehat M}$ is defined by (Definition 10.1 in [4])
$$
\log T_{{\widehat M}}(g^{{\widehat M}}, \nabla) \hspace{0.1 cm} = \hspace{0.1 cm} \log \operatorname{Det}_{\operatorname{gr}, \theta} ( \mathcal{B}_{\operatorname{even}}^{{\widehat M}} )
\hspace{0.1 cm} + \hspace{0.1 cm} \pi i \hspace{0.1 cm} (\operatorname{rank}({\widehat E})) \hspace{0.1 cm}
\eta(\mathcal{B}_{\operatorname{even}}^{{\widehat M}, \operatorname{trivial}}).
$$
\vspace{0.2 cm}
\noindent
The refined analytic torsion $T_{M_{1}, {\mathcal P}_{+}}(g^{M_{1}}, \nabla)$ and $T_{M_{2}, {\mathcal P}_{-}}(g^{M_{2}}, \nabla)$ on $M_{1}$, $M_{2}$
with respect to the boundary conditions ${\mathcal P}_{+}$ and ${\mathcal P}_{-}$ are defined similarly (Dfinition 4.9 in [14]).
Lemma \ref{Lemma:4.1} and Lemma \ref{Lemma:4.2} lead to the following theorem, which is the main result of this paper.
\vspace{0.2 cm}
\begin{theorem} \label{Theorem:4.3}
Let $({\widehat M}, g^{\widehat M})$ be a closed Riemannian manifold of dimension $m = 2r -1$ and $Y$ be a hypersurface so that
${\widehat M} = M_{1} \cup_{Y} M_{2}$. We assume that $g^{\widehat M}$ is a product metric near $Y$ and
for each $0 \leq q \leq m$, $i = 1, 2$, $H^{q}({\widehat M}, {\widehat E}) = H^{q}(M_{i}, Y; E_{i}) = H^{q}(M_{i};E_{i}) = 0$. Then,
\begin{eqnarray*}
\log T_{{\widehat M}}(g^{{\widehat M}}, \nabla) & = & \frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \log \operatorname{Det}_{2\theta} (\mathcal{B}^{{\widehat M}}_{q})^{2}
- i \pi \hspace{0.1 cm} \eta(\mathcal{B}^{{\widehat M}}_{\operatorname{even}}) + i \pi (\operatorname{rank}({\widehat E})) \eta(\mathcal{B}_{\operatorname{even}}^{{\widehat M}, \operatorname{trivial}}) \\
& \equiv & \frac{1}{2} \sum_{q=0}^{m} (-1)^{q+1} \cdot q \cdot \left( \log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{1}}_{q, {\widetilde {\mathcal P}_{1}}})^{2} +
\log \operatorname{Det}_{2\theta} (\mathcal{B}^{M_{2}}_{q, {\widetilde {\mathcal P}_{0}}})^{2} \right) \\
& & - i \pi \hspace{0.1 cm} \left( \eta(\mathcal{B}^{M_{1}}_{\operatorname{even}, {\mathcal P}_{+}}) + \eta(\mathcal{B}^{M_{2}}_{\operatorname{even}, {\mathcal P}_{-}}) \right)
+ i \pi (\operatorname{rank}({\widehat E}))
\hspace{0.1 cm} \left( \eta(\mathcal{B}^{M_{1}, \operatorname{trivial}}_{\operatorname{even}, {\mathcal P}_{+}}) + \eta(\mathcal{B}^{M_{2}, \operatorname{trivial}}_{\operatorname{even}, {\mathcal P}_{-}}) \right) \\
& = & \log T_{M_{1}, {\mathcal P}_{+}}(g^{M_{1}}, \nabla) \hspace{0.1 cm} + \hspace{0.1 cm}
\log T_{M_{2}, {\mathcal P}_{-}}(g^{M_{2}}, \nabla) \qquad (\operatorname{mod} \hspace{0.1 cm} 2\pi i {\Bbb Z}).
\end{eqnarray*}
\noindent
Equivalently, we have
$$
T_{{\widehat M}}(g^{{\widehat M}}, \nabla) \hspace{0.1 cm} = \hspace{0.1 cm} T_{M_{1}, {\mathcal P}_{+}}(g^{M_{1}}, \nabla) \cdot
T_{M_{2}, {\mathcal P}_{-}}(g^{M_{2}}, \nabla).
$$
\end{theorem}
\vspace{0.3 cm}
\vspace{0.5 cm}
|
\section{Introduction}
The structure of multipartite entanglement has been a subject of much research in recent years. Much work has been done on trying to classify the entanglement in many body states (for example \cite{DVC3qubit, 4qubit, 2x2xn}), and investigating the properties and uses of particular multipartite entangled states, for example graph states in measurement based quantum computing \cite{HEB04}. Three qubit states with genuine tripartite entanglement fall into two SLOCC\footnote{SLOCC: Stochastic Local Operations and Classical Communication.}-classes: one inhabited by the GHZ-state, which is a graph states, and one inhabited by the W-state \cite{DVC3qubit}.
But beyond these states, there is little structural understanding of general multipartite entangled states; even their classification remains mysterious. Such states also remain virtually untapped as resources for quantum information processing protocols.
Most of the work described above has so far employed rather technical arguments using linear algebra. This paper describes some tentative steps towards a different, \emph{compositional} approach, where we view entangled states as being built up from simpler components. This approach springs from the programme, initiated by Abramsky and one of the authors \cite{CatSem}, to analyse quantum mechanics in terms of symmetric monoidal categories. We assume that the reader is familiar with the basics of this approach; for an introductory account see \cite{ContPhys}. One feature of this programme is that it allows us to use a very intuitive graphical language to describe quantum states and processes, and this is utilised in much of what follows.
Here we show how to build up examples from the W SLOCC class using simple graphical building blocks; the GHZ-state is itself one of the building blocks. These building blocks are categorical structures called \emph{basis structures} \cite{wosums}; in the categorical quantum mechanics programme these provide an abstract counterpart to orthonormal bases. In particular we will use a pair of \emph{complementary} basis structures - this notion was introduced by Duncan and one of the authors \cite{RedGreen} to model mutually unbiased bases. Their \em graphical Z/X-calculus \em employed a vivid graphical convention where two complementary basis structures were depicted using green and red dots respectively.
To understand the difference between the GHZ class states and W class states within the context of abstract categorical quantum mechanics, we introduce a concept of the \em supplementarity \em of certain elements relative to complementary basis structures; W states then arise in situations of supplementarity. We sketch how this concept leads to distinct subclasses of more general multipartite qubit states.
One important advantage of the graphical presentation of entangled states is that it is subject to automated reasoning, by means of the {\tt quantomatic} software \cite{quanto}, developed by Dixon, Duncan, Kissinger and Merry.
\begin{figure}
\begin{center}
\epsfig{figure=Quanto,width=320pt}
\caption{Screenshot of some suggested rewrites in {\tt quantomatic}.}
\end{center}
\end{figure}
Meanwhile, the results in this paper have also led to a new graphical calculus, which takes both the GHZ and W state as its primitives \cite{CK}. Rather than being mediated by the laws of complementarity, this graphical GHZ/W-calculus is mediated by the laws of basic arithmetic \cite{CKMS}.
This paper is structured as follows: in section \ref{3qubitbackgroundSec} we give a brief introduction to some key aspects of three-qubit entanglement, and section \ref{redgreenbackgroundSec} reviews some essential properties of complementary basis structures. Next, in section \ref{GHZWgraphicallySec} we show (via concrete linear algebra calculations) how certain states from the GHZ and W classes can be built up from the morphisms of a pair of complementary basis structures.
In section \ref{sec:topoclass} we show how one can identify W class states within the graphical calculus by means of the notion of supplementarity.
We conclude with some speculations on whether our methods can be used for the study of general multipartite entanglement.
\section{Background: 3-qubit entanglement}\label{3qubitbackgroundSec}
Entangled states are classified primarily according to the following criterion: If state $\ket{\psi}$ can be transformed into state $\ket{\phi}$ via local operations on the components of the system (we also allow classical communication between the agents acting on the components, so that they can condition their operations on the outcomes of measurements performed by other agents, for example) then $\ket{\psi}$ is more entangled than, or as entangled as $\ket{\phi}$.
In the case of bipartite systems (with components whose state spaces have arbitrary dimension) the entanglement of quantum states has been completely classified, in that there exists a well-defined mathematical criterion for determining whether one state can be transformed into another via LOCC (local operations and classical communication). This leads to the well-known \emph{majorization order} \cite{NieMaj}.
For systems with more than two subsystems (henceforth we will refer to this as \emph{multipartite} entanglement) no such classification has been achieved. However in certain cases a weaker classification has been achieved. This is based on whether one state can be converted into another via LOCC \emph{with some non-zero probability}. In this case we say that the two states are related via SLOCC (\emph{stochastic} local operations and classical communication).
Mathematically SLOCC translates into a very simple condition \cite{DVC3qubit}. For the $n$-partite state $\ket{\psi}$ to be transformable into $\ket{\phi}$ via SLOCC there must exist local linear operations $A_1,A_2,\dots,A_n$ such that:
\begin{equation} \label{sloccconditionEq}
\ket{\phi} = A_1\otimes A_2 \otimes \dots \otimes A_n \ket{\psi}
\end{equation}
In practice determining whether this condition is satisfied for two states is not straightforward. However in the case of three qubits the states have been completely classified under SLOCC \cite{DVC3qubit}. There are six classes of states arranged in a hierarchy (in fact a partial order). The states of a given class are all interconvertible under SLOCC. States from higher classes can be converted via SLOCC into states from lower classes.
\[
\begin{tikzpicture}[scale=1.1]
\begin{pgfonlayer}{nodelayer}
\node [style=white dot] (0) at (-7, 0.25) {\small GHZ};
\node [style=white dot] (1) at (-5, 0.25) {\small \ W \ };
\node [style=white dot] (2) at (-8, -1) {\small A-BC};
\node [style=white dot] (3) at (-6, -1) {\small B-CA};
\node [style=white dot] (4) at (-4, -1) {\small C-AB};
\node [style=white dot] (5) at (-6, -2.25) {\small $\!$A-B-C$\!$};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (3) to (5);
\draw (1) to (4);
\draw (2) to (5);
\draw (0) to (3);
\draw (4) to (5);
\draw (3) to (1);
\draw (2) to (1);
\draw (2) to (0);
\draw (0) to (4);
\end{pgfonlayer}
\end{tikzpicture}\]
The bottom class contains the completely separable states. The middle three classes consist of states where one system is unentangled, while the other two are entangled with each other. There are two classes of true tripartite entanglement, not interconvertible, each named after a particular member state:
\begin{eqnarray}
|\Psi_{GHZ}\rangle & = & \frac{1}{\sqrt{2}} ( |000\rangle + |111\rangle ) \\
|\Psi_{W}\rangle & = & \frac{1}{\sqrt{3}} ( |011\rangle + |101\rangle + |110\rangle )
\end{eqnarray}
Is there any intuitive difference between the GHZ-class and W-class states? One difference is mostly clearly seen using a measure of entanglement called the \emph{tangle}, introduced in \cite{CKWtangle}. Label the three qubits $A$, $B$ and $C$. Strictly the tangle is defined on mixed states of two qubits. However it is possible to sensibly extend this definition and calculate the tangle, $\tau_{A(BC)}$ for the entanglement between qubit $A$ and qubits $B$ and $C$ viewed as a single 4-dimensional system. The tangle for the entanglement between $A$ and $B$, $\tau_{AB}$ and between $A$ and $C$, $\tau_{AC}$, can also be calculated. The following inequality then always holds:
\begin{equation}
\tau_{A(BC)} \geq \tau_{AB} + \tau_{AC}
\end{equation}
and likewise for permutations of $A$, $B$ and $C$. Interestingly the quantities:
\begin{eqnarray}
\tau_{A(BC)} - \tau_{AB} - \tau_{AC}\\
\tau_{B(AC)} - \tau_{AB} - \tau_{BC}\\
\tau_{C(AB)} - \tau_{BC} - \tau_{AC}
\end{eqnarray}
are all equal. This quantity is termed the \emph{3-tangle} and is denoted $\tau_{ABC}$. We can now write, for example:
\begin{equation}
\tau_{A(BC)} = \tau_{AB} + \tau_{AC} + \tau_{ABC}
\end{equation}
This seems to say that the entanglement between $A$ and the combined system of $B$ and $C$ consists of the pairwise entanglement of $A$ with $B$ and $A$ with $C$, plus some kind of global tripartite entanglement, quantified by the 3-tangle.
It can be shown \cite{DVC3qubit} that all GHZ-class states have non-zero 3-tangle, while all W states have zero 3-tangle. Thus it would seem that in some sense, the entanglement in W-class states is all pairwise, or local, while in the case of GHZ-class states, at least some of the entanglement is genuinely global, shared between all three qubits.
\section{Background: Red-Green calculus}\label{redgreenbackgroundSec}
We consider as given a dagger symmetric monoidal category ($\dag$-SMC) \cite{CatSem}, that is, a symmetric monoidal category with an identity-on-objects involutive contravariant functor $f\mapsto f^\dagger$. We will work within its corresponding graphical calculus \cite{JoyalStreet}.
In a $\dag$-SMC a \emph{basis structure} is a dagger special commutative Frobenius algebra \cite{wosums} - i.e.~a refinement of Carboni and Walters' frobenius algebras \cite{CarboniWalters}. It consists of an internal commutative monoid
\[
(A, \delta:A\rightarrow A\otimes A, \epsilon: A\rightarrow I)
\]
for which $\delta$ and $\delta^\dagger$ satisfy the specialness and Frobenius equations, that is:
\[
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-3.5, -1.5) {};
\node [style=none] (1) at (-1.75, -1.5) {};
\node [style=none] (2) at (0.5, -1.5) {};
\node [style=none] (3) at (1.5, -1.5) {};
\node [style=none] (4) at (3, -1.5) {};
\node [style=none] (5) at (4, -1.5) {};
\node [style=green dot] (6) at (-3.5, -1.75) {};
\node [style=green dot] (7) at (1.5, -2) {};
\node [style=green dot] (8) at (3.5, -2) {};
\node [style=none] (9) at (-2.5, -2.25) {=};
\node [style=none] (10) at (2.5, -2.25) {=};
\node [style=green dot] (11) at (1, -2.5) {};
\node [style=green dot] (12) at (3.5, -2.5) {};
\node [style=green dot] (13) at (-3.5, -2.75) {};
\node [style=none] (14) at (-3.5, -3) {};
\node [style=none] (15) at (-1.75, -3) {};
\node [style=none] (16) at (1, -3) {};
\node [style=none] (17) at (2, -3) {};
\node [style=none] (18) at (3, -3) {};
\node [style=none] (19) at (4, -3) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (16.center) to (11);
\draw (7) to (3.center);
\draw (0.center) to (6);
\draw (13) to (14.center);
\draw[bend right=315, looseness=1.50] (6) to (13);
\draw[bend right=45, looseness=1.50] (6) to (13);
\draw (8) to (5.center);
\draw (1.center) to (15.center);
\draw (8) to (12);
\draw (8) to (4.center);
\draw[bend left=15] (11) to (2.center);
\draw (18.center) to (12);
\draw (7) to (11);
\draw[bend left=15, looseness=1.25] (7) to (17.center);
\draw (12) to (19.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
where we represented $\delta$, $\delta^\dagger$, $\epsilon$ and $\epsilon^\dagger$ graphically as:
\[
\begin{tikzpicture}[scale=0.8]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (3, -1.5) {};
\node [style=none] (1) at (4, -1.5) {};
\node [style=none] (2) at (6, -1.5) {};
\node [style=green dot] (3) at (8, -1.75) {};
\node [style=none] (4) at (10, -1.75) {};
\node [style=green dot] (5) at (3.5, -2) {};
\node [style=green dot] (6) at (6, -2.25) {};
\node [style=none] (7) at (8, -2.5) {};
\node [style=green dot] (8) at (10, -2.5) {};
\node [style=none] (9) at (3.5, -2.75) {};
\node [style=none] (10) at (5.5, -2.75) {};
\node [style=none] (11) at (6.5, -2.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (3) to (7.center);
\draw (6) to (10.center);
\draw (6) to (2.center);
\draw (5) to (1.center);
\draw (6) to (11.center);
\draw (8) to (4.center);
\draw (5) to (9.center);
\draw (5) to (0.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
A key result holding for any basis structure is the \emph{spider theorem} - a high-level abstract account of which is due to Lack \cite{Lack}. This spider theorem essentially allows us, when working with the graphical language, to fuse together any directly connected dots representing $\delta, \delta^\dag, \epsilon$ and $\epsilon^\dag$ from the same basis structure: the theorem guarantees that all morphisms which \emph{look} the same graphically after such a fusing procedure \emph{are} indeed equal. In fact, this property provides an equivalent definition of basis structures, since all the defining axioms of a basis structure are implied by it \cite{ContPhys, RedGreen}.
Among many other things, the spider theorem implies that we have a compact structure \cite{KellyLaplaza}:
\[
\begin{tikzpicture}[scale=0.8]
\begin{pgfonlayer}{nodelayer}
\node [style=green dot] (0) at (4, -1.25) {};
\node [style=none] (1) at (4, -1.25) {};
\node [style=none] (2) at (8.25, -1.25) {};
\node [style=none] (3) at (10.25, -1.25) {};
\node [style=none] (4) at (-1.75, -1.5) {};
\node [style=none] (5) at (-0.75, -1.5) {};
\node [style=green dot] (6) at (6.75, -1.5) {};
\node [style=none] (7) at (-4.25, -1.75) {};
\node [style=none] (8) at (-3.25, -1.75) {};
\node [style=green dot] (9) at (1.5, -1.75) {};
\node [style=none] (10) at (-2.5, -2) {$=$};
\node [style=green dot] (11) at (-1.25, -2) {};
\node [style=none] (12) at (2.75, -2) {$=$};
\node [style=green dot] (13) at (4, -2) {};
\node [style=none] (14) at (6.25, -2) {};
\node [style=none] (15) at (7.25, -2) {};
\node [style=none] (16) at (7.25, -2) {};
\node [style=none] (17) at (8.25, -2) {};
\node [style=none] (18) at (9.25, -2) {$=$};
\node [style=green dot] (19) at (-3.75, -2.25) {};
\node [style=none] (20) at (1, -2.25) {};
\node [style=none] (21) at (2, -2.25) {};
\node [style=none] (22) at (3.5, -2.5) {};
\node [style=none] (23) at (4.5, -2.5) {};
\node [style=green dot] (24) at (7.75, -2.5) {};
\node [style=none] (25) at (-1.25, -2.75) {};
\node [style=green dot] (26) at (-1.25, -2.75) {};
\node [style=none] (27) at (6.25, -2.75) {};
\node [style=none] (28) at (10.25, -2.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (2.center) to (17.center);
\draw[bend left=45, looseness=1.25] (24) to (16.center);
\draw (13) to (1.center);
\draw[bend left=45, looseness=1.25] (19) to (7.center);
\draw[bend right=45, looseness=1.25] (9) to (20.center);
\draw[bend left=45, looseness=1.25] (6) to (15.center);
\draw (11) to (25.center);
\draw (3.center) to (28.center);
\draw (11) to (4.center);
\draw (27.center) to (14.center);
\draw (13) to (23.center);
\draw (13) to (22.center);
\draw[bend left=45, looseness=1.25] (9) to (21.center);
\draw (11) to (5.center);
\draw[bend right=45, looseness=1.25] (19) to (8.center);
\draw[bend right=45, looseness=1.25] (6) to (14.center);
\draw[bend right=45, looseness=1.25] (24) to (17.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
Such a compact structure allows one to exchange the roles of inputs to outputs, and vice versa, so we can essentially ignore those roles. When writing equations it suffices to identify open-ended wires in the LHS with those in the RHS.
Compact structure induces
a covariant involutive \em conjugation \em functor:
\[
\begin{tikzpicture}[scale=1]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (4.25, -1.25) {};
\node [style=none] (1) at (8.25, -1.25) {};
\node [style=green dot] (2) at (6.75, -1.5) {};
\node [style=white dot] (3) at (4.25, -2) {$f$};
\node [style=none] (4) at (5.25, -2) {$\mapsto$};
\node [style=none] (5) at (6.25, -2) {};
\node [style=white dot] (6) at (7.25, -2) {$f^\dagger$};
\node [style=none] (7) at (7.25, -2) {};
\node [style=none] (8) at (8.25, -2) {};
\node [style=green dot] (9) at (7.75, -2.5) {};
\node [style=none] (10) at (4.25, -2.75) {};
\node [style=none] (11) at (6.25, -2.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (1.center) to (8.center);
\draw[bend left=45, looseness=1.25] (9) to (7.center);
\draw[bend right=45, looseness=1.25] (9) to (8.center);
\draw (0.center) to (3.center);
\draw[bend right=45, looseness=1.25] (2) to (5.center);
\draw (3.center) to (10.center);
\draw (11.center) to (5.center);
\draw[bend left=45, looseness=1.25] (2) to (6.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
In the category \textbf{FdHilb} of finite dimensional Hilbert spaces, linear maps, tensor products and adjoints, basis structures are in bijective correspondence with orthonormal bases \cite{CocPavVic}. Explicitly, for a Hilbert space $\mathcal{H}$, $\delta$ `copies' basis vectors and $\epsilon$ `uniformly erases' them:
\begin{equation}\label{eq:concreteBS}
\delta:\mathcal{H}\rightarrow\mathcal{H}\otimes\mathcal{H}:: \ket{i}\mapsto\ket{ii}\qquad\textrm{and}\qquad
\epsilon:\mathcal{H}\rightarrow\mathbb{C}:: \ket{i}\mapsto 1\,.
\end{equation}
Basis structures act as the abstract counterparts to orthonormal bases, and in analogy to the concrete case, we define the \em basis elements \em of general basis structures as the elements $\psi:I\to A$ that are self-conjugate comonoid homomorphisms \cite{RedGreen}, that is, explicitly:
\[
\begin{tikzpicture}[scale=1]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-7.75, -1.5) {};
\node [style=none] (1) at (-6.75, -1.5) {};
\node [style=none] (2) at (-5.25, -1.5) {};
\node [style=none] (3) at (-4.25, -1.5) {};
\node [style=green dot] (4) at (-2, -1.5) {};
\node [style=green dot] (5) at (4, -1.5) {};
\node [style=white dot] (6) at (6.25, -1.5) {$\psi$};
\node [style=green dot] (7) at (-7.25, -2) {};
\node [style=none] (8) at (-6.25, -2) {$\stackrel{C_1}{=}$};
\node [style=none] (9) at (-1.25, -2) {$\stackrel{C_2}{=}$};
\node [style=none] (10) at (0.5, -2) {\mbox{\rm ``empty picture"}};
\node [style=none] (11) at (3.5, -2) {};
\node [style=white dot] (12) at (3.5, -2) {$\psi^\dagger$};
\node [style=none] (13) at (4.5, -2) {};
\node [style=none] (14) at (5.25, -2) {$\stackrel{C_3}{=}$};
\node [style=white dot] (15) at (-5.25, -2.25) {$\psi$};
\node [style=white dot] (16) at (-4.25, -2.25) {$\psi$};
\node [style=none] (17) at (-2, -2.5) {};
\node [style=white dot] (18) at (-2, -2.5) {$\psi$};
\node [style=none] (19) at (4.5, -2.5) {};
\node [style=none] (20) at (6.25, -2.5) {};
\node [style=white dot] (21) at (-7.25, -2.75) {$\psi$};
\node [style=none] (22) at (-7.25, -2.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (13.center) to (19.center);
\draw (4) to (17.center);
\draw (7) to (1.center);
\draw[bend left=45, looseness=1.25] (5) to (13.center);
\draw[bend right=45, looseness=1.25] (5) to (11.center);
\draw (2.center) to (15);
\draw (7) to (0.center);
\draw (6) to (20.center);
\draw (7) to (22.center);
\draw (3.center) to (16);
\end{pgfonlayer}
\end{tikzpicture}\]
In the concrete basis structure of equation \ref{eq:concreteBS}), conjugation boils down to conjugating matrix entries when matrices are expressed in the corresponding orthonormal basis.
In a $\dag$-SMC $\mathcal{C}$, any basis structure induces a bijection $\Lambda: \mathcal{C}(I,A) \rightarrow \mathcal{C}(A,A)$ between states and endomorphisms of $A$, depicted graphically as:
\[
\Lambda(\psi)=\ \ \ \raisebox{-9mm}{
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (4, -1) {};
\node [style=none] (1) at (6.25, -1) {};
\node [style=green dot] (2) at (4, -2) {$\psi$};
\node [style=none] (3) at (5, -2) {$=$};
\node [style=green dot] (4) at (6.25, -2) {};
\node [style=green dot] (5) at (6.75, -2.5) {$\psi$};
\node [style=none] (6) at (4, -3) {};
\node [style=none] (7) at (5.5, -3) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (4) to (1.center);
\draw (0.center) to (2);
\draw (6.center) to (2);
\draw (4) to (5);
\draw (4) to (7.center);
\end{pgfonlayer}
\end{tikzpicture}
}
\]
In \textbf{FdHilb} $\Lambda(\psi)$ is unitary iff $\ket{\psi}$ is unbiased with respect to the basis $\{\ket{i}\}$ copied by $\delta$. Inspired by this, in the general setting a state $\psi$ is defined to be \emph{unbiased with respect to a basis structure $(A,\delta, \epsilon)$} iff $\Lambda(\psi)$ is \em unitary \em i.e.~the dagger and the inverse coincide.
States which are unbiased w.r.t. a particular basis structure are filled with the same colour as the basis structure. In the explicit case of qubits (i.e. the object $\mathbb{C}^2$ in \textbf{FdHilb}) the unbiased states are those of the form $\ket{e_1}+e^{i\alpha}\ket{e_2}$ where $\{\ket{e_i}\}$ is the copied basis, and we label these states with the phase, $\alpha$. For example, if the green basis structure corresponds to $\{\ket{0},\ket{1}\}$ we respectively depict $\ket{0}+e^{i\alpha}\ket{1}$ and $\Lambda(\ket{0}+e^{i\alpha}\ket{1})$ by:
\[
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (10, -1.75) {};
\node [style=none] (1) at (13, -1.75) {};
\node [style=green dot] (2) at (10, -2.25) {$\alpha$};
\node [style=green dot] (3) at (13, -2.25) {$\alpha$};
\node [style=none] (4) at (13, -2.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (2) to (0.center);
\draw (3) to (4.center);
\draw (3) to (1.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
It's easy to show that the unitary $\Lambda(\psi)$s form a group. In the case of any basis structure on the object $\mathbb{C}^2$ in \textbf{FdHilb} (i.e. the qubit case), these unitaries are exactly the phase rotations w.r.t.~the copied basis; for this reason the group formed by the $\Lambda(\psi)$ unitaries is termed the \emph{phase group} \cite{RedGreen}.
In quantum mechanics the relationship between \emph{different} orthonormal bases is clearly of crucial importance -- they represent incompatible observables. Work has been done on abstractly characterising \emph{mutually unbiased} basis structures \cite{RedGreen} -- those corresponding to bases which are unbiased w.r.t.~one another. It was shown that the basis elements of one basis structure are unbiased w.r.t.~the other basis structure -i.e.~imitating the concrete definition of unbiased bases- if and only if \cite{RedGreen}:
\[
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-3.5, -1.25) {};
\node [style=none] (1) at (-1.75, -1.25) {};
\node [style=green dot] (2) at (-3.5, -1.75) {};
\node [style=green dot] (3) at (-1.75, -1.75) {};
\node [style=none] (4) at (-2.5, -2.25) {$\stackrel{H}{=}$};
\node [style=red dot] (5) at (-3.5, -2.75) {};
\node [style=red dot] (6) at (-1.75, -2.75) {};
\node [style=none] (7) at (-3.5, -3.25) {};
\node [style=none] (8) at (-1.75, -3.25) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (5) to (7.center);
\draw (1.center) to (3);
\draw[bend right=45, looseness=1.50] (2) to (5);
\draw (0.center) to (2);
\draw[bend right=315, looseness=1.50] (2) to (5);
\draw (6) to (8.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
where conventionally we denote one such basis structure with green dots and the other with red dots.
As also shown in \cite{RedGreen}, certain importaint pairs of mutually unbiased bases in quantum theory (e.g.~the Z and X spin observables) obey a strictly stronger set of equations,\footnote{That is, strictly stronger provided that we are considering basis structures rather than a monoid-comonoid pair.} namely those that define up-to-scalar-multiples a so-called bialgebra \cite{StreetBook}:
\[
\begin{tikzpicture}[scale=0.8]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-6, -0.75) {};
\node [style=none] (1) at (-5, -0.75) {};
\node [style=none] (2) at (-3, -0.75) {};
\node [style=none] (3) at (-2, -0.75) {};
\node [style=green dot] (4) at (-6, -1.25) {};
\node [style=green dot] (5) at (-5, -1.25) {};
\node [style=red dot] (6) at (-2.5, -1.25) {};
\node [style=none] (7) at (0.5, -1.25) {};
\node [style=none] (8) at (1.5, -1.25) {};
\node [style=none] (9) at (3.75, -1.25) {};
\node [style=none] (10) at (4.75, -1.25) {};
\node [style=none] (11) at (7.25, -1.25) {};
\node [style=none] (12) at (8.25, -1.25) {};
\node [style=none] (13) at (10.5, -1.25) {};
\node [style=none] (14) at (11.5, -1.25) {};
\node [style=none] (15) at (-3.75, -1.75) {$\stackrel{B_1}{=}$};
\node [style=red dot] (16) at (1, -1.75) {};
\node [style=none] (17) at (2.5, -1.75) {$\stackrel{B_2}{=}$};
\node [style=green dot] (18) at (7.75, -1.75) {};
\node [style=none] (19) at (9.25, -1.75) {$\stackrel{B_3}{=}$};
\node [style=green dot] (20) at (3.75, -2) {};
\node [style=green dot] (21) at (4.75, -2) {};
\node [style=red dot] (22) at (10.5, -2) {};
\node [style=red dot] (23) at (11.5, -2) {};
\node [style=red dot] (24) at (-6, -2.25) {};
\node [style=red dot] (25) at (-5, -2.25) {};
\node [style=green dot] (26) at (-2.5, -2.5) {};
\node [style=green dot] (27) at (1, -2.5) {};
\node [style=red dot] (28) at (7.75, -2.5) {};
\node [style=none] (29) at (-6, -2.75) {};
\node [style=none] (30) at (-5, -2.75) {};
\node [style=none] (31) at (-3, -3) {};
\node [style=none] (32) at (-2, -3) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (11.center) to (18);
\draw (3.center) to (6);
\draw (25) to (4);
\draw (14.center) to (23);
\draw (9.center) to (20);
\draw (10.center) to (21);
\draw (13.center) to (22);
\draw (32.center) to (26);
\draw (12.center) to (18);
\draw (25) to (30.center);
\draw (0.center) to (4);
\draw (26) to (31.center);
\draw (1.center) to (5);
\draw (18) to (28);
\draw (6) to (26);
\draw (5) to (24);
\draw (24) to (29.center);
\draw (2.center) to (6);
\draw (16) to (27);
\draw (5) to (25);
\draw (4) to (24);
\draw (7.center) to (16);
\draw (8.center) to (16);
\end{pgfonlayer}
\end{tikzpicture}
\]
below we will omit specifying `up-to-scalar-multiples' when referring to these laws.
The key feature of the graphical calculus is that particular behaviors correspond to radical changes of the topology of the picture e.g.~being `an eigenstate' or being `mutually unbiased' both correspond to pictures decomposing into disconnected components - cf.~equations ($C_1$) and ($H$) respectively. This is the very heart of categorical quantum mechanics: essential concepts are expressed in a language for which all can be reduced to tensor product structure, `disconnecting' then standing for `disentanglement'. It goes without saying that topological distinctions come with clear behavioral differences. In Section \ref{sec:topoclass} we will classify tripartite entanglement also according to this paradigm.
\section{GHZ and W states represented graphically}\label{GHZWgraphicallySec}
We now move to consider the concrete case of qubits. We will be using two mutually unbiased basis structures, $\Delta_Z = (\mathbb{C}^2, \delta_Z, \epsilon_Z)$ which corresponds to the $\ket{0},\ket{1}$ basis and which will be represented graphically by green dots, and $\Delta_X = (\mathbb{C}^2, \delta_X, \epsilon_X)$ which corresponds to the $\ket{+},\ket{-}$ basis and which will be represented graphically by red dots. Viewed as a tripartite state $\mathbb{C}\rightarrow \mathbb{C}^2\otimes\mathbb{C}^2\otimes\mathbb{C}^2$ the green $\delta_Z$ is in fact the GHZ state. Thus the GHZ state can be depicted graphically as:
\begin{equation}\label{GHZ}
\begin{tikzpicture}[scale=0.7]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (4.5, 1) {};
\node [style=green dot] (1) at (4.5, 0) {};
\node [style=none] (2) at (3.75, -0.75) {};
\node [style=none] (3) at (5.25, -0.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (0.center) to (1);
\draw (1) to (2.center);
\draw (3.center) to (1);
\end{pgfonlayer}
\end{tikzpicture}
\end{equation}
The same diagram but with a red dot is also a GHZ class state: it's the state $\ket{+++}+\ket{---}$ which is clearly obtained from the standard GHZ state via local basis transformations.
Since any state in the GHZ class is related to the GHZ state via local linear operations (recall equation \ref{sloccconditionEq}), such states can be depicted as:
\begin{equation}\label{GHZclass}
\begin{tikzpicture}[scale=0.8]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (4.5, 1.75) {};
\node [style=white dot] (1) at (4.5, 1) {$A_3$};
\node [style=green dot] (2) at (4.5, 0) {};
\node [style=white dot] (3) at (3.75, -0.75) {$A_1$};
\node [style=white dot] (4) at (5.25, -0.75) {$A_2$};
\node [style=none] (5) at (3.25, -1.25) {};
\node [style=none] (6) at (5.75, -1.25) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (2) to (5.center);
\draw (0.center) to (2);
\draw (6.center) to (2);
\end{pgfonlayer}
\end{tikzpicture}
\end{equation}
where $A_1$, $A_2$ and $A_3$ are linear maps. Now consider the following diagram:
\begin{equation}\label{W}
\begin{tikzpicture}[scale=0.65]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (4.5, 2.25) {};
\node [style=red dot] (1) at (4.5, 1.5) {};
\node [style=green dot] (2) at (3.75, 0.25) {${\pi\over 3}$};
\node [style=green dot] (3) at (5.25, 0.25) {${\pi\over 3}$};
\node [style=red dot] (4) at (3, -1) {};
\node [style=green dot] (5) at (4.5, -1) {${\pi\over 3}$};
\node [style=red dot] (6) at (6, -1) {};
\node [style=none] (7) at (2.25, -1.5) {};
\node [style=none] (8) at (6.75, -1.5) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (6) to (8.center);
\draw (7.center) to (4);
\draw (1) to (4);
\draw (4) to (6);
\draw (1) to (6);
\draw (0.center) to (1);
\end{pgfonlayer}
\end{tikzpicture}
\end{equation}
When evaluated in \textbf{FdHilb} (i.e. concretely composing the linear maps represented by the graphical components), and ignoring global phase and normalisation, we get the W state.
Comparing the two diagrams \ref{GHZ} and \ref{W} it is striking to see how they seem to embody the global/local entanglement distinction discussed in section \ref{3qubitbackgroundSec}. In the GHZ state, \ref{GHZ}, the three systems are all connected together via the green dot, mirroring the genuine global tripartite entanglement which they share in this state. In contrast, in the W state, \ref{W}, the systems are connected in a pairwise fashion, mirroring the pairwise entanglement in this state.
If we change the phases, from $\pi / 3$ to $0$ then we would wind up with a GHZ-class state:
\begin{equation}\label{spider}
\begin{tikzpicture}[scale=0.5]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (4.5, 2.25) {};
\node [style=red dot] (1) at (4.5, 1.5) {};
\node [style=none] (2) at (9.25, 1.5) {};
\node [style=none] (3) at (7.25, 0.5) {$=$};
\node [style=red dot] (4) at (9.25, 0.5) {};
\node [style=none] (5) at (8.5, -0.25) {};
\node [style=none] (6) at (10, -0.25) {};
\node [style=red dot] (7) at (3, -1) {};
\node [style=red dot] (8) at (6, -1) {};
\node [style=none] (9) at (2.25, -1.5) {};
\node [style=none] (10) at (6.75, -1.5) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (4) to (6.center);
\draw (2.center) to (4);
\draw (8) to (10.center);
\draw (9.center) to (7);
\draw (0.center) to (1);
\draw (7) to (8);
\draw (4) to (5.center);
\draw (1) to (8);
\draw (1) to (7);
\end{pgfonlayer}
\end{tikzpicture}\end{equation}
So depending on the choice of phases we may end up with a W-class or a GHZ-class state. For
the general case:
\begin{equation}\label{alphabetagamma}
\begin{tikzpicture}[scale=0.6]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (4.5, 2.25) {};
\node [style=red dot] (1) at (4.5, 1.5) {};
\node [style=green dot] (2) at (3.75, 0.25) {$\alpha$};
\node [style=green dot] (3) at (5.25, 0.25) {$\beta$};
\node [style=red dot] (4) at (3, -1) {};
\node [style=green dot] (5) at (4.5, -1) {$\gamma$};
\node [style=red dot] (6) at (6, -1) {};
\node [style=none] (7) at (2.25, -1.5) {};
\node [style=none] (8) at (6.75, -1.5) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (0.center) to (1);
\draw (6) to (8.center);
\draw (1) to (4);
\draw (7.center) to (4);
\draw (1) to (6);
\draw (4) to (6);
\end{pgfonlayer}
\end{tikzpicture}
\end{equation}
we will now verify for which values of $\alpha$, $\beta$ and $\gamma$ this state is a GHZ-class state, and for which it is a W-class state. Concretely composing linear maps gives:
\begin{equation}
\ket{\psi}=(1 + e^{i(\alpha + \beta + \gamma)})\ket{000} + (e^{i\alpha} + e^{i(\beta + \gamma)})\ket{011} + (e^{i\beta} + e^{i(\alpha + \gamma)})\ket{101} + (e^{i\gamma} + e^{i(\alpha + \beta)})\ket{110}
\end{equation}
for which the 3-tangle is equal to:
\begin{equation}
\tau_{ABC} = 16 |a| |b| |c| |d|
\end{equation}
where
\begin{equation}
a =1 + e^{i(\alpha + \beta + \gamma)}\quad
b =e^{i\gamma} + e^{i(\alpha + \beta)}\quad
c =e^{i\alpha} + e^{i(\beta + \gamma)}\quad
d =e^{i\beta} + e^{i(\alpha + \gamma)}
\end{equation}\par\noindent
Thus, the tangle is zero (and $\ket{\psi}$ not a GHZ-class state) when one of the following conditions holds:
\begin{equation}\label{constraints}
\begin{array}{l}
\alpha + \beta + \gamma = \pi\\
\gamma - \alpha - \beta = \pi\\
\alpha - \beta - \gamma = \pi\\
\beta - \alpha - \gamma = \pi
\end{array}\end{equation}
So unless one of these conditions holds, we must be able to find linear maps $A_1$, $A_2$, $A_3$, such that:
\begin{equation}
\begin{tikzpicture}[scale=0.7]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-5.75, 2) {};
\node [style=none] (1) at (5.75, 2) {};
\node [style=none] (2) at (0, 1.5) {};
\node [style=red dot] (3) at (-5.75, 1.25) {};
\node [style=white dot] (4) at (5.75, 1.25) {$A_3$};
\node [style=white dot] (5) at (0, 0.75) {$A_3$};
\node [style=red dot] (6) at (5.75, 0.25) {};
\node [style=green dot] (7) at (-6.5, 0) {$\alpha$};
\node [style=green dot] (8) at (-5, 0) {$\beta$};
\node [style=none] (9) at (-2.5, 0) {$=$};
\node [style=none] (10) at (2.5, 0) {$=$};
\node [style=red dot] (11) at (0, -0.25) {};
\node [style=red dot] (12) at (5.25, -0.5) {};
\node [style=red dot] (13) at (6.25, -0.5) {};
\node [style=white dot] (14) at (-0.75, -1) {$A_1$};
\node [style=white dot] (15) at (0.75, -1) {$A_2$};
\node [style=red dot] (16) at (-7.25, -1.25) {};
\node [style=green dot] (17) at (-5.75, -1.25) {$\gamma$};
\node [style=red dot] (18) at (-4.25, -1.25) {};
\node [style=white dot] (19) at (4.25, -1.25) {$A_1$};
\node [style=white dot] (20) at (7.25, -1.25) {$A_2$};
\node [style=none] (21) at (-1.25, -1.5) {};
\node [style=none] (22) at (1.25, -1.5) {};
\node [style=none] (23) at (-8, -1.75) {};
\node [style=none] (24) at (-3.5, -1.75) {};
\node [style=none] (25) at (3.5, -1.75) {};
\node [style=none] (26) at (8, -1.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (16) to (18);
\draw (3) to (16);
\draw (18) to (24.center);
\draw (25.center) to (12);
\draw (0.center) to (3);
\draw (11) to (21.center);
\draw (22.center) to (11);
\draw (6) to (12);
\draw (12) to (13);
\draw (23.center) to (16);
\draw (6) to (13);
\draw (3) to (18);
\draw (2.center) to (11);
\draw (13) to (26.center);
\draw (1.center) to (6);
\end{pgfonlayer}
\end{tikzpicture}
\end{equation}
i.e. we should be able to `pull out' the three phases, in the process transforming them into linear maps, leaving a free central triangle, which can then be closed down to a single red dot using the spider law. In the case where one of the four conditions does hold, something stops us from pulling the phases out, and they cause a `log jam', preventing us from closing down the triangle to a dot, and dooming the state to be without genuine global tripartite entanglement.
\section{GHZ and W states analyzed graphically}\label{sec:topoclass}
\newcommand{\begin{color}{red}\bullet\end{color}}{\begin{color}{red}\bullet\end{color}}
\newcommand{\rpt}{
\raisebox{-1.8mm}{
\begin{tikzpicture}[scale=0.44]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (0, 0.5) {};
\node [style=red dot] (1) at (0, -0.25) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (1) to (0.center);
\end{pgfonlayer}
\end{tikzpicture}
}}
\newcommand{\rppt}{
\raisebox{-1.8mm}{
\begin{tikzpicture}[scale=0.44]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (0, 0.5) {};
\node [style=red dot] (1) at (0, -0.25) {\small$\!\pi\!$};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (1) to (0.center);
\end{pgfonlayer}
\end{tikzpicture}
}}
\newcommand{\gbs}{\bigl(\!\!
\raisebox{-2.3mm}{
\begin{tikzpicture}[scale=0.44]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (2.75, -1.5) {};
\node [style=none] (1) at (3.75, -1.5) {};
\node [style=green dot] (2) at (3.25, -2) {};
\node [style=green dot] (3) at (4.75, -2) {};
\node [style=none] (4) at (4, -2.25) {$,$};
\node [style=none] (5) at (3.25, -2.75) {};
\node [style=none] (6) at (4.75, -2.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (2) to (0.center);
\draw (2) to (5.center);
\draw (2) to (1.center);
\draw (3) to (6.center);
\end{pgfonlayer}
\end{tikzpicture}
}\!\bigr)}
\newcommand{\gbr}{\bigl(\!\!
\raisebox{-2.3mm}{
\begin{tikzpicture}[scale=0.44]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (2.75, -1.5) {};
\node [style=none] (1) at (3.75, -1.5) {};
\node [style=red dot] (2) at (3.25, -2) {};
\node [style=red dot] (3) at (4.75, -2) {};
\node [style=none] (4) at (4, -2.25) {$,$};
\node [style=none] (5) at (3.25, -2.75) {};
\node [style=none] (6) at (4.75, -2.75) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (2) to (0.center);
\draw (2) to (5.center);
\draw (2) to (1.center);
\draw (3) to (6.center);
\end{pgfonlayer}
\end{tikzpicture}
}\!\bigr)}
In this section we wish to produce the constraints (\ref{constraints}) not by direct computation but within the diagrammatic language. We will show how these four constraints can be classified.
First observe that:
\[
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-3.5, -0.75) {};
\node [style=none] (1) at (-0.25, -0.75) {};
\node [style=none] (2) at (3.75, -0.75) {};
\node [style=none] (3) at (8, -0.75) {};
\node [style=none] (4) at (10.5, -0.75) {};
\node [style=red dot] (5) at (-3.5, -1.25) {};
\node [style=red dot] (6) at (-0.25, -1.25) {};
\node [style=green dot] (7) at (3.75, -1.5) {};
\node [style=green dot] (8) at (8, -1.5) {};
\node [style=red dot] (9) at (7, -1.75) {};
\node [style=green dot] (10) at (10.5, -1.75) {};
\node [style=green dot] (11) at (-4, -2) {$\xi$};
\node [style=green dot] (12) at (-3, -2) {$\zeta$};
\node [style=none] (13) at (-2.25, -2) {$=$};
\node [style=green dot] (14) at (-1.5, -2) {$\xi$};
\node [style=green dot] (15) at (-0.75, -2) {};
\node [style=green dot] (16) at (0.25, -2) {};
\node [style=green dot] (17) at (1, -2) {$\zeta$};
\node [style=none] (18) at (1.75, -2) {$\stackrel{B_1}{=}$};
\node [style=green dot] (19) at (2.5, -2) {$\xi$};
\node [style=red dot] (20) at (3.5, -2) {};
\node [style=green dot] (21) at (5, -2) {$\zeta$};
\node [style=none] (22) at (5.75, -2) {$=$};
\node [style=none] (23) at (8.75, -2) {$=$};
\node [style=green dot] (24) at (6.5, -2.5) {$\xi$};
\node [style=green dot] (25) at (7.5, -2.5) {$\zeta$};
\node [style=white dot] (26) at (9.75, -2.5) {\!\!$\xi\begin{color}{red}\bullet\end{color}\zeta$\!\!};
\node [style=red dot] (27) at (-3.5, -2.75) {};
\node [style=red dot] (28) at (-0.25, -2.75) {};
\node [style=none] (29) at (-3.5, -3.25) {};
\node [style=none] (30) at (-0.25, -3.25) {};
\node [style=none] (31) at (3.75, -3.25) {};
\node [style=none] (32) at (8, -3.25) {};
\node [style=none] (33) at (10.5, -3.25) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw[bend right=45, looseness=1.50] (5) to (27);
\draw (8) to (9);
\draw (4.center) to (10);
\draw (27) to (29.center);
\draw (10) to (33.center);
\draw (19) to (20);
\draw (14) to (15);
\draw (8) to (32.center);
\draw (16) to (17);
\draw[bend right=315, looseness=1.50] (6) to (28);
\draw (3.center) to (8);
\draw (2.center) to (7);
\draw (20) to (21);
\draw (1.center) to (6);
\draw (7) to (31.center);
\draw (7) to (20);
\draw[bend right=45, looseness=1.50] (6) to (28);
\draw (24) to (9);
\draw (9) to (25);
\draw (0.center) to (5);
\draw[bend right=315, looseness=1.50] (5) to (27);
\draw (28) to (30.center);
\draw (10) to (26);
\end{pgfonlayer}
\end{tikzpicture}\]
where we set $\xi\begin{color}{red}\bullet\end{color}\zeta:=\delta_X\circ(\xi\otimes\zeta)$. It then follows by equation $C_1$ that:
\begin{proposition}\label{MainProp}
If basis structures $\gbr$ and $\gbs$ form a bialgebra then the endomorphism:
\[
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-3.5, -0.75) {};
\node [style=red dot] (1) at (-3.5, -1.25) {};
\node [style=green dot] (2) at (-4, -2) {$\xi$};
\node [style=green dot] (3) at (-3, -2) {$\zeta$};
\node [style=red dot] (4) at (-3.5, -2.75) {};
\node [style=none] (5) at (-3.5, -3.25) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (0.center) to (1);
\draw[bend right=315, looseness=1.50] (1) to (4);
\draw[bend right=45, looseness=1.50] (1) to (4);
\draw (4) to (5.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
is disconnected whenever $\xi\begin{color}{red}\bullet\end{color}\zeta$ is a basis element of $\gbs$. Explicitly, up-to-a-scalar we obtain:
\[
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-3.5, -0.75) {};
\node [style=red dot] (1) at (-3.5, -1.5) {\!\!$\xi\begin{color}{red}\bullet\end{color}\zeta$\!\!};
\node [style=red dot] (2) at (-3.5, -2.75) {\!\!$\xi\begin{color}{red}\bullet\end{color}\zeta$\!\!};
\node [style=none] (3) at (-3.5, -3.5) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (0.center) to (1);
\draw (2) to (3.center);
\end{pgfonlayer}
\end{tikzpicture}\]
\end{proposition}
\begin{definition}\em
For basis structures $\gbr$ and $\gbs$ on $A$ which form a bialgebra we call a pair of $\gbs$-phases $\xi, \zeta:I\to A$ \em supplementary \em when $i:=\xi\begin{color}{red}\bullet\end{color}\zeta:I\to A$ is (up-to-a-scalar) a basis element of $\gbr$. More specifically, we say that $\xi$ and $\zeta$ are \em $i$-supplementary\em.
\end{definition}
In the case of the concrete basis structures $\Delta_Z$ and $\Delta_X$, since $\Delta_Z$ has two basis elements $\ket{0}$ and $\ket{1}$, there will be two kinds of supplementary. We have
\begin{eqnarray*}
\xi\begin{color}{red}\bullet\end{color}\zeta
&=& \delta_Z\left((\ket{0}+ e^{i\xi} \ket{1})\otimes (\ket{0}+ e^{i\zeta} \ket{1}) \right)\\
&=& \delta_Z\left((\ket{00}+ e^{i(\xi+ \zeta)} \ket{11})+( e^{i\xi} \ket{01} + e^{i\zeta} \ket{10}) \right)\\
&=& (1+ e^{i(\xi+ \zeta)} )\ket{0}+( e^{i\xi} + e^{i\zeta} )\ket{1} \\
\end{eqnarray*}\par\noindent
Hence $\xi\begin{color}{red}\bullet\end{color}\zeta$ is either proportional to $\ket{0}$ and $\ket{1}$ respectively when:
\[
\xi+\zeta= \pi\mbox{\ \ $\Rightarrow$~$\ket{1}$-supplementary}\qquad\mbox{or}\qquad\xi+ \pi= \zeta \mbox{\ \ $\Rightarrow$~$\ket{0}$-supplementary}
\]
We can now reproduce equations (\ref{constraints}) in terms of `disconnectedness of a picture'. We have:
\[
\begin{tikzpicture}[scale=1]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-2.5, 1.25) {};
\node [style=none] (1) at (1.25, 1.25) {};
\node [style=none] (2) at (5.25, 1.25) {};
\node [style=red dot] (3) at (-2.5, 0.75) {};
\node [style=red dot] (4) at (1.25, 0.75) {};
\node [style=green dot] (5) at (5.25, 0.25) {};
\node [style=green dot] (6) at (-3, 0) {$\beta$};
\node [style=green dot] (7) at (-2, 0) {$\gamma$};
\node [style=none] (8) at (-0.5, 0) {$=$};
\node [style=green dot] (9) at (0.75, 0) {\scriptsize$\!\!\alpha+\beta\!\!$};
\node [style=green dot] (10) at (1.75, 0) {$\gamma$};
\node [style=none] (11) at (2.75, 0) {$=$};
\node [style=green dot] (12) at (4.25, -0.5) {\scriptsize$\!(\alpha+\beta)\begin{color}{red}\bullet\end{color}\gamma$\!};
\node [style=red dot] (13) at (-3.5, -0.75) {};
\node [style=green dot] (14) at (-2.5, -0.75) {$\alpha$};
\node [style=red dot] (15) at (-1.5, -0.75) {};
\node [style=red dot] (16) at (1.25, -0.75) {};
\node [style=none] (17) at (-1, -1) {};
\node [style=red dot] (18) at (-4.5, -1.25) {};
\node [style=none] (19) at (1.25, -1.25) {};
\node [style=none] (20) at (5.25, -1.25) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (3) to (13);
\draw (0.center) to (3);
\draw (16) to (19.center);
\draw (3) to (15);
\draw (17.center) to (15);
\draw (2.center) to (5);
\draw (13) to (18);
\draw (5) to (20.center);
\draw (15) to (13);
\draw[bend right=315, looseness=1.50] (4) to (16);
\draw (5) to (12);
\draw[bend right=45, looseness=1.50] (4) to (16);
\draw (1.center) to (4);
\end{pgfonlayer}
\end{tikzpicture}
\]
so we obtain the two first of equations (\ref{constraints}); the remaining two equations arise when `plugging' $\rpt$ in the other corners of the W-state.
Now, since by plugging $\rpt$ we obtain equations that are characteristic for the W class states, one expects that one can find a derivation of $\rpt$ from such a state. This is indeed the case. As shown in \cite{CK} also the W state can be endowed with a commutative Frobenius algebra structure, although this structure is not dagger, and more importantly, not special. We still have some kind of a spider theorem, but now it also accounts for the number of loops in a picture, which has to be preserved. It is indeed `the value of the loop' that is the characteristic difference between GHZ class states and W class states; we have:
\[
\raisebox{-7mm}{\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-4.75, -1.5) {};
\node [style=none] (1) at (-3, -1.5) {};
\node [style=dot] (2) at (-3, -1.75) {};
\node [style=dot] (3) at (-4.75, -2.25) {};
\node [style=none] (4) at (-4, -2.25) {=};
\node [style=none] (5) at (-2, -2.25) {=};
\node [style=dot] (6) at (-5.25, -2.5) {};
\node [style=dot] (7) at (-3, -2.75) {};
\node [style=none] (8) at (-4.75, -3) {};
\node [style=none] (9) at (-3, -3) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw[out=-45, in=-135, loop] (6) to ();
\draw[bend left=315, looseness=1.50] (2) to (7);
\draw (8.center) to (0.center);
\draw[bend left=45, looseness=1.50] (2) to (7);
\draw (6) to (3);
\draw (7) to (9.center);
\draw (1.center) to (2);
\end{pgfonlayer}
\end{tikzpicture}}
\ \ \ \ \left\{
\begin{array}{cc}
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-4.75, -1.5) {};
\node [style=none] (1) at (-4.75, -3) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (0.center) to (1.center);
\end{pgfonlayer}
\end{tikzpicture}
& \raisebox{6mm}{\mbox{for the GHZ state}}\vspace{3mm}\\
\begin{tikzpicture}
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (0, 2) {};
\node [style=dot] (1) at (0, 1.75) {};
\node [style=dot] (2) at (0, 0.5) {};
\node [style=none] (3) at (0, 0.25) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw[out=45, in=135, loop] (2) to ();
\draw (2) to (3.center);
\draw (1) to (0.center);
\draw[out=-45, in=-135, loop] (1) to ();
\end{pgfonlayer}
\end{tikzpicture}
& \raisebox{7mm}{\mbox{for the W state}}
\end{array}
\right.
\]
The commutative Frobenius algebra structure on W is given by:
\[
\left(\raisebox{-1.1cm}{\begin{tikzpicture}[scale=0.92]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (0, 1) {};
\node [style=none] (1) at (2, 1) {};
\node [style=none] (2) at (4.25, 1) {};
\node [style=red dot] (3) at (0, 0.5) {$\pi$};
\node [style=green dot] (4) at (1, 0.5) {${\pi\over 3}$};
\node [style=red dot] (5) at (2, 0.5) {$\pi$};
\node [style=red dot] (6) at (4.25, 0.5) {$\pi$};
\node [style=green dot] (7) at (0.5, -0.25) {${\pi\over 3}$};
\node [style=green dot] (8) at (1.5, -0.25) {${\pi\over 3}$};
\node [style=green dot] (9) at (3.75, -0.25) {${\pi\over 3}$};
\node [style=green dot] (10) at (4.75, -0.25) {${\pi\over 3}$};
\node [style=red dot] (11) at (7, -0.25) {};
\node [style=none] (12) at (8.5, -0.25) {};
\node [style=none] (13) at (2.5, -0.75) {$,$};
\node [style=none] (14) at (6.25, -0.75) {$,$};
\node [style=none] (15) at (7, -0.75) {};
\node [style=none] (16) at (7.75, -0.75) {$,$};
\node [style=red dot] (17) at (8.5, -0.75) {$\pi$};
\node [style=red dot] (18) at (1, -1) {};
\node [style=red dot] (19) at (3.25, -1) {};
\node [style=green dot] (20) at (4.25, -1) {${\pi\over 3}$};
\node [style=red dot] (21) at (5.25, -1) {};
\node [style=none] (22) at (1, -1.5) {};
\node [style=none] (23) at (3.25, -1.5) {};
\node [style=none] (24) at (5.25, -1.5) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (22.center) to (18);
\draw (18) to (5);
\draw (21) to (24.center);
\draw (0.center) to (3);
\draw (2.center) to (6);
\draw (5) to (1.center);
\draw (6) to (21);
\draw (3) to (5);
\draw (23.center) to (19);
\draw (6) to (19);
\draw (18) to (3);
\draw (12.center) to (17);
\draw (15.center) to (11);
\draw (19) to (21);
\end{pgfonlayer}
\end{tikzpicture}}\right)
\]
and one can compute that:
\[
\begin{tikzpicture}[scale=0.92]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-0.25, 2.5) {};
\node [style=red dot] (1) at (-0.25, 2) {$\pi$};
\node [style=green dot] (2) at (-0.75, 1.25) {${\pi\over 3}$};
\node [style=green dot] (3) at (0.25, 1.25) {${\pi\over 3}$};
\node [style=none] (4) at (-3.5, 0.75) {};
\node [style=dot] (5) at (-3.5, 0.5) {};
\node [style=red dot] (6) at (-1.25, 0.5) {};
\node [style=green dot] (7) at (-0.25, 0.5) {${\pi\over 3}$};
\node [style=red dot] (8) at (0.75, 0.5) {};
\node [style=none] (9) at (-5.5, 0.25) {};
\node [style=none] (10) at (2.75, 0.25) {};
\node [style=dot] (11) at (-5.5, 0) {};
\node [style=none] (12) at (-4.5, 0) {$=$};
\node [style=none] (13) at (-2.25, 0) {$=$};
\node [style=none] (14) at (1.75, 0) {$=$};
\node [style=red dot] (15) at (2.75, -0.25) {};
\node [style=dot] (16) at (-3.5, -0.5) {};
\node [style=red dot] (17) at (-1.25, -0.5) {$\pi$};
\node [style=none] (18) at (-1.25, -0.5) {$\pi$};
\node [style=green dot] (19) at (-0.25, -0.5) {${\pi\over 3}$};
\node [style=red dot] (20) at (0.75, -0.5) {$\pi$};
\node [style=none] (21) at (0.75, -0.5) {$\pi$};
\node [style=dot] (22) at (-3.5, -1) {};
\node [style=green dot] (23) at (-0.75, -1.25) {${\pi\over 3}$};
\node [style=green dot] (24) at (0.25, -1.25) {${\pi\over 3}$};
\node [style=red dot] (25) at (-0.25, -2) {};
\node [style=red dot] (26) at (-0.25, -3) {$\pi$};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (26) to (25);
\draw (18.center) to (6);
\draw[bend left=45, looseness=1.50] (5) to (16);
\draw (0.center) to (1);
\draw (4.center) to (5);
\draw (16) to (22);
\draw[out=-45, in=-135, loop] (11) to ();
\draw (11) to (9.center);
\draw (25) to (20);
\draw (1) to (6);
\draw (8) to (21.center);
\draw (17) to (20);
\draw (25) to (17);
\draw (1) to (8);
\draw (15) to (10.center);
\draw[bend left=315, looseness=1.50] (5) to (16);
\draw (6) to (8);
\end{pgfonlayer}
\end{tikzpicture}
\]
We can establish the last equality within graphical calculus by showing that, like RHS, also LHS is orthogonal to $\rppt$. Indeed:
\[
\begin{tikzpicture}[scale=0.92]
\begin{pgfonlayer}{nodelayer}
\node [style=red dot] (0) at (-6.5, 3) {$\pi$};
\node [style=red dot] (1) at (-6.5, 2) {$\pi$};
\node [style=red dot] (2) at (-2.5, 2) {};
\node [style=green dot] (3) at (-7, 1.25) {${\pi\over 3}$};
\node [style=green dot] (4) at (-6, 1.25) {${\pi\over 3}$};
\node [style=green dot] (5) at (-3, 1.25) {${\pi\over 3}$};
\node [style=green dot] (6) at (-2, 1.25) {${\pi\over 3}$};
\node [style=red dot] (7) at (-7.5, 0.5) {};
\node [style=green dot] (8) at (-6.5, 0.5) {${\pi\over 3}$};
\node [style=red dot] (9) at (-5.5, 0.5) {};
\node [style=red dot] (10) at (-3.5, 0.5) {};
\node [style=green dot] (11) at (-2.5, 0.5) {${\pi\over 3}$};
\node [style=red dot] (12) at (-1.5, 0.5) {};
\node [style=red dot] (13) at (0.5, 0.5) {$\pi$};
\node [style=red dot] (14) at (2.5, 0.5) {$\pi$};
\node [style=none] (15) at (-4.5, 0) {$=$};
\node [style=none] (16) at (-0.5, 0) {$\stackrel{Prop.~\ref{MainProp}}{=}$};
\node [style=red dot] (17) at (-7.5, -0.5) {$\pi$};
\node [style=none] (18) at (-7.5, -0.5) {$\pi$};
\node [style=green dot] (19) at (-6.5, -0.5) {${\pi\over 3}$};
\node [style=none] (20) at (-5.5, -0.5) {$\pi$};
\node [style=red dot] (21) at (-5.5, -0.5) {$\pi$};
\node [style=none] (22) at (-3.5, -0.5) {$\pi$};
\node [style=red dot] (23) at (-3.5, -0.5) {$\pi$};
\node [style=green dot] (24) at (-2.5, -0.5) {${\pi\over 3}$};
\node [style=none] (25) at (-1.5, -0.5) {$\pi$};
\node [style=red dot] (26) at (-1.5, -0.5) {$\pi$};
\node [style=red dot] (27) at (0.5, -0.5) {$\pi$};
\node [style=none] (28) at (0.5, -0.5) {$\pi$};
\node [style=green dot] (29) at (1.5, -0.5) {${\pi\over 3}$};
\node [style=none] (30) at (2.5, -0.5) {$\pi$};
\node [style=red dot] (31) at (2.5, -0.5) {$\pi$};
\node [style=none] (32) at (4.5, -0.5) {};
\node [style=red dot] (33) at (4.5, -0.5) {};
\node [style=green dot] (34) at (5.5, -0.5) {${\pi\over 3}$};
\node [style=none] (35) at (6.5, -0.5) {};
\node [style=red dot] (36) at (6.5, -0.5) {};
\node [style=green dot] (37) at (8.5, -0.5) {$\pi$};
\node [style=green dot] (38) at (-7, -1.25) {${\pi\over 3}$};
\node [style=green dot] (39) at (-6, -1.25) {${\pi\over 3}$};
\node [style=green dot] (40) at (-3, -1.25) {${\pi\over 3}$};
\node [style=green dot] (41) at (-2, -1.25) {${\pi\over 3}$};
\node [style=green dot] (42) at (1, -1.25) {${\pi\over 3}$};
\node [style=green dot] (43) at (2, -1.25) {${\pi\over 3}$};
\node [style=none] (44) at (3.5, -1.25) {$=$};
\node [style=green dot] (45) at (5, -1.25) {${\pi\over 3}$};
\node [style=green dot] (46) at (6, -1.25) {${\pi\over 3}$};
\node [style=none] (47) at (7.5, -1.25) {$=$};
\node [style=red dot] (48) at (8.5, -1.25) {};
\node [style=red dot] (49) at (-6.5, -2) {};
\node [style=red dot] (50) at (-2.5, -2) {};
\node [style=red dot] (51) at (1.5, -2) {};
\node [style=red dot] (52) at (5.5, -2) {};
\node [style=red dot] (53) at (8.5, -2.25) {$\pi$};
\node [style=red dot] (54) at (-6.5, -3) {$\pi$};
\node [style=red dot] (55) at (-2.5, -3) {$\pi$};
\node [style=red dot] (56) at (1.5, -3) {$\pi$};
\node [style=red dot] (57) at (5.5, -3) {$\pi$};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (1) to (7);
\draw (33) to (36);
\draw (0) to (1);
\draw (12) to (25.center);
\draw (7) to (9);
\draw (28.center) to (13);
\draw (2) to (12);
\draw (54) to (49);
\draw[bend right=60] (37) to (48);
\draw[bend left=45] (37) to (48);
\draw (51) to (31);
\draw (49) to (17);
\draw (23) to (26);
\draw (52) to (33);
\draw (22.center) to (10);
\draw (50) to (26);
\draw (55) to (50);
\draw (17) to (21);
\draw (53) to (48);
\draw (56) to (51);
\draw (27) to (31);
\draw (52) to (36);
\draw (10) to (12);
\draw (57) to (52);
\draw (50) to (23);
\draw (49) to (21);
\draw (18.center) to (7);
\draw (51) to (27);
\draw (1) to (9);
\draw (2) to (10);
\draw (14) to (30.center);
\draw (9) to (20.center);
\end{pgfonlayer}
\end{tikzpicture}
\]
and a $\pi$-gate of one color within a loop-point of the other color is always $0$.
\section{Outlook}
We showed that W class states occur when a singularity takes place in the graphical calculus: plugging with a certain point `special' structural value leeds to disconnectedness. The methods in this paper for analyzing the threepartite qubit states admit extrapolation to multiple qubits; for example, for:
\[
\begin{tikzpicture}[scale=0.6]
\begin{pgfonlayer}{nodelayer}
\node [style=none] (0) at (-3.5, 1.5) {};
\node [style=none] (1) at (-0.5, 1.5) {};
\node [style=red dot] (2) at (-3, 1) {};
\node [style=green dot] (3) at (-2, 1) {$\delta$};
\node [style=red dot] (4) at (-1, 1) {};
\node [style=green dot] (5) at (-3, 0) {$\alpha$};
\node [style=green dot] (6) at (-1, 0) {$\gamma$};
\node [style=red dot] (7) at (-3, -1) {};
\node [style=green dot] (8) at (-2, -1) {$\beta$};
\node [style=red dot] (9) at (-1, -1) {};
\node [style=none] (10) at (-3.5, -1.5) {};
\node [style=none] (11) at (-0.5, -1.5) {};
\end{pgfonlayer}
\begin{pgfonlayer}{edgelayer}
\draw (9) to (11.center);
\draw (7) to (9);
\draw (2) to (7);
\draw (10.center) to (7);
\draw (4) to (9);
\draw (2) to (4);
\draw (0.center) to (2);
\draw (1.center) to (4);
\end{pgfonlayer}
\end{tikzpicture}
\]
by plugging two corners, analogous singularities occur when either $(\alpha + \beta) \begin{color}{red}\bullet\end{color} (\gamma +\delta)$ or $\alpha \begin{color}{red}\bullet\end{color} (\beta + \gamma +\delta)$ are basis elements for $\gbs$. In this case the status of $\rpt$ isn't as clear as in the previous section where we could rely on the results of \cite{CK}, and hence requires further investigation.
On the other hand, while the results in \cite{CK} are restricted to the highly symmetrical so-called Frobenius states, the methods presented here don't have such a constraint, and may lead to an algebraic account on more general states than tripartite ones.
|
\section{Introduction}
A convex equal-area polygon is a simple polygon bounding a convex planar domain such that all triangles formed by three consecutive vertices have the same area (\cite{Harel03}). As a consequence, each four consecutive vertices form a trapezium.
It is worthwhile to mention that the class of equal-area polygons includes also non-convex polygons
and appears as a set of maximizers of some area functionals (see \cite{Gronchi08}).
{\it Affinely regular polygons} are affine transforms of convex regular polygons and thus are convex equal-area polygons.
The affinely regular polygons appear as maximizers of area functionals and in many other different contexts (see \cite{Fischer98}).
We begin with the study of the space ${\mathcal T}_n$ of convex equal area polygons with $n$ vertices, which can be thought as a subset of the $(2n)$-dimensional affine space.
The space ${\mathcal T}_n$ has dimension $(n-5)$ and in section \ref{section:TrapezoidalPolygons} we describe in detail its structure.
We also consider the problem of approximating a closed convex curve by polygon in ${\mathcal T}_n$. We show how to
construct a convex equal-area $n$-gon whose vertices, for $n$ large, are close to a uniform distribution of points on the curve with respect to affine arc-length.
For convex equal-area polygons, we define in section \ref{section:DiscreteAffineGeometry} the affine tangent vector, affine normal vector at each vertex and the affine curvature $\mu$ at each edge in a straightforward way.
In this context, we define a sextactic edge as an edge where $\Delta\mu$ is changing sign. Then we show that there are at least six sextactic edges in a convex equal-area polygon,
which is a discrete analog of the well-known six vertices theorem.
In section \ref{section:Evolutas} we define the affine evolute and the $\lambda$-parallel curves for convex equal-area polygons. As in the smooth case, the six vertices theorem may be re-phrased as the existence of at least six cusps on the affine evolute (\cite{Ovsienko01},\cite{Tabach00}). Besides, we show that cusps of the parallels belong to the evolute, as in the smooth case.
The set of self-intersections of the parallels is called the affine distance symmetry set (ADSS) (\cite{Giblin98}). Again we prove here two properties analogous to the smooth case: cusps of the ADSS occur at points of the
affine evolute and endpoints of the ADSS are exactly the cusps of the evolute (\cite{Giblin08}).
Finally in section \ref{section:Isoperimetric} we prove an isoperimetric inequality by using a Minkowski mixed area inequality for parallel polygons. This inequality has a well-known smooth counterpart
saying that the integral of the affine curvature with respect to affine arc-length is at most $\frac{L^2}{2A}$, where $L$ is the affine perimeter
of the convex curve and $A$ is the area that it bounds, with equality holding only for ellipses (see \cite{Sapiro94}). In the convex equal-area polygon case, the equality holds only for affinely regular polygons.
We have used the free software GeoGebra (\cite{GeoGebra}) for all figures and many experiments during the preparation of the paper. Applets of some of these experiments are available at \cite{Applets}.
We would like to thank
the GeoGebra team for this excellent mathematical tool.
\section{Review of some concepts of affine geometry of convex curves}
In this section we review the basic affine differential concepts associated with convex curves. Since in this paper we deal
with a discrete model, these concepts are not strictly necessary for understanding the paper. Nevertheless,
they are the inspiration for most definitions and results.
\paragraph*{Affine arc-length, tangent, normal and curvature} Let $\gamma:[a,b]\to\mathbb{R}^2$ be a regular parameterized strictly convex curve, i.e.,
$\left[\gamma'(t), \gamma''(t) \right]\neq 0$. The parameter $s$ defined by
$s(t)=\int_a^t\left[\gamma'(t), \gamma''(t) \right]^{1/3}dt\ $
is called the affine arc-length parameter. If $\gamma$ is parameterized by affine arc-length,
\begin{equation}\label{affinearclength}
\left[\gamma'(s), \gamma''(s) \right]=1.
\end{equation}
The vector $\gamma'(s)$ is called the affine tangent and $\gamma''(s)$ is called the affine normal.
Differentiating \eqref{affinearclength} we obtain that $\gamma'''(s)$ is parallel to $\gamma'(s)$.
Thus we write
$\gamma'''(s)+\mu(s)\gamma'(s)=0$,
where the scalar function $\mu(s)$ is called the affine curvature. A {\it vertex} or {\it sextactic point} of $\gamma$ is a point
where the derivative of the curvature is zero.
\paragraph*{Affine evolute, parallels and affine distance symmetry set}
The envelope of the family of affine normal lines is called the affine evolute of $\gamma$. For a fixed $\lambda\in\mathbb{R}$, the set
of points of the form
$\gamma(s)+\lambda\gamma''(s)$, $s\in[a,b]$,
is called the $\lambda$-parallel. The ADSS is the locus of self-intersection of parallels (\cite{Giblin98},\cite{Giblin08}).
\paragraph*{Six vertices theorem}
A well-known theorem of global affine planar geometry says that any convex closed planar curve must have at least six vertices (\cite{Buchin83}).
Equivalently, the affine evolute must have at least six cusps.
\paragraph*{An affine isoperimetric inequality} There are some different inequalities referred as affine isoperimetric inequalities. In this paper
we shall consider the following one:
for a closed convex curve $\gamma$ with affine perimeter $L$ enclosing a region of area $A$,
\begin{equation}\label{SmoothIsoIneq}
\int_{\gamma}\mu ds\leq \frac{L^2}{2A},
\end{equation}
with equality if and only if $\gamma$ is an ellipsis (\cite{Sapiro94}).
\section{Convex Equal-Area Polygons}\label{section:TrapezoidalPolygons}
\subsection{Basic definitions}
Consider a closed convex polygon $P$ with vertices $P_1,P_2,...,P_n$, $n\geq 5$, and oriented sides
$$
\v_{i+\frac{1}{2}}=P_{i+1}-P_{i},
$$
$1\leq i\leq n$ (throughout the paper, all indices are taken modulo $n$). The polygon $P$ is said to be {\it equal-area} if
\begin{equation}\label{DefEqualArea}
[\v_{i-\frac{1}{2}}, \v_{i+\frac{1}{2}}]=[\v_{i+\frac{1}{2}}, \v_{i+\frac{3}{2}}],
\end{equation}
for every $1\leq i\leq n$.
This is equivalent to say that $P_{i+2}-P_{i-1}$ is parallel to $\v_{i+\frac{1}{2}}$, for every $1\leq i\leq n$ (see figure \ref{sides10}).
We write
\begin{equation}\label{RatioDiagonalSide}
P_{i+2}-P_{i-1}=(3-\mu_{i+\frac{1}{2}})\v_{i+\frac{1}{2}},
\end{equation}
for some scalar $\mu_{i+\frac{1}{2}}<2$.
The following lemma will be useful below:
\begin{lemma}\label{Hyperbola}
Consider a trapezium $ABCD$ with $AD\parallel BC$. Then $A$ and $D$ belong to a hyperbola whose asymptotes are the lines parallel to $AB$ and $CD$ passing through $C$ and $B$, respectively.
As a consequence, in a convex equal-area polygon, $P_{i}$ and $P_{i+3}$ belong to a hyperbola whose asymptotes are $P_ {i-1}P_{i+2}$ and $P_{i+1}P_{i+4}$, for any $1\leq i\leq n$.
\end{lemma}
\begin{figure}[htb]
\centering
\includegraphics[width=0.45\linewidth]{TwoPointsrev.eps}
\caption{$A$ and $D$ belong to a hyperbola with asymptotes $BO$ and $CO$.}
\label{TwoPoints}
\end{figure}
\begin{proof}
Denote by $O$ be the intersection of the proposed asymptotes, by $A_1$ the intersection of the parallel to $CD$ through $A$ with $CO$ and by $D_1$ the intersection of the parallel
to $AB$ through $D$ with $BO$ (see figure \ref{TwoPoints}). We must prove that the area of the parallelograms $AA_1OB$ and $DD_1OC$ area equal. But
$$
Area(AA_1OB)=2Area(ABO), \ \ \ Area(DD_1OC)=2Area(DOC)
$$
and
$$
Area(ABO)=Area(ABC)=Area(DBC)=Area(DOC),
$$
thus proving the lemma.
\end{proof}
\subsection{ The space of convex equal-area polygons}
Denote by ${\mathcal T}_n$ the space of convex equal-area $n$-gons modulo affine equivalence. By taking the coordinates of the vertices of the polygon,
we can consider ${\mathcal T}_n\subset\mathbb{R}^{2n}$. For a polygon to be equal area, conditions \eqref{DefEqualArea} must be satisfied. Since these
conditions are cyclic, one of them is redundant, and thus they amount to $(n-1)$ equations. Also, the affine group is $6$-dimensional and thus we expect ${\mathcal T}_n$ to be a $(n-5)$-dimensional space. In fact, this is proved in \cite{Harel03} for the space of (not necessarily convex) equal-area polygons.
In this section we study the space ${\mathcal T}_n$ in some detail.
We prove here the following proposition:
\begin{proposition}\label{CEAspace}
There exists an open set $U\subset\mathbb{R}^{n-5}$ and $W\subset\partial U$ with injective smooth maps $\phi_1:U\cup W\to{\mathcal T}_n$, $\phi_2:U\cup W\to{\mathcal T}_n$ such that
$Im(\phi_1)\cup Im(\phi_2)={\mathcal T}_n$ and $\phi_1|_W=\phi_2|_W$.
Moreover there exists an open set $U_0\subset U$, $\overline{U_0}\cap W=\emptyset$,
such that $\left.\phi_1\right|_{U_0}=\left.\phi_2\right|_{U_0}$.
\end{proposition}
\begin{proof}
Begin with three non-collinear points $P_1,P_2, P_3$. Since we are working modulo affine equivalence, we may assume they are fixed. Let $\mu=(\mu_{2+\frac{1}{2}},...,\mu_{n-4+\frac{1}{2}})\in\mathbb{R}^{n-5}$.
Since $P_4$ must be in a line parallel to $P_2P_3$ passing through $P_1$, so it is determined by the choice of $\mu_{2+\frac{1}{2}}$. In other words, we use \eqref{RatioDiagonalSide} with $i=2$ to define $P_4$.
Similarly, we determine
$P_5,P_6,....,P_{n-2}$.
Now that we know the first $(n-2)$ vertices $\{P_1,P_2,....,P_{n-2}\}$, we want to find the vertices $P_{n-1}$ and $P_n$ to complete the convex equal-area $n$-gon.
Denote by $r_1$ the line through $P_{n-2}$ and $P_{1}$, by $r_2$ the line parallel to $P_{n-3}P_{n-2}$ passing through $P_{n-4}$ and by $r_3$ the line parallel to $P_1P_2$ passing through $P_3$
(see figure \ref{sides10} and applet \cite{Applets}).
The following five conditions must be satisfied:
$$
P_{n-1}P_n\parallel r_1,\ \ P_{n-1}\in r_2,\ \ P_n\in r_3,\ \ P_{n-2}P_{n-1}\parallel P_{n-3}P_{n},\ \ P_{n-1}P_{2}\parallel P_nP_1.
$$
The fifth condition is redundant, since the $n$ equations \eqref{DefEqualArea} are cyclic. Thus we have to verify only the first four.
According to lemma \ref{Hyperbola}, the first and fourth conditions say that $P_n$ belongs to the hyperbola passing through $P_{n-3}$ with asymptotes $r_1$ and $r_2$.
Thus the points $P_n$ and $P'_n$ that close the convex equal-area polygon
are the intersections of the branch $h$ of this hyperbola that does not contain $P_{n-3}$ with the line $r_3$, when these intersections exist.
\begin{figure}[htb]
\centering
\includegraphics[width=0.6\linewidth]{sides10rev.eps}
\caption{Conditions for the existence of a convex equal-area $n$-gon. }
\label{sides10}
\end{figure}
Let $O=r_1\cap r_2$, $A=r_1\cap r_3$ and $B=r_2\cap r_3$.
It is clear that $A$ must be on the segment
$P_{1}O$.
We consider some different cases: (0) $B$ is in the half-line $AP_3$. In this case the branch of hyperbola $h$ intersects $r_3$ exactly once.
(1) $r_2$ is parallel to $r_3$. In this case the branch of hyperbola $h$ also intersects $r_3$ exactly once.
(2) When $B$ is on the half-line $P_3A$, the branch of hyperbola $h$ touches $r_3$ if and only if the ratio
\begin{equation}\label{AreaCondition}
R(\mu)=\frac{Area(AOB)}{Area(P_1P_2P_3)}
\end{equation}
is greater than or equal to $4$. This is not difficult to verify by arguments similar to the ones used in lemma \ref{Hyperbola}. (2a) When $R(\mu)$ is strictly greater than $4$,
the segment $AB$ intersect $h$ is two points (see figure \ref{sides10}).
(2b) When this ratio is exactly $4$, the segment $AB$ is tangent to $h$ and thus intersect it in one point.
We shall denote by $U_0$, $U_1$, $U_2$ and $W$ the subsets of $\mu\in\mathbb{R}^{n-5}$ such that (0), (1), (2a) or (2b) occurs, respectively.
It is now clear how to define the smooth maps $\phi_1$ and $\phi_2$. For $\mu\in U_0\cup U_1\cup W$, $\phi_1=\phi_2$ correspond to the unique convex equal area polygon
corresponding to $\mu$. For $\mu\in U_2$, let $\phi_1$ and $\phi_2$ correspond to the two convex equal-area polygons associated with $\mu$. To conclude the proof of the proposition,
take $U=U_0\cup U_1\cup U_2$.
\end{proof}
Next lemma says that the affinely regular $n$-gon corresponds to $\mu\in U$:
\begin{lemma}\label{AreasNgon}
Consider an affine regular $n$-gon:
\begin{itemize}
\item For $n=5,6,7,8$, $\mu\in \overline{U_0}$.
\item For $n\geq 9$,
the ratio $R(\mu)$ defined by equation \eqref{AreaCondition}
is strictly greater than $4$ and thus $\mu\in U-\overline{U_0}$.
\end{itemize}
\end{lemma}
\begin{proof}
The first item is immediate. For the second item, one can obtain
$$
R=\frac{(X+Y)^2}{XY},
$$
where $X=\frac{1}{2\cos(\alpha_n)}$, $Y=\frac{1}{\sqrt{2(1+\cos(\alpha_n))}}$ and $\alpha_n=\frac{4\pi}{n}$.
We conclude that $R\geq 4$, and since $\cos{\alpha_n}<1$, the inequality is strict.
\end{proof}
We conjecture that the set ${\mathcal T}_n$ is connected, as suggested by experiments done with GeoGebra.
\subsection{ Approximating a convex curve by a convex equal-area polygon}
In this section we propose an algorithm for approximating a closed convex curve by convex equal-area polygons. Although the resulting polygons may be neither inscribed nor circumscribed,
they are asymptotically close to the inscribed polygon whose vertices are equally spaced with respect to the affine arc-length of the curve. For asymptotic results concerning optimal approximations
of convex curves by inscribed or circumscribed polygons, we refer to \cite{Ludwig94} and \cite{McClure75}. For surveys on approximation of convex curves by polygons,
we refer to \cite{Gruber83} and \cite{Gruber93}.
Given a closed convex planar curve $\gamma$, we consider the following algorithm: Fix any three points $P_1,P_2,P_3$ in a positive orientation at $\gamma$
such that $d(P_1,P_2)=d(P_2,P_3)=\frac{L}{n}$, where $d(\cdot,\cdot)$ denotes affine distance along $\gamma$ and $L$ is the affine perimeter of $\gamma$. Then $P_4$ is obtained as the intersection of a parallel to $P_2P_3$ at $P_1$ with $\gamma$. Proceeding in this way we obtain $P_k$, $k=1,...,m$ at the curve, and we continue until condition (2) in the proof of proposition \ref{CEAspace} holds with $R(\mu)<4$.
Denote by $o(s^k)$ any quantity satisfying
$\lim_{s\to 0}\frac{o(s^k)}{s^{k}}=0$.
\begin{lemma}
Assume that $d(P_{i-1},P_{i})=s$ and $d(P_{i},P_{i+1})=s+o(s^2)$. Then $d(P_{i+1},P_{i+2})=s+o(s^2)$.
\end{lemma}
\begin{proof}
We may assume that $\gamma(s)$ is an affine arc-length parameterization of the curve $\gamma$, $P_{i}=\gamma(0)=(0,0)$, $\gamma'(0)=(1,0)$ and $\gamma''(0)=(0,1)$. Then $\gamma'''(0)=(-\mu(0),0)$ and thus
we write $\gamma(s)=(s,s^2/2)+(o(s^2),o(s^3))$ with $P_{i-1}=\gamma(-s)$, $P_{i+1}=\gamma(s+o(s^2))$. Denoting $\gamma(t)=P_{i+2}$, we must show that $t=2s+o(s^2)$.
Write
\begin{eqnarray*}
P_{i+2}-P_{i-1}&=&\left(t+s+o(s^2),\frac{t^2-s^2}{2}+o(s^3)\right)\\
&=&(t+s)\left( 1+o(s),\frac{t-s}{2}+o(s^2) \right).
\end{eqnarray*}
Also,
\begin{eqnarray*}
P_{i+1}-P_{i}&=&\left( s+o(s^2),\frac{s^2}{2}+o(s^3)\right)\\
&=&s\left(1+o(s),\frac{s}{2}+o(s^2)\right).
\end{eqnarray*}
Since $P_{i+2}-P_{i-1}\parallel P_{i+1}-P_i$, we have
\begin{equation}\label{eq:tequal2s}
(1+o(s))(\frac{s}{2}+o(s^2))=(1+o(s))(\frac{t-s}{2}+o(s^2)).
\end{equation}
Thus $t=2s+o(s)$. Using this result in \eqref{eq:tequal2s} we obtain that in fact \linebreak $t=2s+o(s^2)$, thus proving the lemma.
\end{proof}
\begin{corollary}\label{Cor:UniformlySampling}
For a closed convex curve $\gamma$ with affine length $L$, let $\{P_1,...,P_m\}$ be the trapezoidal polygon obtained by the algorithm described
in the beginning of this section, with $d(P_1,P_2)=d(P_2,P_3)=\frac{L}{n}$.
Denote by $\{P_1,...,{\overline P}_4,...,{\overline P}_n\}$ the affinely uniform sample along $\gamma$, i.e., $d({\overline P}_i,P_1)=(i-1)\frac{L}{n}$. Then
$$
\lim_{n\to\infty}\frac{m}{n}=1
$$
and, for any $1\leq i\leq \min(m,n)$,
$$
\lim_{n\to\infty} d(P_i,{\overline P}_i)=0.
$$
\end{corollary}
\begin{proof}
The above lemma says that $d(P_{i+1},P_i)=s+o(s)$, $1\leq i\leq m$, where $s=\frac{L}{n}$. Thus
$$
d(P_i,P_1)=\frac{(i-1)L}{n}+o(1),
$$
thus proving the corollary.
\end{proof}
Corollary \ref{Cor:UniformlySampling} says that, for $n$ large, the convex equal-area polygon constructed above gives a sampling of the curve $\gamma$ that is approximately uniform with respect to affine arc-length.
It is natural to ask if, given a closed convex curve $\gamma$, a point $P_1$ on it and $n$, there exists a convex equal-area $n$-gon inscribed in $\gamma$ with $P_1$ as a vertex.
We believe that this is true for odd $n$, but not for even $n$. We plan to consider this question in a future work.
\section{ Discrete planar affine geometry and the six sextactic edges theorem}\label{section:DiscreteAffineGeometry}
\subsection{Affine curvature and sextactic edges}
At each vertex, one defines the affine normal vector $\mathbf{n}_i$ by
$$
\mathbf{n}_i=\v_{i+\frac{1}{2}}-\v_{i-\frac{1}{2}}=P_{i-1}+P_{i+1}-2P_{i}.
$$
If we assume that the polygon is convex equal-area, the determinants $[\v_{i+\frac{1}{2}},\mathbf{n}_i]=[\v_{i-\frac{1}{2}},\mathbf{n}_i]$ equal some constant $l$, independent of $i$.
By applying an affine transformation, we may assume that $l=1$, and from now on we shall assume so.
Note that $\mathbf{n}_{i+1}-\mathbf{n}_i$ is parallel to $\v_{i+\frac{1}{2}}$ and
\begin{equation}\label{def:AffineCurvature}
\mathbf{n}_{i+1}-\mathbf{n}_i=-\mu_{i+\frac{1}{2}}\v_{i+\frac{1}{2}},
\end{equation}
where $\mu_{i+\frac{1}{2}}$ is defined by \eqref{RatioDiagonalSide}. We shall call $\mu_{i+\frac{1}{2}}$ the {\it discrete affine curvature} of the edge $\v_{i+\frac{1}{2}}$.
Let $\Delta\mu(i)=\mu_{i+\frac{1}{2}}-\mu_{i-\frac{1}{2}}$. A {\it sextactic edge} is an edge $\v_{i+\frac{1}{2}}$ such that $\Delta\mu(i)\cdot\Delta\mu(i+1)\leq 0$.
\subsection{ The six sextactic edges theorem}
In this section we prove a discrete analog of the six vertices theorem (\cite{Buchin83},\cite{Ovsienko01}). We begin with the following lemma:
\begin{lemma}\label{quadratic}
Consider a convex equal-area polygon $\{P_1,...,P_n\}$, $P_i=(x_i,y_i)$. Then, for any quadratic function $q(x,y)$,
$$
\sum_{i=1}^n\Delta\mu(i)q(x_i,y_i)=0
$$
\end{lemma}
\begin{proof}
The case $q(x,y)=1$ is trivial. Consider now $q(x,y)=x$, the case $q(x,y)=y$ being analogous. Denote $\mathbf{n}_i=(n_i^1,n_i^2)$. A straightforward calculation shows that
\begin{eqnarray*}
\sum_{i=1}^n [\mu_{i+\frac{1}{2}}-\mu_{i-\frac{1}{2}}]\ x_i&=& -\sum_{i=1}^n\mu_{i+\frac{1}{2}}(x_{i+1}-x_i)\\
&=& \sum_{i=1}^n(n_{i+1}^1-n_i^1)=0.
\end{eqnarray*}
Now consider $q(x,y)=x^2$, which is similar to the case $q(x,y)=y^2$.
\begin{eqnarray*}
&&\sum_{i=1}^n [\mu_{i+\frac{1}{2}}-\mu_{i-\frac{1}{2}}]\ x_i^2=-\sum_{i=1}^n\mu_{i+\frac{1}{2}}(x_{i+1}^2-x_i^2)\\
&=& -\sum_{i=1}^n\mu_{i+\frac{1}{2}}(x_{i+1}-x_i)(x_{i+1}+x_i)\\
&=& \sum_{i=1}^n(n_{i+1}^1-n_i^1)(x_{i+1}+x_i)\\
&=& -\sum_{i=1}^n n_i^1\left((x_{i+1}+x_i)-(x_{i}+x_{i-1})\right)\\
&=& -\sum_{i=1}^n\left((x_{i+1}-x_i)-(x_{i}-x_{i-1})\right)\left((x_{i+1}-x_i)+(x_{i}-x_{i-1})\right)\\
&=&-\sum_{i=1}^n\left( (x_{i+1}-x_i)^2-(x_{i}-x_{i-1})^2 \right)=0.
\end{eqnarray*}
Since a rotation of the plane preserves the convex equal-area property we do not need to consider the term in $xy$, and so the lemma is proved.
\end{proof}
\begin{theorem}\label{SextaticEdges}
Any convex equal-area $n$-gon, $n\geq 6$, admits at least six sextactic edges.
\end{theorem}
\begin{proof}
Suppose by contradiction that $\Delta\mu(i)$ changes sign four times or less.
Then there exists a quadratic function $q(x,y)$ that is positive
in a region that contains the vertices where $\Delta\mu(i)$ is positive and negative in a region that contains the vertices where $\Delta\mu(i)$ is
negative. In fact, if there no changes of sign, just take $q=constant$. If there are just two changes of sign, take $q$ to be a linear function
whose zero line divides the vertices with positive $\Delta\mu$ from the vertices with negative $\Delta\mu$. In the case of four changes of sign,
consider lines $l_1$ and $l_2$ passing through edges where $\Delta\mu$ changes sign and whose intersection occurs inside the polygon. Then take
$q$ as the product of linear functions whose zero lines are $l_1$ and $l_2$.
The existence of such a quadratic function contradicts lemma \ref{quadratic} and so the theorem is proved.
\end{proof}
\section{Affine evolutes, parallel polygons and the affine distance symmetry set}\label{section:Evolutas}
\subsection{Parallel polygons and the affine evolute}
The affine normal vectors $\mathbf{n}_i$ generate affine normal lines $P_i(\lambda)=P_i+\lambda\mathbf{n}_i$, $\lambda\in\mathbb{R}$. The polygon $P(\lambda)$ whose vertices are $P_i(\lambda)$ is called the {\it $\lambda$-parallel polygon}.
The edges $\v_{i+\frac{1}{2}}(\lambda)=P_{i+1}(\lambda)-P_i(\lambda)$ of $P(\lambda)$ are parallel to the edges $\v_{i+\frac{1}{2}}$ of $P$, and in fact
\begin{equation}\label{eq:parallel}
\v_{i+\frac{1}{2}}(\lambda)=(1-\lambda\mu_{i+\frac{1}{2}})\ \v_{i+\frac{1}{2}}.
\end{equation}
A vertex $P_i(\lambda)$ of the parallel polygon is a {\it cusp} if the determinants $[\v_{i-\frac{1}{2}},\v_{i+\frac{1}{2}}]$ and $[\v_{i-\frac{1}{2}}(\lambda),\v_{i+\frac{1}{2}}(\lambda)]$
have different signs. By equation \eqref{eq:parallel}, this is equivalent to
\begin{equation}\label{eq:parallelcusp}
(1-\lambda\mu_{i-\frac{1}{2}})\cdot(1-\lambda\mu_{i+\frac{1}{2}})<0.
\end{equation}
This condition is equivalent to
$\v_{i-\frac{1}{2}}(\lambda)$ and $\v_{i+\frac{1}{2}}(\lambda)$ being at the same side relative to the normal line $P_i(\lambda),\ \lambda\in\mathbb{R}$ (see figure \ref{ParallelCusp}).
\begin{figure}[htb]
\centering
\includegraphics[width=0.7\linewidth]{ParallelCusp.eps}
\caption{A cusp of a parallel at the normal line at $P_3$. The segment $\mathbf{e}_3$ is part of the affine evolute.}
\label{ParallelCusp}
\end{figure}
For each pair $(i,i+1)$, define the node $Q_{i+\frac{1}{2}}$ as the intersection of the normal lines at $P_i$ and $P_{i+1}$. Then
$$
Q_{i+\frac{1}{2}}=P_i+\frac{1}{\mu_{i+\frac{1}{2}}}\mathbf{n}_i=P_{i+1}+\frac{1}{\mu_{i+\frac{1}{2}}}\mathbf{n}_{i+1}.
$$
Let $\mathbf{e}_i$ denote the subset of the normal line $P_i(\lambda)$ with boundary $\{Q_{i-\frac{1}{2}},Q_{i+\frac{1}{2}}\}$ that does not contain $P_i$. Observe that if $\mu_{i-\frac{1}{2}}\mu_{i+\frac{1}{2}}>0$, then $\mathbf{e}_i$
is the segment $\overline{Q_{i-\frac{1}{2}}Q_{i+\frac{1}{2}}}$, while if $\mu_{i-\frac{1}{2}}\mu_{i+\frac{1}{2}}<0$, then $\mathbf{e}_i$ is the complement of this segment. If $\mu_{i-\frac{1}{2}}$ or $\mu_{i+\frac{1}{2}}$ is zero, then $\mathbf{e}_i$
is a half-line. The graph with nodes $Q_{i+\frac{1}{2}}$ and edges $\mathbf{e}_i$ is called the {\it affine evolute} of $P$. Observe that the affine evolute is a bounded polygon if and only if $\mu_{i+\frac{1}{2}}>0$, for any $1\leq i\leq n$.
From equation \eqref{eq:parallelcusp} one can easily prove the following proposition, which is well-known for smooth curves:
\begin{proposition}\label{ParallelEvolute}
Every cusp of a parallel belong to the affine evolute.
\end{proposition}
The following proposition suggests that affinely regular polygons are the discrete counterparts of ellipses:
\begin{proposition}\label{PropPolRegular}
The affine evolute of a convex equal-area polygon reduces to a point if and only if the polygon is affinely regular.
\end{proposition}
\begin{proof}
If the polygon is affinely regular, then it is obvious that the affine evolute reduces to a point. Now suppose that the affine evolute of $P$ reduces to a point.
Since $Q_{i-\frac{1}{2}}=Q_{i+\frac{1}{2}}$, we obtain
$$
P_i+\frac{1}{\mu_{i+\frac{1}{2}}}\mathbf{n}_i=P_i+\frac{1}{\mu_{i-\frac{1}{2}}}\mathbf{n}_i.
$$
Thus $\mu_{i+\frac{1}{2}}=\mu_{i-\frac{1}{2}}$. Since this holds for any $i$, we conclude that $\mu_{i+\frac{1}{2}}$ is independent of $i$. Now theorem 2 of \cite{Harel03} implies that $P$ is affinely regular.
\end{proof}
\subsection{ Cusps of the affine evolute}
We now give an orientation to the edges of the affine evolute.
The orientation of $\mathbf{e}_i$ is defined as the orientation of $\mathbf{n}_i$.
A {\it cusp} of the affine evolute is a node $Q_{i+\frac{1}{2}}$ whose adjacent pair of edges is not coherently oriented (see figure \ref{EvolutaAfim2}).
\begin{figure}[htb]
\centering
\includegraphics[width=0.7\linewidth]{EvolutaAfim2.eps}
\caption{An oriented affine evolute. The filled dots are its cusps, while the unfilled dots are its ordinary vertices.}
\label{EvolutaAfim2}
\end{figure}
\begin{proposition}\label{SextaticCusps}
$Q_{i+\frac{1}{2}}$ is a cusp of the affine evolute if and only if $\v_{i+\frac{1}{2}}$ is a sextactic edge of the polygon.
\end{proposition}
\begin{proof}
We shall assume $\mu_{i-\frac{1}{2}}>0$, $\mu_{i+\frac{1}{2}}>0$ and $\mu_{i+\frac{3}{2}}>0$, the other cases being analogous. Since
\begin{equation*}
Q_{i+\frac{1}{2}}-Q_{i-\frac{1}{2}}=\frac{\Delta\mu(i)}{\mu_{i-\frac{1}{2}}\mu_{i+\frac{1}{2}}}\mathbf{n}_i,
\end{equation*}
$\mathbf{e}_i$ is oriented coherently with $Q_{i+\frac{1}{2}}-Q_{i-\frac{1}{2}}$ if and only if $\Delta\mu(i)>0$. The same holds for $\mathbf{e}_{i+1}$ and thus
$Q_{i+\frac{1}{2}}$ is a cusp if and only if $\Delta\mu(i)\Delta\mu(i+1)<0$, thus proving the proposition.
\end{proof}
\begin{figure}[htb]
\centering
\includegraphics[width=0.6\linewidth]{EvolutaAfim4.eps}
\caption{Correspondence between the cusp $Q_{3+\frac{1}{2}}$ of the affine evolute (big dot) and the sextactic edge $\v_{3+\frac{1}{2}}$ of the polygon. In the figure, $\mu_{3+\frac{1}{2}}>\mu_{2+\frac{1}{2}}>0$ and $\mu_{3+\frac{1}{2}}>\mu_{4+\frac{1}{2}}>0$.}
\label{EvolutaAfim4}
\end{figure}
Following \cite{Tabach00}, the normal lines $P_i(\lambda),\ \lambda\in\mathbb{R}$ form an exact system, which means that the parallel polygons are closed. Then the discrete four vertex theorem
of \cite{Tabach00} says that the affine evolute admits at least four cusps.
The following corollary, which is a direct consequence of proposition \ref{SextaticCusps} and theorem \ref{SextaticEdges}, says that, in the context of convex equal-area polygons, the affine evolute has at least six cusps.
\begin{corollary}\label{cor:evolutecusps}
The affine evolute of a convex equal-area polygon has at least six cusps.
\end{corollary}
\subsection{ Affine distance symmetry set }
The set of self-intersections of the parallel curves is called {\it affine distance symmetry set} (ADSS). More precisely, the ADSS is the closure of the set of points $X$ such that
there exist two edges
$\v_{i+\frac{1}{2}}$ and $\v_{j+\frac{1}{2}}$, $|i-j|\geq 2$, and $\lambda\in\mathbb{R}$ such that the edges $\v_{i+\frac{1}{2}}(\lambda)$ and $\v_{j+\frac{1}{2}}(\lambda)$ of $P(\lambda)$ intersect
at $X$ (see figure \ref{ADSS} and applet \cite{Applets}).
\begin{figure}[htb]
\centering
\includegraphics[width=0.8\linewidth]{ADSS.eps}
\caption{The dashed line is one branch of the ADSS. Observe the endpoints of the ADSS (round dots) at the cusps of the affine evolute and the cusps of the ADSS (square dots) at the affine evolute.
This ADSS has two other branches.}
\label{ADSS}
\end{figure}
Denote by $\mathbf{h}(i+\frac{1}{2},j+\frac{1}{2})$ the points of the ADSS associated with the edges
$\v_{i+\frac{1}{2}}$ and $\v_{j+\frac{1}{2}}$.
Since a point of $\mathbf{h}(i+\frac{1}{2},j+\frac{1}{2})$ is affinely equidistant to the edges $\v_{i+\frac{1}{2}}$ and $\v_{j+\frac{1}{2}}$, one can verify that $\mathbf{h}(i+\frac{1}{2},j+\frac{1}{2})$ is a
segment contained in the line through the intersection
of the sides $\v_{i+\frac{1}{2}}$ and $\v_{j+\frac{1}{2}}$ with direction given by $\v_{i+\frac{1}{2}}-\v_{j+\frac{1}{2}}$. Moreover, the endpoints of $\mathbf{h}(i+\frac{1}{2},j+\frac{1}{2})$ are points of the normal lines
at $P_i,P_{i+1},P_j$ or $P_{j+1}$ (see figure \ref{ADSSedge}). It is possible to give a coherent orientation for $\mathbf{h}(i+\frac{1}{2},j+\frac{1}{2})$, but we shall not need it in this paper.
\begin{figure}[htb]
\centering
\includegraphics[width=0.5\linewidth]{ADSSedge.eps}
\caption{The dashed line is the parallel to the vector $P_6-P_7+P_2-P_1$ through the intersection of $P_1P_2$ and $P_6P_7$. This line is not the normal at $P_8$, although, by the equal-area property, $P_8$ belongs to it.
The segment $\mathbf{h}(1+\frac{1}{2},6+\frac{1}{2})$ have endpoints at the normal lines at $P_7$ and $P_2$. }
\label{ADSSedge}
\end{figure}
Let $M$ be a node of the ADSS. Note that $M$ must belong to a normal line at some vertex $P_i$. More specifically, either $M$ is a common vertex of $\mathbf{h}(i-\frac{1}{2},j+\frac{1}{2})$ and $\mathbf{h}(i+\frac{1}{2},j+\frac{1}{2})$,
in which case we denote it by $M_{i,j+\frac{1}{2}}$ or M is connected to only one edge $\mathbf{h}(i-\frac{1}{2},i+\frac{3}{2})$, in which case we denote it simply by $M_{i+\frac{1}{2}}$. In the latter case, we say that M is an {\it endpoint} of the ADSS (see figure \ref{EndPoints}). The following proposition is a discrete counterpart of a well-known result of smooth curves:
\begin{figure}[htb]
\centering \hspace*{\fill} \subfigure[ An endpoint of the ADSS corresponding to a cusp of the affine evolute.] {
\includegraphics[width=.45
\linewidth,clip =false]{EndPoints1.eps}} \hspace*{\fill}\subfigure[
An ordinary vertex of the affine evolute that does not belong to the ADSS. ] {
\includegraphics[width=.48\linewidth,clip
=false]{EndPoints2.eps}}\hspace*{\fill}
\caption{Endpoints of the ADSS corresponds to cusps of the affine evolute.}
\label{EndPoints}
\end{figure}
\begin{proposition}
$M_{i+\frac{1}{2}}$ is an endpoint of the ADSS if and only if $M_{i+\frac{1}{2}}$ is a cusp of the affine evolute.
\end{proposition}
\begin{proof}
$M_{i+\frac{1}{2}}$ is an endpoint of the ADSS if it is the limit of cusps of the parallels $P(\lambda)$ at the normal lines $P_i(\lambda),\ \lambda\in\mathbb{R}$ and $P_{i+1}(\lambda),\ \lambda\in\mathbb{R}$, with
$\lambda$ converging to $(\mu_{i+\frac{1}{2}})^{-1}$.
It is now easy to see that this occurs if and only if $M_{i+\frac{1}{2}}$ is a cusp of the affine evolute.
\end{proof}
As a consequence of the above proposition and corollary \ref{cor:evolutecusps}, we conclude that any ADSS has at least three branches.
\bigskip
Take now a node $M_{i,j+\frac{1}{2}}$ of the ADSS that is not an endpoint.
Then $M_{i,j+\frac{1}{2}}$ is at the boundary of $\v_{i-\frac{1}{2}}(\lambda)$
and $\v_{i+\frac{1}{2}}(\lambda)$, for a certain $\lambda$, and also belongs to $\v_{j+\frac{1}{2}}(\lambda)$. We say that $M_{i,j+\frac{1}{2}}$ is a {\it cusp} of the ADSS if it is a cusp of the
parallel $P(\lambda)$, which is equivalent to say that the edges $\v_{i-\frac{1}{2}}(\lambda)$
and $\v_{i+\frac{1}{2}}(\lambda)$ are on the same side of the normal line at $P_i$ (see figure \ref{ADSSZoom}). It follows then from proposition \ref{ParallelEvolute} that a node of the
ADSS is a cusp if and only if belongs to the affine evolute.
\begin{figure}[htb]
\centering
\includegraphics[width=0.4\linewidth]{ADSSZoom.eps}
\caption{A zoom of figure \ref{ADSS}. Observe that, in the neighborhood of a cusp $M$ of the ADSS, both $\mathbf{h}(i+\frac{1}{2},j+\frac{1}{2})$ and $\mathbf{h}(i-\frac{1}{2},j+\frac{1}{2})$ (dashed)
are on the same side of the normal at $P_i$ (dotted) as $\v_{i+\frac{1}{2}}(\lambda)$ and $\v_{i-\frac{1}{2}}(\lambda)$ (black). The remaining line in grey is $\v_{j+\frac{1}{2}}(\lambda)$, which crosses
the normal line.}
\label{ADSSZoom}
\end{figure}
\section{An affine isoperimetric inequality}\label{section:Isoperimetric}
Denote by $A=A_0$ the area bounded by the convex equal-area $n$-gon $P$. Assuming that $[\v_{i-\frac{1}{2}},\mathbf{n}_i]=[\v_{i+\frac{1}{2}},\mathbf{n}_i]=1$, the affine perimeter is defined as $L=n$. The
following inequality is a discrete counterpart of inequality \eqref{SmoothIsoIneq}.
\begin{theorem}\label{isoperimetric} The following isoperimetric inequality holds,
$$
\sum_{i=1}^n \mu_{i+\frac{1}{2}}\leq \frac{L^2}{2A},
$$
with equality if and only if $P$ is affinely regular.
\end{theorem}
The proof of this theorem is based on the inequality of Minkowski for mixed areas (\cite{Flanders68,Schneider93}).
Define the mixed area of two parallel $n$-gons $P$ and $P'$ by
$$
A(P,P')=\frac{1}{2}\sum_{i=1}^n[P'_i,P_{i+1}-P_i].
$$
We remark that $A(P,P')=A(P',P)$, since
\begin{eqnarray*}
\sum_{i=1}^n[P_i,P'_{i+1}-P'_i]&=&\sum_{i=1}^n[P'_{i+1},P_{i+1}-P_{i}]\\
&=&\sum_{i=1}^n[P'_i,P_{i+1}-P_{i}],
\end{eqnarray*}
where we have used the parallelism of $P_iP_{i+1}$ with $P'_{i}P'_{i+1}$ in the last equality. When $P=P$ we write $A(P)=A(P,P)$.
Since, for small $\lambda$, $P+\lambda\mathbf{n}$ is convex, we can apply the Minkowski inequality for the convex parallel polygons $P$ and $P+\lambda\mathbf{n}$, which says that
$$
A(P,P+\lambda\mathbf{n})^2\geq A(P)A(P+\lambda\mathbf{n}),
$$
with equality only in case $P$ and $P+\lambda\mathbf{n}$ are homothetic (see \cite{Schneider93}, p.321, note 1). Now
\begin{equation*}
A(P+\lambda\mathbf{n})=A(P)+2\lambda A(P,\mathbf{n})+\lambda^2 A(\mathbf{n})
\end{equation*}
and
\begin{equation*}
A(P,P+\lambda\mathbf{n})=A(P)+\lambda A(P,\mathbf{n}).
\end{equation*}
Thus we conclude that
\begin{equation*}
A(P,\mathbf{n})^2\geq A(P)A(\mathbf{n}),
\end{equation*}
which is the Minkowski inequality for $P$ and the possibly non-convex polygon $\mathbf{n}$.
We can now complete the proof of theorem \ref{isoperimetric}.
We have that
$$
A(\mathbf{n})=\frac{1}{2}\sum [\mathbf{n}_i,\mathbf{n}_{i+1}-\mathbf{n}_i]=\frac{1}{2}\sum \mu_{i+\frac{1}{2}}[\v_{i+\frac{1}{2}},\mathbf{n}_i]=\frac{1}{2}\sum \mu_{i+\frac{1}{2}}.
$$
On the other hand
$$
A(P,\mathbf{n})=\frac{1}{2}\sum_{i=1}^n[\mathbf{n}_i,P_{i+1}-P_i]=-\frac{n}{2}=-\frac{L}{2}.
$$
Since $A(P)=A$, the Minkowski inequality for $P$ and $\mathbf{n}$ implies that
$$
\frac{L^2}{4}\geq A\frac{1}{2}\sum \mu_{i+\frac{1}{2}},
$$
which proves the isoperimetric inequality. Equality holds if and only if $P$ and $P+\lambda\mathbf{n}$ are homothetic, which is equivalent to the affine evolute of $P$ being a single point.
By proposition \ref{PropPolRegular}, this fact occurs if and only if the polygon $P$ is affinely regular.
|
\section{Introduction}
In this paper we consider the simplest of electrical networks, namely those that consist of only resistors. The electrical properties of such a network $N$ are completely described by the {\it response matrix} $L(N)$, which computes the current that flows through the network when certain voltages are fixed at the boundary vertices of $N$. The study of the response matrices of {\it planar} electrical networks has led to a robust theory, see Curtis-Ingerman-Morrow \cite{CIM}, or de Verdi\`{e}re-Gitler-Vertigan \cite{dVGV}. In 1899, Kennelly \cite{Ken} described a local transformation (see Figure \ref{fig:elec11}) of a network, called the {\it star-triangle} or {\it $Y-\Delta$} transformation, which preserves the response matrix of a network. This transformation is one of the many places where a Yang-Baxter style transformation occurs in mathematics or physics.
\medskip
Curtis-Ingerman-Morrow \cite{CIM} studied the operations of {\it adjoining a boundary spike} and {\it adjoining a boundary edge} to (planar) electrical networks.
Our point of departure is to consider these operations as one-parameter subgroups of a Lie group action. The star-triangle transformation then leads to an ``electrical Serre relation'' in the corresponding Lie algebra, which turns out to be a deformation of the Chevalley-Serre relation for $\ssl_{n}$:
$$
\text{Serre relation:}\; [e,[e,e']] = 0 \qquad \text{electrical Serre relation:} \;[e,[e,e']] = -2e
$$
The corresponding one-parameter subgroups satisfy a Yang-Baxter style relation which is a deformation of Lusztig's relation in total positivity:
\begin{align*}
\text{Lusztig's relation:} &\;u_i(a) u_j(b) u_i(c) = u_j({bc}/({a+c})) u_i(a+c) u_j({ab}/({a+c})) \\
\text{electrical relation:} &\; u_i(a) u_j(b) u_i(c) = u_j({bc}/({a+c+abc})) u_i(a+c+abc) u_j({ab}/({a+c+abc})).
\end{align*}
We study in detail the Lie algebra $\el_{2n}$ with $2n$ generators satisfying the electrical Serre relation. An (electrically) nonnegative part $(EL_{2n})_{\geq 0}$ of the
corresponding Lie group $EL_{2n}$ acts on planar electrical networks with $n+1$ boundary vertices, and one obtains a dense subset of all response matrices of planar networks
in this way. This nonnegative part is a rather precise analogue of the totally nonnegative subsemigroup $(U_{2n+1})_{\geq 0}$ of the unipotent subgroup of $SL_{2n+1}$, studied in Lie-theoretic terms by Lusztig \cite{Lus}. We show (Proposition \ref{P:decomp}) that $(EL_{2n})_{\geq 0}$ has a cell decomposition labeled by permutations $w \in S_{2n+1}$, precisely paralleling one of Lusztig's results for $(U_{2n})_{\geq 0}$ and reminiscent of the Bruhat decomposition. This can be considered an algebraic analogue of parametrization results in the theory of electrical networks \cite{CIM,dVGV}. Surprisingly, $EL_{2n}$ itself is isomorphic to the symplectic Lie group $Sp_{2n}(\R)$ (Theorem \ref{thm:elsp}). This semisimplicity does not in general hold for electrical Lie groups: $EL_{2n+1}$ is not semisimple. We also describe (Theorem \ref{thm:stab}) the stabilizer lie algebra of the infinitesimal action of $\el_{2n}$ on electrical networks.
While we focus on the planar case in this paper, we shall connect the results of this paper to the inverse problem for electrical networks on cylinders in a future paper.
There we borrow ideas from representation theory such as that of crystals and $R$-matrices.
\medskip
One obtains the types $B$ and $G$ electrical Serre relations by the standard technique of ``folding'' the type $A$ electrical Serre relation.
These lead to a new species of electrical Lie algebras $\e_D$ defined for any Dynkin diagram $D$. Besides the above results for type $A$, there appear to be other interesting
relations between these $\e_D$ and simple Lie algebras. For example, $\eb_2 := \e_{B_2}$ is isomorphic to $\gl_2$. We conjecture (Conjecture \ref{con:1}) that for a finite
type diagram $D$, the dimension $\dim(\e_D)$ is always equal to the dimension of the maximal nilpotent subalgebra of the semisimple Lie algebra ${\mathfrak g}_D$ with Dynkin
diagram $D$, and furthermore (Conjecture \ref{con:2}) that $\e_D$ is finite dimensional if and only if ${\mathfrak g}_D$ is finite-dimensional. We give the electrical analogue
of Lusztig's relation for type $B$ (see Berenstein-Zelevinsky \cite{BZ}) where we observe similar positivity to the totally nonnegative case.
The most interesting case beyond finite Dynkin types is affine type $A$. The corresponding electrical Lie (semi)-group action on planar electrical networks is perhaps even more natural than that of $EL_{2n}$, since one can obtain {\it {all}} (rather than just a dense subset of) response matrices of planar networks in this way (\cite{dVGV, CIM}), and furthermore the circular symmetry of the planar networks is preserved. We however do not attempt to address this case in the current paper.
\section{Electrical networks}
For more background on electrical networks, we refer the reader to \cite{CIM, dVGV}.
\subsection{Response matrix}
For our purposes, an electrical network is a finite weighted undirected graph $N$, where the vertex set is divided into the {\it boundary} vertices and the {\it interior} vertices. The weight $w(e)$ of an edge is to be thought of as the conductance of the corresponding resistor, and is generally taken to be a positive real number. A $0$-weighted edge would be the same as having no edge, and an edge with infinite weight would be the same as identifying the endpoint vertices.
We define the {\it Kirchhoff matrix} $K(N)$ to be the square matrix with rows and columns labeled by the vertices of $N$ as follows:
$$
K_{ij} = \begin{cases} -\sum_{e \; \text{joins $i$ and $j$}} w(e) &\mbox{for $i \neq j$} \\
\sum_{e \; \text{incident to $i$}} w(e) &\mbox{for $i = j$.}
\end{cases}
$$
We define the {\it response matrix} $L(N)$ to be the square matrix with rows and columns labeled by the boundary vertices of $N$, given by the Schur-complement
$$
L(N) = K/K_I
$$
where $K_I$ denotes the submatrix of $K$ indexed by the interior vertices. The response matrix encodes all the electrical properties of $N$ that can be measured from the boundary vertices.
\subsection{Planar electrical networks}
We shall usually consider electrical networks $N$ embedded into a disk, so that the boundary vertices, numbered $1,2,\ldots,n+1$, lie on the boundary of the disk.
\begin{figure}[h!]
\begin{center}
\input{eLie3.pstex_t}
\end{center}
\caption{The operations $v_i(t)$ acting on a network $N$.}
\label{fig:eLie3}
\end{figure}
Given an odd integer $2k-1$, for $k = 1,2,\ldots,n+1$, and a nonnegative real number $t$, we define $v_{2k-1}(t)(N)$ to be the electrical network obtained from $N$ by adding a new edge from a new vertex $v$ to $k$, with weight $1/t$, followed by treating $k$ as an interior vertex, and the new vertex $v$ as a boundary vertex (now named $k$).
Given an even integer $2k$, for $k = 1,2,\ldots,n+1$, and a nonnegative real number $t$, we define $v_{2k}(t)(N)$ to be the electrical network obtained from $N$ by adding a new edge
from $k$ to $k+1$ (indices taken modulo $n+1$), with weight $t$.
The two operations are shown in Figure \ref{fig:eLie3}. In \cite{CIM}, these operations are called {\it adjoining a boundary spike}, and {\it adjoining a boundary edge} respectively.
Our notation suggests, and we shall establish, that there is some symmetry between these two types of operations.
In \cite[Section 8]{CIM}, it is shown that $L(v_i(a) \cdot N)$ depends only on $L(N)$, giving an operation $v_i(t)$ on response matrices. Denote by $x_{ij}$ the entries of the response matrix, where
$1 \leq i , j \leq n+1$. In particular, we have $x_{ij} = x_{ji}$. Then if $\delta_{ij}$ denotes the Kronecker delta, we have
\begin{align} \label{eq:cim}
v_{2k-1}(t): x_{ij} \mapsto x_{ij} -\frac{t x_{ik} x_{kj}}{tx_{kk}+1} \;\;\; \text{and} \;\;\; v_{2k}(t): x_{ij} \mapsto x_{ij} + (\delta_{ik}-\delta_{(i+1)k}) (\delta_{jk}-\delta_{(j+1)k}) t.
\end{align}
We caution that the parameter $\xi$ in \cite{CIM} is the inverse of our $t$ in the odd case.
\subsection{Electrically-equivalent transformations of networks} \label{ss:trans}
\label{sec:trans}
Series-parallel transformations are shown in Figure \ref{fig:eLie1}. The following proposition is well-known and can be found for example in \cite{dVGV}.
\begin{figure}[h!]
\begin{center}
\input{eLie1.pstex_t}
\end{center}
\caption{Series-parallel transformations.}
\label{fig:eLie1}
\end{figure}
\begin{prop}\label{P:SP}
Series-parallel transformations, removing loops, and removing interior degree 1 vertices, do not change the response matrix of a network.
\end{prop}
\begin{figure}[h!]
\begin{center}
\input{eLie2.pstex_t}
\end{center}
\caption{The $Y-\Delta$, or star-triangle, transformation.}
\label{fig:elec11}
\end{figure}
The following theorem is attributed to Kennelly \cite{Ken}.
\begin{prop}[$Y-\Delta$ transformation] \label{P:YDelta}
Assume that parameters $a$,$b$,$c$ and $A$,$B$,$C$ are related by
$$A = \frac{bc}{a+b+c}, \;\; B = \frac{ac}{a+b+c}, \;\; C= \frac{ab}{a+b+c},$$
or equivalently by
$$a = \frac{AB+AC+BC}{A}, \;\; b= \frac{AB+AC+BC}{B}, \;\; c = \frac{AB+AC+BC}{C}.$$
Then switching a local part of an electrical network between the two options shown in Figure \ref{fig:elec11} does not change the response matrix of the whole network.
\end{prop}
\section{The electrical Lie algebra $\el_{2n}$}
\label{sec:el}
Let $\el_{2n}$ be the Lie algebra ${\mathfrak el}_{2n}$ generated over $\R$ by $e_i$, $i=1, \ldots, 2n$ subject to relations
\begin{itemize}
\item $[e_i, e_j] = 0$ for $|i-j|>1$;
\item $[e_i, [e_i, e_j]] = -2 e_i$, $|i-j|=1$.
\end{itemize}
\begin{theorem} \label{thm:elsp}
The Lie algebra $\el_{2n}$ is isomorphic to the real symplectic Lie algebra $\sp_{2n}$.
\end{theorem}
\begin{proof}
We identify the symplectic algebra $\sp_{2n}$ with the space of $2n \times 2n$ matrices which in block form
$$
\left(\begin{array}{cc} A & B \\ C & D \end{array} \right)
$$
satisfy $A = -D^T$, $B = B^T$ and $C = C^T$ (see \cite[Lecture 16]{FH}). Note that the subLie algebra consisting of the matrices where $B = C = 0$ is naturally isomorphic to $\gl_n$.
Let $\ep_i \in \R^n$ denote the standard basis column vector with a $1$ in the $i$-th position. Let $a_1 = \ep_1, a_2 = \ep_1 + \ep_2,a_3 = \ep_2 + \ep_3, \ldots,a_n = \ep_{n-1} + \ep_n$, and let $b_1 = \ep_1, b_2 = \ep_2,\ldots, b_n = \ep_n$. For $1 \leq i \leq n$, define elements of $\sp_{2n}$ by the formulae
$$
e_{2i-1} = \left(\begin{array}{cc} 0 & a_i \cdot a_i^T \\ 0 & 0 \end{array} \right) \qquad e_{2i} = \left(\begin{array}{cc} 0 & 0 \\ b_i \cdot b_i^T & 0 \end{array} \right).
$$
We claim that this gives a symplectic representation $\phi$ of $\el_{2n}$ which is an isomorphism of $\el_{2n}$ with $\sp_{2n}$. We first check that the relations of $\el_{2n}$ are satisfied. It is clear from the block matrix form that $[e_{2i-1},e_{2j-1}]=0=[e_{2i},e_{2j}]$ for any $i,j$. Now,
$$
[e_{2i-1},e_{2j}] = \left(\begin{array}{cc}(a_i \cdot a_i^T) (b_j \cdot b_j^T) & 0\\ 0 & -(b_j \cdot b_j^T)(a_i \cdot a_i^T) \end{array} \right).
$$
But by construction, $b_j^T \cdot a_i= 0 = a_i^T \cdot b_j$ unless $j = i$, or $j = i-1$. Thus $[e_k,e_l] = 0$ unless $|k-l|= 1$. Finally,
$$
[e_{2i-1},[e_{2i-1},e_{2i}]] = \left(\begin{array}{cc}0 &-2(a_i \cdot a_i^T)(b_i \cdot b_i^T)(a_i \cdot a_i^T)\\ 0 & 0 \end{array} \right) = \left(\begin{array}{cc}0 &-2 a_i \cdot a_i^T\\ 0 & 0 \end{array} \right) = -2 e_{2i-1}
$$
using the fact that $a_i^T \cdot b_i = 1 = b_i^T \cdot a_i$. Similarly, one obtains $[e_{k},[e_{k},e_{k \pm1}]]= -2e_k$. Thus we have a symplectic representation $\phi: \el_{2n} \to \sp_{2n}$.
Now we show that this map is surjective. First we verify that the $
\gl_n \subset \phi(\el_{2n})$. The non-zero commutators $[e_i,e_j]$ gives matrices of the form $\left(\begin{array}{cc} A & 0 \\ 0&-A^T \end{array}\right)$ where $A$ is a scalar multiple of one of the matrices $E_{1,1}$, or $E_{i,i}+E_{i+1,i}$ or $E_{i+1,i}+E_{i+1,i+1}$. Here $E_{i,j}$ denotes the $n \times n$ matrix with a single non-zero entry equal to one in the $(i,j)$-th position. All the matrices of the above form occur. It is easy to see that $\phi(\el_{2n})$ must then contain the matrices where $A= E_{i,i}$, $E_{i,i+1}$, or $E_{i+1,i}$ for each $i$. But these matrices generate $\gl_n$ as a Lie algebra. However, it is known \cite[Proposition 8.4(d)]{Hum} that for a semisimple Lie algebra $L$, one has $[L_{\alpha},L_{\beta}] = L_{\alpha+\beta}$ when $\alpha,\beta,\alpha+\beta$ are all roots, and $L_\alpha$ denotes a root subspace. It follows easily from the explicit description of the root system of $\sp_{2n}$ and the definition of $e_i$ that every root subspace of $\sp_{2n}$ is contained in $\phi(\el_{2n})$, completing the proof.
To see that the map $\phi:\el_{2n} \to \sp_{2n}$ is injective, we note:
\begin{lem}\label{lem:dimA}
The dimension of $\el_{2n}$ is $n(2n+1)$.
\end{lem}
\noindent {\it Proof.} According to Lemma \ref{lem:dim} the dimension of $\el_{2n}$ is at most $n(2n+1)$. On the other hand, we just saw that the map $\el_{2n} \to \sp_{2n}$ is surjective. The statement follows.
\end{proof}
\section{The electrical Lie group $EL_{2n}$}
Let $EL_{2n}$ be the simply-connected (real) Lie group with Lie algebra $\el_{2n}$. Let $u_i(a) = \exp(a e_i)$ denote the the one-parameter subgroups corresponding to the generators of $\el_{2n}$.
\begin{theorem} \label{thm:rels}
The elements $u_i(a)$ satisfy the relations
\begin{enumerate}
\item $u_i(a) u_i(b) = u_i(a+b)$,
\item $u_i(a) u_j(b) = u_j(b) u_i(a)$, if $|i-j|>1$;
\item $u_i(a) u_j(b) u_i(c) = u_j({bc}/({a+c+abc})) u_i(a+c+abc) u_j({ab}/({a+c+abc}))$, if $|i-j|=1$.
\end{enumerate}
\end{theorem}
\begin{proof} The first two relations are clear. For the third, observe that for each $1 \leq i \leq 2i-1$, the elements $e_i$, $e_{i+1}$, and $[e_i,e_{i+1}]$ are the usual Chevalley generators of a Lie subalgebra isomorphic to $\ssl_2$. Thus we can verify the relation inside $SL_2(\R)$:
$$\left(\begin{matrix} 1 & a\\ 0 & 1\\ \end{matrix} \right) \left(\begin{matrix} 1 & 0\\ b & 1\\ \end{matrix} \right) \left(\begin{matrix} 1 & c\\ 0 & 1\\ \end{matrix} \right) =
\left(\begin{matrix} 1+ab & a+c+abc\\ b & 1+bc\\ \end{matrix} \right) =$$
$$\left(\begin{matrix} 1 & 0\\ {bc}/({a+c+abc}) & 1\\ \end{matrix} \right) \left(\begin{matrix} 1 & a+c+abc\\ 0 & 1\\ \end{matrix} \right) \left(\begin{matrix} 1 & 0\\ {ab}/({a+c+abc}) & 1\\ \end{matrix} \right).$$
\end{proof}
\begin{remark}
There is a one-parameter deformation which connects the relation Theorem \ref{thm:rels}(3) with Lusztig's relation \cite{Lus} in total positivity:
\begin{equation}\label{E:lus}
u_i(a) u_j(b) u_i(c) = u_j({bc}/({a+c})) u_i(a+c) u_j({ab}/({a+c})).
\end{equation}
Consider the associative algebra $U_{\tau}({\el}_{2n})$ where the generators $\ep_1,\ep_2,\ldots,\ep_{2n}$ satisfy the following deformed Serre relations:
\begin{itemize}
\item $\ep_i \ep_j = \ep_j \ep_i$ if $|i-j|>1$;
\item $\ep_i \ep_j \ep_i = \tau \ep_i + (\ep_i^2 \ep_j + \ep_j \ep_i^2)/2$, if $|i-j|=1$.
\end{itemize}
This is a one-parameter family of algebras which at $\tau=0$ reduces to $U({\mathfrak n^+})$, where ${\mathfrak sl}_{2n+1} = {\mathfrak n^-} \oplus {\mathfrak h} \oplus {\mathfrak n^+}$
is the Cartan
decomposition, while at $\tau=1$ it gives the universal enveloping algebra $U({\mathfrak el}_{2n})$
of the electrical Lie algebra. For $U_{\tau}({\el}_{2n})$, the ``braid move'' for the elements $\exp(a \ep_i)$
then takes the form
$$u_i(a) u_j(b) u_i(c) = u_j({bc}/({a+c+\tau abc})) u_i(a+c+\tau abc) u_j({ab}/({a+c+\tau abc})) \;\;\; \text{if $|i-j|=1$}.$$
At $\tau = 0$ this reduces to \eqref{E:lus}, and at $\tau=1$ it is the relation Theorem \ref{thm:rels}(3).
\end{remark}
\subsection{Nonnegative part of $EL_{2n}$}
Define the {\it {nonnegative}} part $(EL_{2n})_{\geq 0}$ of $EL_{2n}$, or equivalently the {\it {electrically nonnegative}} part of $Sp_{2n}$, to be the subsemigroup generated by the $u_i(a)$ with nonnegative parameters $a$.
For a reduced word $\i = {i_1} \dotsc {i_\ell}$ of $w \in S_{2n+1}$ denote by $E(w) \subset (EL_{2n})_{\geq 0}$ the image of the map
$$
\psi_\i: (a_1, \ldots, a_\ell) \in \mathbb R_{>0}^\ell \mapsto u_{i_1}(a_1) \dotsc u_{i_\ell}(a_\ell).
$$
It follows from the relations in Theorem \ref{thm:rels} that the set $E(w)$ depends only on $w$, and not on the chosen reduced word.
The following proposition is similar to \cite[Proposition 2.7]{Lus} which gives a decomposition $U_{\geq 0} = \sqcup_w U^w_{\geq 0}$ of the totally nonnegative part of the unipotent group.
\begin{prop}\label{P:decomp}
The sets $E(w)$ are disjoint and cover the whole $(EL_{2n})_{\geq 0}$. Each of the maps $\psi_\i:\mathbb R_{>0}^\ell \to E(w)$ is a homeomorphism.
\end{prop}
\begin{proof}
Using Theorem \ref{thm:rels}, we can rewrite any product of generators $u_i(a)$ by performing braid moves similar to those in the symmetric group $S_{2n+1}$. Any product of the generators can be transformed into a product that corresponds to a reduced word in $S_{2n+1}$, and thus it belongs to one of the $E(w)$.
If the map $\psi_\i$ is not injective for some reduced word $\i$, then we can find two reduced products
$$u_{i_1}(a_1) \dotsc u_{i_\ell}(a_\ell) = u_{i_1}(b_1) \dotsc u_{i_\ell}(b_\ell)$$ for two $\ell$-tuples of positive numbers
such that $a_1 \neq b_1$. Without loss of generality we can assume
$a_1 > b_1$, and thus $$u_{i_1}(a_1-b_1) u_{i_2}(a_2) \dotsc u_{i_\ell}(a_\ell) = u_{i_2}(b_2) \dotsc u_{i_\ell}(b_\ell).$$
This shows that two different $E(w)$-s have non-empty intersection. Thus it suffices to show that the latter is impossible.
Furthermore, it suffices to prove that the top cell corresponding to the longest element $w_0$ does not intersect any other cell. Indeed, if any two cells intersect, by adding some extra factors to both
we can lift it to the top cell intersecting one of the other cells. Assume we have a product of the form
\begin{equation}\label{E:ured}
u = [u_1(a_1)] [u_2(a_2) u_1(a_3)] \dotsc [u_k(a_{\ell-k+1}) \dotsc u_1(a_{\ell})]
\end{equation} where $\ell = {k \choose 2}$. Let $\Phi(u) \in Sp_{2n}$ be the image of $u$ in $Sp_{2n}$ under the natural map $EL_{2n} \to Sp_{2n}$ induced by the map $\phi:\el_{2n} \to \sp_{2n}$ of Theorem \ref{thm:elsp}. We argue that the positive parameters $a_{\ell-k+1}, \ldots, a_\ell$ can be recovered uniquely
from just looking at the $n+k/2$-th row of $\Phi(u)$ for $k$ even, or the $(k+1)/2$-th row for $k$ odd. Furthermore, if exactly one of them is equal to zero, then the same calculation will tell us that.
Indeed, assume $k$ is odd. Then the $n+(k+1)/2$-th entry in the $(k+1)/2$-th row is just equal to $a_{\ell-k+1}$. Next, once we know $a_{\ell-k+1}$,
we can use $(k-1)/2$-th entry of the same row to recover $a_{\ell-k+2}$, after which we can use $n+(k-1)/2$ entry to recover $a_{\ell-k+3}$, and so on. For each step we solve a linear equation
where the parameter we divide by is strictly positive as long as all previous parameters $a_i$ recovered are positive. For example, let $n=2$ and take the product
$$u_1(a_1) u_2(a_2) u_1(a_3) u_3(a_4) u_2(a_5) u_1(a_6) .$$
Then the second row of this product is $(a_4 a_5, 1, a_4 + a_4 a_5 a_6, a_4)$. The last entry tells us $a_4$, then from
the first entry we solve for $a_5$, then from the third entry we solve for $a_6$. The case of even $k$ is similar, the order in which we have to read the entries of the $n+k/2$-th row
in this case is $k/2$-th, $n+k/2$-th, $k/2-1$-th, $n+k/2-1$-th, etc. For example, in the product
$$u_1(a_1) u_2(a_2) u_1(a_3) u_3(a_4) u_2(a_5) u_1(a_6) u_4(a_7) u_3(a_8) u_2(a_9) u_1(a_{10}),$$ the last row is $(a_7 a_8 a_9, a_7, a_7 a_8 + a_7 a_8 a_9 a_{10}, 1+a_7 a_8)$.
By looking at second, last, first and third entries we can solve for $a_7, a_8, a_9$ and $a_{10}$ one after the other.
Now we are ready to complete the argument. Assume that $E(w_0)$ intersects some other $E(v)$. For any reduced word of $w_0$, one can find a subword which is a reduced word for $v$, so that we have
\begin{align*}
u &= [u_1(a_1)] [u_2(a_2) u_1(a_3)] \dotsc [u_k(a_{\ell-k+1}) \dotsc u_1(a_{\ell})] \\ &= [u_1(b_1)] [u_2(b_2) u_1(b_3)] \dotsc [u_k(b_{\ell-k+1}) \dotsc u_1(b_{\ell})],
\end{align*}
where the $a_i$ are all positive, but the $b_i$ are nonnegative, with at least one zero. The above algorithm of recovering the $a_i$'s will arrive at a contradiction. Thus the $E(w_0)$ is disjoint from all other $E(v)$.
It remains to show that $\phi_\i^{-1}$ is continuous for any reduced word. If $\i$ is the reduced word of $w_0$ used in \eqref{E:ured}, this follows from the algorithm above: the $a_i$ depend continuously on the matrix entries of $\Phi(u)$. But any two reduced words are connected by braid and commutation moves, so it follows from the (continuously invertible) formulae in Theorem \ref{thm:rels} that $\phi_\i^{-1}$ is continuous for any reduced word of $w_0$. But for any other $v \in S_{2n+1}$, a reduced word $\j$ for $v$ can be found as an initial subword of some reduced word $\i$ for $w_0$. The map $\phi_\j^{-1}$ can then be expressed as a composition of: right multiplication by a fixed element of $EL_{2n}$, the map $\phi_\i^{-1}$, and the projection from $\R^{\ell(w_0)}_{>0}$ to $\R^{\ell(v)}_{>0}$, all of which are continuous.
\end{proof}
\subsection{Action on electrical networks}
Let $\P(n+1) \subset \R^{(n+1)^2}$ denote the set of response matrices of planar electrical networks with $n+1$ boundary vertices. In \cite{CIM}, it is shown that $\P(n+1)$ is exactly the
set of symmetric, circular totally-nonnegative $(n+1) \times (n+1)$ matrices. In \cite[Th\'{e}or\`{e}me 4]{dVGV}, it is shown that $\P(n+1)$ can be identified with the set of planar
electrical networks with $n+1$ boundary vertices modulo the local transformations of Section \ref{ss:trans}. Let $N_0$ denote the empty network (with $n+1$ boundary vertices) and let $L_0 = L(N_0)$ denote the zero matrix.
\begin{theorem}\label{T:action}
The nonnegative part $(EL_{2n})_{\geq 0}$ of the electrical group acts on $\P(n+1)$ via:
$$
u_i(a) \cdot L(N) = L(v_i(a)(N)).
$$
\end{theorem}
\begin{proof}
For a single generator $u_i(a)$, the stated action is well-defined because it can be described explicitly on the level of response matrices. The formulae for $L( v_i(a)(N))$ in terms of $L(N)$ is given by the equation \eqref{eq:cim}.
We first show that the relations of Theorem \ref{thm:rels} hold for this action. Relation (1) follows from the series-parallel relation for networks. Relation (2) is immediate: the corresponding networks are identical without any transformations. Relation (3) follows from the $Y-\Delta$ transformation (see Example \ref{ex:YDelta}).
But now suppose we have two different expressions for $u \in (EL_{2n})_{\geq 0}$ in terms of generators. Then using Relations (1)-(3) of Theorem \ref{thm:rels}, we may assume that both expressions are products corresponding to a reduced word. By Proposition \ref{P:decomp}, the two products come from reduced words for the same $w \in S_{2n+1}$. It follows that they are related by relations (1)-(3).
\end{proof}
\begin{example}\label{ex:YDelta}
The products $u_3(a) u_4(b) u_3(c)$ and $u_4(\frac{bc}{a+c+abc}) u_3(a+c+abc) u_4(\frac{ab}{a+c+abc})$ act on a network in exactly the same way, as shown in Figure \ref{fig:elec10}.
\begin{figure}[h!]
\begin{center}
\input{eLie4.pstex_t}
\end{center}
\caption{}
\label{fig:elec10}
\end{figure}
\end{example}
A permutation $w = w(1) w(2) \cdots w(2n+1) \in S_{2n+1}$ is {\it efficient} if
\begin{enumerate}
\item
$w(1) < w(3) < \cdots < w(2n+1)$ and
$w(2) < w(4) < \cdots < w(2n)$
\item
$w(1) < w(2)$, $w(3) < w(4)$, $\ldots$, $w(2n-1) < w(2n)$.
\end{enumerate}
Recall the {\it left weak order} of permutations is given by $w \preceq v$ if and only if there is a $u$ so that $uw = v$ and $\ell(v) = \ell(u) + \ell(w)$. It is a standard fact that $w \preceq v$ if and only if whenever $w(i) > w(j)$ and $i < j$ then $v(i) > v(j)$. Thus the set of efficient permutations has a maximum in left weak order, namely $w = 1 (n+2) 2 (n+3) \cdots (n) (2n+1) (n+1)$ with length $n(n-1)/2$.
\begin{thm}\label{T:efficient}
Let $w \in S_{2n+1}$. The map $\Theta_w: E(w) \to \P(n+1)$ given by $\Theta_w(u) = u \cdot L_0$ is injective if and only if $w$ is efficient. We have $\image(\Theta_w) \cap \image(\Theta_v) = \emptyset$ for $w \neq v$ both efficient. If $w$ is not efficient, there is a unique efficient $v$ such that $\image(\Theta_w) = \image(\Theta_v)$.
\end{thm}
\begin{proof}
Let $w^* \in S_{2n+1}$ be the efficient permutation of maximal length. Then a possible reduced word for $w^*$ is $(n+1) \cdots \left(46 \cdots (2n-2)\right)\left(35 \cdots (2n-1)\right) \left(246 \cdots (2n)\right)$. The graph obtained by taking the corresponding $u_i(a)$ and acting on the empty network is exactly the ``standard graph'' of \cite[Section 7]{CIM} or the graph $C_N$ of \cite{dVGV}. In particular, $\Theta_{w^*}$ is injective by \cite[Theorem 2]{CIM} or \cite[Th\'{e}or\`{e}me 3]{dVGV}.
Suppose $w$ is an arbitrary efficient permutation.
Then since $w^* = uw$ for some $u \in S_{2n+1}$, if $\Theta_w$ is not injective then $\Theta_{w^*}$ is not injective as well, so we conclude that $\Theta_w$ is injective.
For a pair $(i,j)$ with $1 \leq i < j \leq n+1$, let us say that a network $N$ is $(i,j)$-{\it {connected}} if we can find a disjoint set of paths $p_1, p_2,...,p_{\lfloor (j-i+1)/2\rfloor}$
so that for each $k$, the path
$p_k$ connects boundary vertex $i+k-1$ to boundary vertex $j-k+1$ without passing through any other boundary vertex. This is a special case of the connections of circular pairs $(P,Q)$
of \cite{CIM}. Let $N_w$ be a graph constructed from some reduced word of an efficient $w$. We observe that $N_w$ is $(i,j)$-connected if and only if $w(2i) > w(2j-1)$.
\begin{figure}[h!]
\begin{center}
\input{eLie5.pstex_t}
\end{center}
\caption{}
\label{fig:eLie5}
\end{figure}
For example, the first network in Figure \ref{fig:eLie5} corresponds to $w = s_5 s_3 s_6 s_4 s_2 = (1,3,5,2,7,4,6)$. We see that it is $(2,4)$-connected, which agrees with
$w(4)=6>5=w(7)$. If the dashed edge is not there, we have $w = s_3 s_6 s_4 s_2 = (1,3,5,2,4,7,6)$ and $w(4)=5<6=w(7)$ in agreement with the network not being $(2,4)$-connected. Similarly,
the second network in Figure \ref{fig:eLie5} corresponds to $w = s_4 s_5 s_3 s_6 s_4 s_2 = (1,3,5,7,2,4,6)$ with the dashed edge and to $w = s_5 s_3 s_6 s_4 s_2 = (1,3,5,2,7,4,6)$ without. In the first case
it is $(1,4)$-connected, in the second it is not, in agreement with relative order of $w(2)$ and $w(7)$.
Note that the inversions $w(2i) > w(2j-1)$ are exactly the possible inversions of an efficient permutation. Since $w$ is determined by its inversions, it follows that the set of $(i,j)$-connections
of $N_w$ determines $w$, and that $\image(\Theta_w) \cap \image(\Theta_v) = \emptyset$ for $w \neq v$ both efficient.
Suppose $w$ is not efficient. Then it is easy to see that $w$ has a reduced expression $s_{i_1}s_{i_2} \cdots s_{i_\ell}$ where either (1) $i_\ell$ is odd, or (2) $i_\ell$ is even
and $i_\ell = i_{\ell-1} \pm 1$. But $u_i(a) \cdot L_0 = L_0$ for odd $i$ since all the boundary vertices are still disconnected in $u_i(a) \cdot N_0$, and $u_{i\pm1}(a) u_{i}(b) \cdot L_0 = u_i(1/a + b) L_0$, using the series-transformation (Proposition \ref{P:SP}). It follows that $\Theta_w$ is not injective. Furthermore, $\image(\Theta_w) = \image(\Theta_v)$, where $v$ is obtained from $w$ by recursively (1) removing $i_\ell$ from a reduced word of $w$ if $i_\ell$ is odd, or (2) changing the last two letters $(i_{\ell}\pm 1) i_\ell$ to $i_\ell$ when $i_\ell$ is even. An efficient $v$ obtained in this way must be unique, since $\image(\Theta_v) \cap \image(\Theta_{v'}) = \emptyset$ for efficient $v \neq v'$.
\end{proof}
\begin{cor}
The number of efficient $w \in S_{2n+1}$ is equal to the Catalan number $C_{n+1} = \frac{1}{n+2}{2n+2 \choose n+1}$.
\end{cor}
\begin{proof}
It is clear that $(2i,2j+1)$ is an inversion of $w$ only if $(2i,2j-1)$ and $(2i+2,2j+1)$ are also inversions, and this characterizes inversion sets of efficient $w$.
Thus the set of efficient $w \in S_{2n+1}$ is in bijection with the lower order ideals of the positive root poset of $\ssl_{n+1}$ which is well known to be enumerated by the
Catalan number \cite{FR}.
\end{proof}
\begin{cor}
$(EL_{2n})_{\geq 0} \cdot L_0$ is dense in $\P(n+1)$.
\end{cor}
\begin{proof}
Follows from Theorem \ref{T:efficient} and \cite[Th\'{e}or\`{e}me 5]{dVGV}.
\end{proof}
\subsection{Stabilizer}
\begin{lem}
The stabilizer subsemigroup of $(EL_{2n})_{\geq 0}$ acting on the zero matrix $L_0$ is the subsemigroup generated by $u_{2i+1}(a)$.
\end{lem}
\begin{proof}
It is clear that $u_{2i+1}(a)$ lies in the stabilizer. But the action of any $u_{2i}(a)$ will change the connectivity of the network, and it is impossible to return to trivial connectivity by adding more edges, or by relations.
\end{proof}
The semigroup stabilizer is too small in the sense that it does not detect the relations $u_{i\pm1}(a) u_{i}(b) \cdot L_0 = u_i(1/a + b) L_0$ used in the proof of Theorem \ref{T:efficient}. We shall calculate the stabilizer subalgebra of the corresponding infinitesimal action of the Lie algebra $\el_{2n}$, which will in particular give an algebraic explanation of these relations. The reason we do not work with the whole Lie group $EL_{2n}$ is threefold: (1) non-positive elements of $EL_{2n}$ will produce networks that are ``virtual'', that is, have negative edge weights; (2) the topology of $EL_{2n}$ means that to obtain an action one cannot just check the relations of Theorem \ref{thm:rels}; (3) when the parameters are non-positive, the relation Theorem \ref{thm:rels}(3) develops singularities.
To describe the infinitesimal action of $\el_{2n}$, we give derivations of $\R[x_{ij}]$, the polynomial ring in $(n+1)(n+2)/2$ variables $x_{ij}$ where $1 \leq i , j \leq n+1$ and we set
$x_{ij} = x_{ji}$.
\begin{prop}
The electrical Lie algebra $\el_{2n}$ acts on $\R[x_{ij}]$ via derivations as follows:
\begin{align}\label{E:2i}
e_{2i} &\mapsto \partial_{ii} + \partial_{i+1,i+1} - \partial_{i,i+1}
\\
\label{E:2i1}
e_{2i-1}& \mapsto -\sum_{1 \leq p \leq q \leq n+1} x_{ip}x_{iq} \partial_{pq}
\end{align}
\end{prop}
\begin{proof}
These formulae can be checked by directly verifying the defining relations of $\el_{2n}$. Alternatively, they can be deduced by differentiating the formulae \eqref{eq:cim}.
\end{proof}
We calculate that
\begin{equation}\label{E:2ii}
[e_{2i},e_{2i-1}] \mapsto -x_{ii}\partial_{ii} + x_{i,i+1} \partial_{i+1,i+1}+\sum_{1 \leq p \leq n+1} \left( x_{i+1,p} \partial_{i+1,p}-x_{ip}\partial_{ip}\right).
\end{equation}
\begin{thm}\label{thm:stab}
The stabilizer subalgebra $\el^0_{2n}$, at the zero matrix $L_0$, of the infinitesimal action of $\el_{2n}$ on the space of response matrices is generated by $e_i$ for $i$ odd, and $[e_{2i-1},e_{2i}]$ for $i=1,2,\ldots,n$.
\end{thm}
\begin{proof}
The fact the stated elements lie in $\el^0_{2n}$ follows from \eqref{E:2i1} and \eqref{E:2ii}, since $x_{ij} = 0$ at the zero matrix. By Lemma \ref{lem:dimA}, the elements
$e_\alpha$ in the proof of Lemma \ref{lem:dim} form a basis of $\el_{2n}$. Write $\alpha_{ij} := \alpha_i + \alpha_{i+1}+ \cdots + \alpha_j$ for $ 1 \leq i \leq j \leq 2n$
to denote the positive roots of $\ssl_{2n+1}$. Then we know that $e_{\alpha_{i,i}} \in \el^0_{2n}$ for $i$ odd, and $e_{\alpha_{i,i+1}} \in \el^0_{2n}$ for each $i$. It follows easily that $e_{\alpha_{i,j}} \in \el^0_{2n}$ for every pair $1 \leq i \leq j \leq 2n$ where at least one of $i$ and $j$ are odd. It follows that
$$
\dim_\R(\el_{2n}) - \dim_\R(\el^0_{2n}) \leq \#\{(i,j) \mid 1 \leq i \leq j \leq 2n \; \text{and $i$ and $j$ are even}\} =
n(n+1)/2.
$$
By Theorem \ref{T:efficient} the action of $(EL_{2n})_{\geq 0}$ on the zero matrix $L_0$ gives a space of dimension $n(n+1)/2$. It follows that the above inequality is an equality, and that $\el^0_{2n}$ is generated by the stated elements.
\end{proof}
Note that a basis for $\el_{2n}/\el^0_{2n}$ is given by the $e_{\alpha_{i,j}}$'s where $i$ and $j$ are both even. These $\alpha_{i,j}$'s are exactly the inversions of efficient permutations (cf. proof of Theorem \ref{T:efficient}).
\section{Electrical Lie algebras of finite type}
\subsection{Dimension}
Let $D$ be a Dynkin diagram of finite type and let $A = (a_{ij})$ be the associated Cartan matrix.
To each node $i$ of $D$ associate a generator $e_i$, and define $\mathfrak e_D$ to be the Lie algebra generated over $\R$ by the $e_i$ modulo for each $i \neq j$ the relations
$ad(e_i)^{1-a_{ij}}(e_j) = 0$ for $a_{ij} \neq -1$ and $ad(e_i)^2(e_j) = -2e_i$ for $a_{ij} = -1$.
These ``electrical Serre relations'' can be deduced from the type $A$ electrical Serre relations of Section \ref{sec:el} by {\it folding}. Namely, the relation for an edge of multiplicity two ($a_{ij} = -2$) can be obtained by finding the relation for the elements $e_2$ and $e_1 + e_3$ inside $\el_3$. Similarly, the $\e_{G_2}$ relation ($a_{ij} = - 3$) can be obtained by finding the relation for the elements $e_1+e_2+e_3$ and $e_4$ inside $\e_{D_4}$, where $e_4$ corresponds to the node of valency three in $D_4$.
\begin{lemma} \label{lem:dim}
The dimension of the Lie algebra $\mathfrak e_D$ does not exceed the number of positive roots in the finite root system associated to $D$.
\end{lemma}
\begin{proof}
Consider the nilpotent subalgebra $\mathfrak u_D$ of the usual Lie algebra associated to $D$. It is well known that it is generated by Chevalley generators $\tilde e_i$ subject to the
Serre relations: $ad(\tilde e_i)^{1-a_{ij}}(\tilde e_j) = 0$ for each $i \neq j$.
The Lie algebra $\mathfrak u_D$ has a basis $\tilde e_{\alpha}$ labeled by positive roots $\alpha \in R^+$ of the root system, where the $\tilde e_i$ correspond to simple roots. Let us fix an
expression $$\tilde e_{\alpha} = [\tilde e_{i_1},[\tilde e_{i_2}, [\ldots, [\tilde e_{i_{l-1}}, \tilde e_{i_l}]\ldots]]]$$ of shortest possible length for each $\tilde e_{\alpha}$,
and define the corresponding
\begin{equation}\label{E:alpha}
e_{\alpha} = [e_{i_1},[e_{i_2}, [\ldots, [e_{i_{l-1}}, e_{i_l}]\ldots]]]
\end{equation}
in $\mathfrak e_D$. We claim that the $e_{\alpha}$ span $\mathfrak e_D$. Indeed, it is enough to show that any expression $ [e_{j_1},[e_{j_2}, [\ldots, [e_{j_{\ell-1}}, e_{j_\ell}]\ldots]]]$
lies in the linear span of the $e_{\alpha}$-s.
Assume otherwise, and take a counterexample of smallest possible total length $\ell$.
Let us define $\hat e_\alpha$ in the free Lie algebra $\mathfrak f_D$ with generators $\hat e_i$ using \eqref{E:alpha}.
Then in $\mathfrak f_D$ we have a relation of the form
$$
[\hat e_{j_1},[\hat e_{j_2}, [\ldots, [\hat e_{j_{\ell-1}}, \hat e_{j_\ell}]\ldots]]]
- \sum_\alpha c_{\alpha} \hat e_{\alpha} = \hat x \in I
$$
where $I$ denotes the ideal generated by the Serre relations, so that $\mathfrak u_D = \mathfrak f_D/I$. Now $\mathfrak f_D$ is naturally $\Z$-graded, and $I$ is a graded ideal, so we may assume all terms in the relation are homogeneous with the same degree. Now replacing each instance of the Serre relation in $x$ with the corresponding electrical Serre relation gives us a relation
$$
[e_{j_1},[e_{j_2}, [\ldots, [e_{j_{\ell-1}}, e_{j_\ell}]\ldots]]] - \sum_\alpha c_{\alpha} e_{\alpha} = x
$$
in $\mathfrak e_D$, where $x$ is a sum of terms of the form $ [e_{k_1},[e_{k_2}, [\ldots, [e_{k_{\ell-1}}, e_{k_{\ell'}}]\ldots]]]$ where $\ell' < \ell$, and thus by assumption lies in the span of the $e_\alpha$-s. The statement of the lemma follows.
\end{proof}
\begin{conjecture} \label{con:1}
The dimension of $\mathfrak e_D$ coincides with the number of positive roots in the root system of $D$.
\end{conjecture}
One can also define the Lie algebras $\e_D$ for any Dynkin diagram $D$, finite type or not.
\begin{conjecture} \label{con:2}
The Lie algebra $\e_D$ is finite dimensional if and only if $D$ is of finite type.
\end{conjecture}
\subsection{Examples}
We illustrate Conjecture \ref{con:1} with some examples. For electrical type $A_{2n}$ it has already been verified in Theorem \ref{thm:elsp}.
\subsubsection{Electrical $B_2$}
Consider the case when $D$ has two nodes connected by a double edge. In that case we denote by $\mathfrak e_D = \mathfrak {eb}_2$ the Lie algebra generated by two generators $e$ and $f$ subject to the relations $$[e,[e,[e,f]]]=0, \;\;\; [f,[f,e]] = -2f.$$
\begin{lemma}\label{L:GL2}
The Lie algebras $\mathfrak {eb}_2$ and $\mathfrak {gl}_2$ are isomorphic.
\end{lemma}
\begin{proof}
Consider the map $$e \mapsto \left(\begin{matrix} 1 & 1\\ 0 & 1\\ \end{matrix} \right), \;\;\; f \mapsto \left(\begin{matrix} 0 & 0\\ 1 & 0\\ \end{matrix} \right).$$ One easily checks
that it is a Lie algebra homomorphism, and that it is surjective. By Lemma \ref{lem:dim} we know the dimension of $\mathfrak {eb}_2$ is at most four, so the map must be an isomorphism.
\end{proof}
\subsubsection{Electrical $G_2$}
Consider the Lie algebra $\mathfrak {eg}_2$ generated by two generators $e$ and $f$ subject to the relations
$$[e,[e,[e,[e,f]]]]=0, \;\;\; [f,[f,e]] = -2f.$$
\begin{lemma}
The Lie algebra $\mathfrak {eg}_2$ is six-dimensional.
\end{lemma}
\begin{proof}
According to Lemma \ref{lem:dim} the elements $e, f, [ef], [e[ef]], [e[e[ef]]], [f[e[e[ef]]]]$ form a spanning set for $\mathfrak {eg}_2$. Thus it remains to check that they are linearly independent.
This is easily done inside the following faithful representation of $\mathfrak {eg}_2$ in $\mathfrak {gl}_4$:
$$e \mapsto \left(\begin{matrix} 1 & 1 & 0 & 1\\ 0 & 1 &1 & 0 \\ 0 & 0 &1 & 0 \\0 & 0 &0 & 1 \\ \end{matrix} \right), \;\;\;
f \mapsto \left(\begin{matrix} 0 & 0 & 0 & 0\\ 0 & 0 &0 & 0 \\ 0 & 0 &0 & 0 \\1 & 0 &0 & 0\\ \end{matrix} \right).$$
\end{proof}
\subsubsection{Electrical $C_3$}
Consider the Lie algebra $\mathfrak {ec}_3$ generated by three generators $e$, $f$ and $g$ subject to the relations
$$[e,[e,[e,f]]]=0, \;\;\; [f,[f,e]] = -2f, \;\;\; [f,[f,g]] = -2f, \;\;\; [g,[g,f]] = -2g, \;\;\; [e,g]=0.$$
\begin{lemma}
The Lie algebra $\mathfrak {ec}_3$ is nine-dimensional.
\end{lemma}
\begin{proof}
We consider two representations of $\mathfrak {ec}_3$: one inside $\mathfrak {sl}_9$ and one inside $\mathfrak {sl}_2$. Let $E_{ij}$ denote a matrix with a $1$ in the $(i,j)$-th position and $0$'s elsewhere. For the first representation, we define:
\begin{align*}
e &\mapsto E_{42} + E_{54}+E_{76}+E_{87}+E_{89}\\
f& \mapsto 2E_{24} -2 E_{26}+2E_{41}+2E_{45}-E_{47}-2E_{63}+E_{67}+E_{98} \\
g &\mapsto -E_{36} - E_{62} - E_{74} - E_{85} + E_{89}
\end{align*}
For the second representation we define
$$e \mapsto \left(\begin{matrix} 1 & 1\\ 0 & 1\\ \end{matrix} \right), \;\;\; f \mapsto \left(\begin{matrix} 0 & 0\\ 1 & 0\\ \end{matrix} \right),
\;\;\; g \mapsto \left(\begin{matrix} 0 & 1\\ 0 & 0\\ \end{matrix} \right).$$
One verifies directly that these are indeed representations of $\mathfrak {ec}_3$, and the direct sum of these two representations is faithful, from which the dimension is easily
calculated. In fact, the first representation is simply the adjoint representation of $\ec_3$, in the basis
$$e, f, g, [ef], [e[ef]], [fg], [e[fg]], [e[e[fg]]], [f[e[e[fg]]]],$$ and $\ec_3$ has a one-dimensional center which acts non-trivially in the second representation.
\end{proof}
\subsection{$Y-\Delta$ transformation of type $B$}
Let $e$ and $f$ be the generators of $\mathfrak {eb}_2$ as before, and denote by $u(t) = \exp(t e)$ and $v(t) = \exp(t f)$ the corresponding one-parameter subgroups. The following
proposition is a type $B$ analog of the star-triangle transformation (Proposition \ref{P:YDelta}).
\begin{prop}
We have $$u(t_1) v(t_2) u(t_3) v(t_4) = v(p_1) u(p_2) v(p_3) u(p_4),$$ where $$p_1 = \frac{t_2 t_3^2 t_4}{\pi_2}, \;\; p_2 = \frac{\pi_2}{\pi_1}, \;\; p_3 = \frac{\pi_1^2}{\pi_2}, \;\;
p_4 = \frac{t_1 t_2 t_3}{\pi_1},$$ where $$\pi_1 = t_1 t_2 + (t_1+ t_3) t_4 + t_1 t_2 t_3 t_4, \;\;\; \pi_2 = t_1^2t_2+(t_1+t_3)^2t_4+t_1t_2t_3t_4(t_1+t_3).$$
\end{prop}
\begin{proof}
Direct calculation inside $GL_2$, using Lemma \ref{L:GL2}. We have $$u(t) = \left(\begin{matrix} e^t & e^t t\\ 0 & e^t\\ \end{matrix} \right), \;\;\; v(t)= \left(\begin{matrix} 1 & 0\\ t & 1\\ \end{matrix} \right),$$
and both sides of the equality are equal to $$\left(\begin{matrix} e^{t_1 + t_3} (1 + t_3 t_4 + t_1 (t_2 + t_4 + t_2 t_3 t_4)) & e^{t_1 + t_3} (t_1 + t_3 + t_1 t_2 t_3)\\
e^{t_1 + t_3} (t_2 + t_4 + t_2 t_3 t_4) & e^{t_1 + t_3} (1 + t_2 t_3)\\ \end{matrix} \right).$$
\end{proof}
\begin{remark}
One can consider a one-parameter family of deformations of the above formulas by taking
$$\pi_1 = t_1 t_2 + (t_1+ t_3) t_4 + \tau t_1 t_2 t_3 t_4, \;\;\; \pi_2 = t_1^2t_2+(t_1+t_3)^2t_4+\tau t_1t_2t_3t_4(t_1+t_3).$$
At $\tau=0$ this specializes
to the transformation found by Berenstein and Zelevinsky in \cite[Theorem 3.1]{BZ}. Furthermore, just like the transformation in \cite{BZ}, this deformation is given by {\it {positive}}
rational formulae.
\end{remark}
|
\section{Introduction}
Even though it has been around for quite a while \cite{Vafa:1996xn}, F-theory has recently received a lot of new attention as a setup where Grand Unified Theories (GUTs) can be conceived from string theory. Starting with \cite{Donagi:2008ca,Beasley:2008dc,Beasley:2008kw} the phenomenology of F-theory GUTs has become an active field of research. The basic idea is that the GUT theory is localized on a $(p,q)$ seven-brane $S$ inside a three-dimensional base manifold $B$ of an F-theory compactification on an elliptically fibered Calabi-Yau fourfold. The location of the GUT brane and the gauge group are determined by the degeneration of the elliptic fibration. Chiral matter localizes on curves inside the GUT brane $S$ where gauge enhancement occurs, Yukawa couplings sit at points. For many phenomenological applications it is sufficient to consider the field theory living on the GUT brane without specifying the details of the global F-theory compactification. However, fluxes, monodromies or consistency constraints such as tadpole cancellation cannot be addressed in a purely local setup. These issues have recently received a lot of attention in the literature \cite{Andreas:2009uf,Blumenhagen:2009yv,Marsano:2009wr,Grimm:2009yu,Blumenhagen:2010ja,Cvetic:2010rq,Hayashi:2010zp,Chen:2010tp,Chen:2010ts,Grimm:2010ez,Marsano:2010ix,Chung:2010bn,Heckman:2010qv,Cvetic:2010ky,Cecotti:2010bp,Marsano:2010sq,Collinucci:2010gz,Chiou:2011js,Ludeling:2011en,Knapp:2011wk,Dolan:2011iu}. Therefore it is interesting to see whether it is possible to embed the local F-theory GUT into a compactification on a Calabi-Yau fourfold. Most known examples of compact Calabi-Yau manifolds are hypersurfaces or complete intersections in a toric ambient space. It is thus natural to look for Calabi-Yau fourfolds within this class of examples. A prescription for constructing elliptically fibered Calabi-Yau fourfolds as complete intersections in a six-dimensional toric ambient space has been given in \cite{Blumenhagen:2009yv,Grimm:2009yu}. Before that complete intersection Calabi-Yau fourfolds in F-theory had already been used in the context of F-theory uplifts of type IIB string theory \cite{Collinucci:2008zs,Collinucci:2009uh,Blumenhagen:2009up}. A similar construction has also been discussed in \cite{Cvetic:2010rq}. It has been shown in examples that it is indeed possible to construct viable F-theory GUTs within this framework.
The construction of \cite{Blumenhagen:2009yv} is very well-suited for a systematic search of a large class of models. This is interesting for several reasons: one goal is to find particularly nice examples of F-theory compactifications. Even though the known examples have been able to incorporate F-theory models, one usually gets much more than just that. In minimal F-theory GUTs one typically needs only very few Yukawa points and a small number of moduli on the matter curves. This is not satisfied in most known global models. A related question deals with the genericity of F-theory GUTs. The geometric configurations used for constructing such models are usually quite special and one may wonder how often they can be realized in elliptically fibered fourfolds. From the point of view of model building it is useful to have some easy-to-check geometric conditions which makes it possible to select suitable models from a large class of geometries. This will be discussed in more details in the text. From a mathematical point of view it might be interesting to obtain a partial classification of Calabi-Yau fourfolds.
This review article discusses selected topics in toric geometry and F-theory GUTs. The paper is organized as follows: in section \ref{sec-fth} we recall the construction of global F-theory models and discuss the basic requirements we would like to impose. In section \ref{sec-toric} we review several notions in toric geometry which are required in order to perform the F-theory calculations. The geometries one has to deal with are usually quite complicated, and very often one has to rely on computer support in order to be able to do explicit calculations. Therefore we discuss how such calculations can be implemented using existing software such as PALP \cite{Kreuzer:2002uu}. We will mainly focus on the application of toric geometry in the context of F-theory model building. For a more complete picture on this vast subject of F-theory phenomenology we refer to other review articles such as \cite{Denef:2008wq,Heckman:2010bq,Weigand:2010wm}. For more extensive discussions of toric geometry we recommend \cite{fultontoric,Kreuzer:2006ax,coxbook}.
\section{Global F-theory models}
\label{sec-fth}
\subsection{Setup}
In this section we introduce the basic concepts and notions used in global F-theory models. In the remainder of this review we will explain the techniques which are necessary to do calculations within this framework. For more details on how the quantities introduced below come about we refer to the original papers or the recent review article \cite{Weigand:2010wm}.
In \cite{Blumenhagen:2009yv} it has been proposed to construct Calabi-Yau fourfolds, which are suitable for F-Theory model building, as complete intersections of two hypersurfaces in a six-dimensional toric ambient space. The hypersurface equations have the following structure:
\begin{equation}
\label{cicydef}
P_{B}(y_i,w)=0\qquad\qquad P_W(x,y,z,y_i,w)=0
\end{equation}
The first equation only depends on the homogeneous coordinates $(y_i,w)$ of the three-dimensional base $B$ of the elliptically fibered Calabi-Yau fourfold $X_4$. Here we have singled out one coordinate $w$, indicating that the divisor given by $w=0$ defines a seven-brane $S$ which supports a GUT theory of the type introduced in \cite{Donagi:2008ca,Beasley:2008dc,Beasley:2008kw}. The second equation in (\ref{cicydef}) defines a Weierstrass model, where $(x,y,z)$ are those coordinates of the six-dimensional ambient space that describe the torus fiber. For this type of elliptic fibrations $P_W$ is of Tate form which is defined as follows:
\begin{equation}
\label{tate}
P_W=x^3-y^2+xyz a_1+x^2z^2a_2+yz^3a_3+xz^4a_4+z^6a_6
\end{equation}
The $a_n(y_i,w)$ are sections of $K_B^{-n}$, where $K_B$ is the canonical bundle of the base manifold. Furthermore $x$ and $y$ can be viewed as sections of $K_B^{-2}$ and $K_B^{-3}$, respectively.
The information about the F-theory model is encoded in the Tate equation (\ref{tate}). In order to have a non-trivial gauge group on the GUT brane the elliptic fibration must degenerate above $S$. The gauge group is determined by the structure of the singularity. The elliptic fibration becomes singular at the zero locus of the discriminant $\Delta$. Defining the polynomials $\beta_2=a_1^2+4a_2$, $\beta_4=a_1a_3+2a_4$ and $\beta_6=a_3^2+4a_6$ the discriminant is given by the following expression:
\begin{equation}
\Delta=-\frac{1}{4}\beta_2^2(\beta_2\beta_6-\beta_4^2)-8\beta_4^3-27\beta_6^2+9\beta_2\beta_4\beta_6
\end{equation}
According to Kodaira's classification \cite{Kodaira} and Tate's algorithm \cite{tate}, the gauge group can be inferred from the factorization properties of the $a_n(y_i,w)$ with respect to $w$. Considering for instance $SU(5)$- and $SO(10)$-models, the factorization looks like this:
\begin{eqnarray}
\label{su5so10tate}
SU(5):&&a_1=b_5w^0\quad a_2=b_4w^1\quad a_3=b_3w^2\quad a_4=b_2w^3\quad a_6=b_0w^5\nonumber\\
SO(10):&& a_1=b_5w^1\quad a_2=b_4w^1\quad a_3=b_3w^2\quad a_4=b_2w^3\quad a_6=b_0w^5
\end{eqnarray}
The $b_i$s are sections of some appropriate line bundle over $B$ that have at least one term independent of $w$.
In F-theory GUT models chiral matter localizes on curves on $S$, where a rank $1$ enhancement of the gauge group appears. In $SU(5)$ models the matter curves are at the following loci inside $S$:
\begin{equation}
b_3^2b_4-b_2b_3b_5+b_0b_5^2=0\quad \textrm{{\bf 5} matter} \qquad b_5=0\quad \textrm{{\bf 10} matter}\,,
\end{equation}
The matter curves for the $SO(10)$ models are at:
\begin{equation}
b_3=0\quad \textrm{{\bf 10} matter} \qquad b_4=0\quad \textrm{{\bf 16} matter}\,.
\end{equation}
Yukawa coupling arise at points inside $S$ where the GUT singularity has an rank $2$ enhancement. In $SU(5)$ models the Yukawa points sit at:
\begin{eqnarray}
b_4=0 \cap b_5=0&\quad& {\bf 10\:10\:5} \textrm{ Yukawas}\quad \textrm{$E_6$ enhancement} \nonumber\\
b_5=0 \cap b_3=0&\quad& {\bf 10\:\bar{5}\:\bar{5}} \textrm{ Yukawas} \quad \textrm{$SO(12)$ enhancement}\,.
\end{eqnarray}
In the $SO(10)$-case we have the following Yukawa couplings:
\begin{eqnarray}
b_3=0 \cap b_4=0&\quad& {\bf 16\:16\:10} \textrm{ Yukawas}\quad \textrm{$E_7$ enhancement} \nonumber\\
b_2^2-4 b_0 b_4=0 \cap b_3=0&\quad& {\bf 16\:10\:10} \textrm{ Yukawas} \quad \textrm{$SO(14)$ enhancement}\,.
\end{eqnarray}
Given a complete intersection Calabi-Yau fourfold of the form (\ref{cicydef}) the expressions for matter curves and Yukawa points are globally defined and can be calculated explicitly. Having a global F-theory compactification we can furthermore calculate the Hodge numbers and the Euler number $\chi_4$ of the Calabi-Yau fourfold $X_4$. The latter enters the D3-tadpole cancellation condition,
\begin{equation}
\frac{\chi_4}{24}=N_{D3}+\frac{1}{2}\int_{X_4}G_4\wedge G_4,
\end{equation}
where $G_4$ denotes the fourform flux on $X_4$ and $N_{D3}$ is the number of $D3$-branes.
\subsection{Geometric Data in F-theory models}
\label{sec-fgeom}
So far we have summarized the basic structure of a global F-theory GUT. In the present section we will discuss which properties of the GUT model are encoded in the geometries of the base manifold $B$ and the Calabi-Yau fourfold $X_4$. We will not go deeply into the phenomenology of F-theory GUTs but rather focus on the basic geometric properties which should be satisfied in order to obtain a viable GUT model.
\subsubsection*{Base Manifold}
Since the GUT brane $S$ is a divisor in a three-dimensional base manifold $B$, a large amount of information about the model can be extracted from the geometry of $B$. The base $B$ is a non-negatively curved manifold of complex dimension three. In our setup it will be given by a hypersurface in a toric ambient space. Note that Fano threefolds are not a good choice for $B$ due to the lack of a decoupling limit \cite{Cordova:2009fg}. In section \ref{sec-toric} we discuss a systematic construction of such base manifolds using toric geometry. In order to have a well defined model we have to make sure that $B$ is non-singular. In contrast to Calabi-Yau threefolds the base manifolds for F-theory GUTs may inherit the singularities of the ambient space. Therefore checks for the regularity of $B$ have to be implemented.
Having found a suitable base manifold the next step is to identify divisors inside $B$ that can support GUT models. The most promising candidates for F-theory model building are del Pezzo surfaces. These are Fano twofolds (see for instance \cite{Griffiths:1994ab}). Note however, that del Pezzos are not the only possibility for the construction of GUT models in F-theory. See \cite{Braun:2010hr} for a recent discussion. There are several motivations to focus on del Pezzo divisors. In local F-theory GUTs the del Pezzo property ensures the existence of a decoupling limit \cite{Beasley:2008dc,Beasley:2008kw}. For $SU(5)$ GUT models, the fact that del Pezzos have $h^{0,1}=h^{2,0}=0$ implies some powerful vanishing theorems which forbid exotic matter after breaking $SU(5)$ to the Standard Model gauge group \cite{Beasley:2008kw}.
We can identify candidates for del Pezzo divisors inside $B$ by their topological data. Suppose the base manifold $B$ is embedded in a toric ambient space which has toric divisors $D_i$. The $D_i$ give a homology basis of the ambient space. In this setup the hypersurface is specified by a divisor, which we will by abuse of notation also call $B$, that is given in terms of a linear combination of the $D_i$. The total Chern class of a particular divisor $S$ in the ambient space is, after restriction to $B$ (for more details see section \ref{sec-intmori}):
\begin{equation}
c(S)=\frac{\prod_i(1+D_i)}{(1+B)(1+S)}
\end{equation}
In order to apply this formula we have to know the intersection ring of $B$. As we will discuss in section \ref{sec-intmori} this can be obtained from the intersection ring of the ambient space.
A necessary condition for a divisor $S$ to be $dP_n$ is that it must have the following topological data:
\begin{equation}
\int_Sc_1(S)^2=9-n\qquad \int_Sc_2(S)=n+3\qquad\Rightarrow\qquad \chi_h=\int_S\mathrm{Td}(S)=1,
\end{equation}
where $\chi_h=\sum_i(-1)^ih^{0,i}$ is the holomorphic Euler characteristic and $\mathrm{Td}$ denotes the Todd class. In the equations above the integration over the four-cycle (representing the divisor) $S$ is equivalent to computing the intersection with $S$.
Since del Pezzos are Fano twofolds, we have a further necessary condition: the intersections of $c_1(S)$ with curves on $S$ have to be positive. In the toric setup we can only check this for curves which are induced from the divisors on the ambient space. In that case the condition is:
\begin{equation}
\label{poscurve}
D_i\cdot S\cdot c_1(S) > 0\qquad D_i\neq S \qquad \forall D_i\cdot S\neq\emptyset\,.
\end{equation}
In order to make these calculations we need to know the homology basis of toric divisors and their intersection numbers.
In local F-theory GUTs the del Pezzo property is sufficient to ensure the existence of a decoupling limit. For global models further checks are in order. Gravity decouples from the gauge degrees of freedom if the mass ratio $M_{GUT}/M_{pl}$ becomes parametrically small. The Planck mass $M_{pl}$ and the mass scale $M_{GUT}$ of the GUT theory are related to the geometry of $B$ and $S$ in the following way:
\begin{equation}
M_{pl}^2\sim \frac{M_s^8}{g_s^2}\mathrm{Vol}(B)\qquad M_{GUT}\sim \mathrm{Vol}(S)^{-\frac{1}{4}}\qquad 1/g^2_\textmd{YM}\sim \frac{M_s^4}{g_s}\mathrm{Vol}(S)\,,
\end{equation}
Therefore one has:
\begin{equation}
\frac{M_{GUT}}{M_{pl}}\sim g^2_\textmd{YM}\frac{\mathrm{Vol}(S)^{3/4}}{\mathrm{Vol}(B)^{1/2}}
\end{equation}
There are two ways to achieve a small value for $M_{GUT}/M_{pl}$, now commonly referred to as the physical and the mathematical decoupling limit. In the physical decoupling limit the volume of the $GUT$ brane $S$ is kept finite while $\mathrm{Vol}(B)\rightarrow\infty$. The mathematical decoupling limit takes $\mathrm{Vol}(S)\rightarrow 0$ for finite volume of $B$. The two limits may not be equivalent in the sense that they may be implemented by tuning different K\"ahler parameters. The volumes of $B$ and $S$ can be determined in terms of the K\"ahler form $J$ of the ambient toric variety restricted to $B$:
\begin{equation}
\label{volumes}
\mathrm{Vol}(B)=J^3\qquad \mathrm{Vol}(S)=S\cdot J^2
\end{equation}
In order for the volumes to be positive we must find a basis of the K\"ahler cone $K_i$ where, by definition, $J$ can be written as $J=\sum_ir_iK_i$ with $r_i>0$. The existence of mathematical and physical decoupling limits can be deduced from the moduli dependence of these volumes.
Having found a suitable base manifold we can also study matter curves and Yukawa couplings. The curve classes $M$ of the matter curves can be expressed in terms of the toric divisors of the ambient space. The genus of the matter curves can be computed using the first Chern class of the matter curve and the triple intersection numbers:
\begin{equation}
c(M)=\frac{\prod_i(1+D_i)}{(1+B)(1+S)(1+M)}
\end{equation}
Here we have assumed that $M$ is irreducible.
After expanding this expression to get $c_1(M)$, the Euler number is obtained by the following intersection product:
\begin{equation}
\chi(M)=2-2g(M)=c_1(M)\cdot M\cdot S
\end{equation}
The genus of a matter curve gives us information about the number of moduli the curve has. Since these moduli have to be stabilized, matter curves of low genus are desirable from a phenomenological point of view. In the generic situation the equations specifying the Yukawa points can be expressed as classes $Y_1,Y_2$ in terms of the toric divisors. The number of Yukawa points is then given by the following intersection product:
\begin{equation}
n_{\textrm{Yukawa}}=S\cdot Y_1\cdot Y_2
\end{equation}
In order to account for the Standard Model Yukawa couplings only a small number of Yukawa points is needed.
\subsubsection*{Fourfold}
Given a base manifold $B$ one can construct a Calabi-Yau fourfold $X_4$ which is an elliptic fibration over $B$. As described in the next section this can be done systematically using toric geometry. However, not all of the desirable features of global F-theory models are automatic in this construction. The main requirement on $X_4$ is that it is a complete intersection of two hypersurfaces. Furthermore these hypersurfaces must have a specific structure (\ref{cicydef}).
In order to be able to use powerful mathematical tools we furthermore have to require that there exists a nef-partition (cf. section \ref{sec-hypercicy}) which is compatible with the elliptic fibration. When this elementary requirement is satisfied we can engineer a GUT model. This is done in two steps: first we have to identify the GUT divisor $S$, given by the equation $w=0$ in $B$ within the Calabi-Yau fourfold. The second step is to impose the GUT group. This amounts to explicitly imposing the factorization conditions such as (\ref{su5so10tate}) on the Tate model. This means that we have to remove all those monomials in (\ref{tate}) which do not satisfy the factorization constraints. This amounts to fixing a number of complex structure moduli of $X_4$.
Recently there has been active discussion in the literature how to globally define fluxes in F-theory models \cite{Blumenhagen:2009yv,Marsano:2009wr,Grimm:2009yu,Hayashi:2010zp,Marsano:2010ix,Chung:2010bn,Marsano:2010sq,Collinucci:2010gz,Dolan:2011iu}. In F-theory model building fluxes enter at several crucial points. Gauge flux along the matter curve is needed in order to generate chiral matter. Breaking of the GUT group to the Standard Model gauge group can be achieved by turning on $U(1)$-flux. Furthermore, in $SU(5)$ F-theory GUTs we need global $U(1)$s in order to forbid dimension $4$ proton decay operators. In $SO(10)$ F-theory GUTs they are needed in order to obtain chiral matter \cite{Chen:2010tg,Chen:2010ts}. A general global description of the fourform flux $G_4$ is still missing. In \cite{Donagi:2009ra} an auxiliary construction involving spectral covers that factorize was used to describe fluxes locally in the vicinity of the GUT brane. It has been shown in \cite{Grimm:2010ez,Marsano:2010ix,Chung:2010bn} that under certain circumstances the information captured by the spectral cover can be encoded in the Tate model, and is therefore global. However, this need not be the case \cite{Hayashi:2010zp}. In \cite{Grimm:2010ez} it has been shown that a spectral cover which factorizes is generically globally defined for ``$U(1)$-restricted Tate models''. This is achieved by imposing a global $U(1)_X$ symmetry in the elliptic fibration. In terms of the Tate model this is achieved by setting $a_6=0$.
\section{Ingredients and Techniques from Toric Geometry}
\label{sec-toric}
In the previous section we have introduced quantities which encode important information about F-theory GUT models in the geometry of the base manifold and the Calabi-Yau fourfold. In this section we will provide the tools to calculate them. The input data needed for these calculations can be obtained by using toric geometry. After giving the basic definitions we will discuss how to describe hypersurfaces and complete intersections of hypersurfaces in toric ambient spaces. Then we explain how to obtain the intersection ring and the K\"ahler cone, or dually, the Mori cone. Finally we will discuss how to use the computer program PALP \cite{Kreuzer:2002uu} for calculations in toric geometry. This discussion of toric geometry has been compiled with a view towards the applications in F-theory model building. It is by no means an exhaustive description of this vast subject which brings together algebraic geometry and combinatorics.
\subsection{Toric Varieties}
\label{sec-toricdef}
We start by defining a toric variety $X$ of dimension $n$ as the following quotient:
\begin{equation}
\label{torvar}
X=(\mathbb{C}^r-Z)/((\mathbb{C}^{\ast})^{r-n}\times G),
\end{equation}
where $G$ is a finite abelian group, $(\mathbb{C}^{\ast})^{r-n}$ describes the action of an algebraic $(r-n)$-torus and $Z\subset\mathbb{C}^r$ is an exceptional set which tells us which combinations of coordinates are not allowed to vanish simultaneously. The simplest example is $\mathbb{CP}^2$, where the $\mathbb{C}^{\ast}$-action is given by $(z_1,z_2,z_3)\sim (\lambda z_1,\lambda z_2, \lambda z_3)$, $\lambda\in\mathbb{C}^{\ast}$, the exceptional set is $Z=\{z_1=z_2=z_3=0\}$ and $G$ is trivial. Thus, as is well-known: $\mathbb{CP}^2=(\mathbb{C}^3-\{z_1=z_2=z_3=0\})/((z_1,z_2,z_3)\sim (\lambda z_1,\lambda z_2, \lambda z_3))$.
The crucial fact about toric geometry is that the geometric data of the toric variety can be described in terms of combinatorics of cones and polytopes in dual pairs of integer lattices. The information about the toric variety is encoded in a fan $\Sigma$, which is a collection of strongly convex rational polyhedral cones where all the faces and intersections of pairs of cones also belong to the fan. 'Strongly convex' means that all the cones of the fan have an apex at the origin, 'rational' means that the rays that span the cone go through points of the lattice. We denote by $\Sigma^{(n)}$ the set of all $n$--dimensional cones. In order to define the fan we use the fact that a toric variety $X$ contains an $n$-torus $T=(\mathbb{C}^{\ast})^n$ as a dense open subset whose action extends to $X$. Parametrizing $T$ by coordinates $(t_1,\ldots,t_n)$, one defines the character group $M=\{\chi:T\rightarrow \mathbb{C}^{\ast}\}$ and the one-parameter subgroups $N=\{\lambda:\mathbb{C}^{\ast}\rightarrow T\}$. $M$ and $N$ can be identified with integer lattices that are isomorphic to $\mathbb{Z}^n$. Given a point $m\in M$, the character is given by $\chi^m(t)=t_1^{m_1}\ldots t_n^{m_n}\equiv t^m$. This is a holomorphic function on $T$, and descends to a rational function on the toric variety $X$. For every $u\in N$, $\lambda$ is defined as $\lambda^u(\tau)=(\tau^{u_1},\ldots,\tau^{u_n})$ for $\tau\in\mathbb{C}^{\ast}$. The fan $\Sigma$ and its cones $\sigma$ are defined on the real extension $N_{\mathbb{R}}$ of $N$. The lattices $M,N$ are dual due to the composition $(\chi\circ\lambda)(\tau)=\chi(\lambda(\tau))=\tau^{\langle\chi,\lambda\rangle}$, where $\langle\chi^m,\lambda^u\rangle=m\cdot u$ is the scalar product.
The $M$-lattice encodes the data about regular monomials in $X$, the $N$-lattice captures the information about the divisors. The divisors defined by $\chi^m=0$ can be decomposed in terms of irreducible divisors $D_j$: $\mathrm{div}(\chi^m)=\sum_{j=1}^ra_jD_j$. These divisors are principal divisors, i.e. divisors of meromorphic function where $D_j$ correspond to poles or zeros and the $a_j$ are orders of the pole/zero. The coefficients $a_j(m)\in\mathbb{Z}$ are unique, and there exists a map $m\rightarrow a_j(m)=\langle m,v_j\rangle$ with $v_j\in N$. Thus there is a vector $v_j$ for every irreducible divisor $D_j$. The $v_j$ are the primitive generators of the one-dimensional cones $\rho_j$ (i.e. rays) in the fan $\Sigma$. The convex hull of the $v_j$ defines a polytope $\Delta^{\ast}=\mathrm{conv}\{v_j\}$. Locally, we can write the divisors as $D_j=\{z_j=0\}$, where $z_j$ is regarded as a local section of a line bundle. $D_j$ are called toric divisors. There are linear relations among the $v_j\in\Delta^{\ast}$ which translate into linear relations among the toric divisors.
In order to make contact with the definition (\ref{torvar}) of $X$, we view the $\{z_j\}$ as global homogeneous coordinates $(z_1:\ldots :z_r)$. If all $z_j$ are non-zero the coordinates $(\lambda^{q_1}z_1:\ldots:\lambda^{q_r}z_r)\sim(z_1:\ldots:z_r)$ with $\lambda\in\mathbb{C}^{\ast}$ describe the same point on the torus $T$, if $\sum q_jv_j=0$ for $v_j\in N$ as above. Since the $v_j$ live in an $n$-dimensional lattice they satisfy $r-n$ linear relations. If the $v_j$ do not span the $N$-lattice there is a finite abelian group $G$ such that $G\simeq N/(\mathrm{span}\{v_1,\ldots,v_r\})$. Identifications coming from the action of $G$ have to be added to the identifications between the homogeneous coordinates coming from the torus action. Having introduced the fan $\Sigma$, we are also able to specify the exceptional set $Z$ that tells us where the homogeneous coordinates are not allowed to vanish simultaneously: a subset of coordinates $z_j$ is allowed to vanish simultaneously if and only if there is a cone $\sigma\in\Sigma$ containing all the corresponding rays $\rho_j$. To be more precise, the exceptional set is the union of sets $Z_I$ with minimal index sets $I$ of rays for which there is no cone that contains them: $Z=\cup_I Z_I$. This is equivalent to the statement that the corresponding divisors $D_j$ intersect in $X$. Putting the pieces together we arrive at the definition (\ref{torvar}).
There are two important properties of the fan $\Sigma$ which translate into crucial properties of the toric variety $X$. Firstly, $X$ is compact if and only if the fan is complete, i.e. if the support of the fan covers the $N$-lattice: $|\Sigma|=\bigcup_{\Sigma}\sigma=N_{\mathbb{R}}$. Secondly, $X$ is non-singular if and only if all cones are simplicial and basic, which means that all cones $\sigma\in\Sigma$ are generated by a subset of a lattice basis of $N$. Singularities can be removed by blow-ups, where singular points are replaced by $\mathbb{P}^{n-1}$s. All the singularities of a toric variety can be resolved by a series of blowups. These correspond to subdivisions of the fan. In order to completely resolve all singularities one must find a maximal triangulation of the fan. In many cases it is sufficient to find a maximal triangulation of the polytope $\Delta^{\ast}$.
Finally, let us emphasize the significance of the homogeneous weights $q_i$. In general there will be a full $(r-n)\times r$ matrix $Q_{ij}$, called weight matrix, whose $(r-n)$ lines encode the $\mathbb{C}^{\ast}$-actions. Since each of the ${z_j}$ corresponds to an irreducible divisor in $X$, the columns of the weight matrix define a homology basis of the divisors $D_j$. In physics language the weights $q_i$ are the $U(1)$-charges in the gauged linear sigma model that defines the toric variety $X$. Note that the weights contain all the information to recover the $M$- and $N$-lattice. With the weight matrix as input this can be done using PALP.
\subsection{Hypersurfaces and complete intersections}
\label{sec-hypercicy}
Having defined a toric variety we go on to discuss hypersurfaces and complete intersections of hypersurfaces in toric varieties. The hypersurface equations are sections of non-trivial line bundles. The information of these bundles can be recovered from their transition functions. In this context we introduce the notions of Cartier divisors and Weil divisors. A Cartier divisor is given, by definition, by rational equations $f_{\alpha}=0$ and regular transition functions $f_{\alpha}/f_{\beta}$ on the overlap of two coordinate patches $U_{\alpha},U_{\beta}$. Cartier divisor classes determine the Picard group $\mathrm{Pic}(X)$ of holomorphic line bundles. Weil divisors are finite formal sums of irreducible varieties of codimension $1$. On a toric variety the Chow group $A_{n-1}(X)$ modulo linear equivalence is generated by the $T$-invariant irreducible divisors $D_j$ modulo the principal divisors $\mathrm{div}(\chi^m)$, $m\in M$. A Weil divisor of the form $D=\sum a_jD_j$ is Cartier if there exists an $m_{\sigma}\in M$ for each maximal cone $\sigma\in\Sigma^{(n)}$ such that $\langle m_{\sigma},v_j\rangle=-a_j$ for all rays $\rho_j\in\sigma$. If $X$ is smooth then all Weil divisors are Cartier. If $X$ is compact and $D$ is Cartier then $\mathcal{O}(D)$ is generated by global sections if and only if $\langle m_{\sigma},v_j\rangle>-a_j$ for $\sigma\in\Sigma^{(n)}$ and $\rho_j\not\subset\sigma$. If this is the case for $v\in\sigma$, $\psi_D(v)=\langle m_{\sigma},v\rangle$ is a strongly convex support function. With that we can define a polytope $\Delta_D=\{m\in M_{\mathbb{R}}: \langle m_{\sigma},v_j\rangle\geq -a_j\}$. This is a convex lattice polytope in $M_{\mathbb{R}}$ whose lattice points provide global sections of the line bundle $\mathcal{O}(D)$ corresponding to the divisor $D$. $D$ is generated by global sections if and only if $\Delta_D$ is the convex hull of $\{m_{\sigma}\}$. Furthermore, $D$ is ample if and only if $\Delta_D$ is $n$-dimensional with vertices $m_{\sigma}$ for $\sigma\in\Sigma^{(n)}$ and with $m_{\sigma}\neq m_{\tau}$ for $\sigma\neq\tau\in\Sigma^{(n)}$. Finally $D$ is called base point free if and only if $m_\sigma\in\Delta_D$ for all $\sigma\in\Sigma^{(n)}$. Base point freedom is a sufficient condition for a hypersurface defined by $D$ to be regular: Bertini's theorem states that the singular points of $D$ are the base locus and the singular points inherited from the ambient space. The absence of base points implies that $D$ can be deformed transversally in every point and therefore generically avoids the singularities of the ambient space. Thus, a base point free $D$ is regular. We emphasize however that base point freedom is not a necessary condition for the regularity of $D$.
Equations for hypersurfaces or complete intersections are sections of line bundles $\mathcal{O}(D)$ given by the following Laurent polynomial:
\begin{equation}
f=\sum_{m\in\Delta_D\cap M}c_m\chi^m=\sum_{m\in\Delta_D\cap M}c_m\prod_jz_j^{\langle m,v_j\rangle}
\end{equation}
In an affine patch $U_{\sigma}$ the local section $f_{\sigma}=f/\chi^{m_{\sigma}}$ is a regular function. Given a polytope $\Delta_D\in M$, we can define the polar polytope $\Delta_D^{\circ}$ by $\Delta_D^{\circ}=\{y\in N_{\mathbb{R}}:\:\langle x,y\rangle\geq -1\:\forall x\in\Delta_D\}$. It can be shown \cite{batyrev93} that the Calabi-Yau condition for hypersurfaces requires that $\Delta_D\subseteq M_{\mathbb{R}}$ is polar to $\Delta^{\ast}=\Delta_D^{\circ}\subseteq N_{\mathbb{R}}$, where $\Delta^{\ast}$ is the convex hull of the $v_j\in N$ as defined in section \ref{sec-toricdef}. A lattice polytope whose polar polytope is again a lattice polytope is called reflexive. For reflexive polytopes $(\Delta,\Delta^{\circ})$ there exists a combinatorial formula for the Hodge numbers \cite{batyrev93}:
\begin{equation}
h_{1,1}(X_{\Delta})=h_{\mathrm{dim}\Delta-2,1}(X_{\Delta^{\circ}})=l(\Delta^{\circ})-1-\mathrm{dim}\Delta-\sum_{\mathrm{codim}(\theta^{\circ})=1}l^{\ast}(\theta^{\circ})+\sum_{\mathrm{codim}(\theta^{\circ})=2}l^{\ast}(\theta^{\circ})l^{\ast}(\theta)
\end{equation}
where $\theta$ and $\theta^{\circ}$ is a dual pair of faces of $\Delta$ and $\Delta^{\circ}$. Furthermore, $l(\theta)$ is the number of lattice points of a face $\theta$ and $l^{\ast}(\theta)$ is the number of its interior lattice points.
In our discussion of F-theory model building we also encounter complete intersection Calabi-Yaus. The concept of polar pairs of reflexive lattice polytopes can be generalized as follows:
\begin{eqnarray} \label{cicy-nef}
\Delta=\Delta_1+\ldots+\Delta_r&& \Delta^{\circ}
=\langle \nabla_1,\ldots,\nabla_r\rangle_{\mathrm{conv}}\nonumber\\
&(\nabla_n,\Delta_m)\geq-\delta_{nm}&\\
\nabla^{\circ}=\langle \Delta_1,\ldots,\Delta_r\rangle_{\mathrm{conv}}
&&\nabla=\nabla_1+\ldots+\nabla_r\nonumber
\end{eqnarray}
Here $r$ is the codimension of the Calabi-Yau and the defining equations $f_i=0$ are sections of $\mathcal{O}(\Delta_i)$. The decomposition of the $M$-lattice polytope $\Delta\subset M_{\mathbb{R}}$ into a Minkowski sum\footnote{The Minkowski sum $A+B$ of two sets $A,B$ is defined as follows: $A+B=\{a+b|a\in A, b\in B\}$.} $\Delta=\Delta_1+\ldots+\Delta_r$ is dual to a nef (numerically effective) partition of the vertices of a reflexive polytope $\nabla\subset N_{\mathbb{R}}$ such that the convex hulls $\langle\nabla_i\rangle_{\mathrm{conv}}$ of the respective vertices and $0\in N$ only intersect at the origin. The nef-property means that the restriction of the line bundles associated to the divisors specified by the N-lattice points to any algebraic curve of the variety are non-negative. There exists a combinatorial formula for the Hodge numbers \cite{batyrevborisov} which has been implemented in PALP.
In many string theory applications, and in particular also in F-theory, the fibration structure of a Calabi-Yau manifold is of great interest. For Calabi-Yaus which can be described in terms of toric geometry the fibration structure can be deduced from the geometry of the lattice polytopes. If we are looking for toric fibrations where the fibers are Calabi-Yaus of lower dimensions, we have to search for reflexive sub-polytopes of $\Delta^{\circ}$ which have appropriate dimension. Given a base $b$ and a fiber $f$, the fibrations descend from toric morphisms of the ambient spaces corresponding to a map $\phi:\Sigma\rightarrow\Sigma_b$ of fans in $N$ and $N_b$, where $\phi:N\rightarrow N_b$ is a lattice homomorphism such that for each cone $\sigma\in\Sigma$ there is a cone $\sigma_b\in\Sigma_b$ that contains the image of $\sigma$. The lattice $N_f$ for the fiber is the kernel of $\phi$ in $N$. The fiber polytope is then defined as follows: $\Delta_f^{\circ}=\Delta^{\circ}\cap N_f$. In order to guarantee the existence of a projection one must find a triangulation of $\Delta_f^{\circ}$ and extend it to a triangulation of $\Delta^{\circ}$. For each choice of triangulation the homogeneous coordinates corresponding to the rays in $\Delta_f^{\circ}$ can be interpreted as coordinates of the fiber.
\subsection{Intersection ring and Mori cone}
\label{sec-intmori}
Two further pieces of data that are necessary in many string theory calculations are the intersection numbers of the toric divisors and the Mori cone, which is the dual of the K\"ahler cone. Inside the K\"ahler cone the volumes such as (\ref{volumes}) are positive. Thus, in the context of F-theory model building the K\"ahler cone is needed in order to make statements about a decoupling limit.
Let us start with discussing the intersection ring. For a compact toric variety $X_{\Sigma}$ the intersection ring is of the form $\mathbb{Z}[D_1,\ldots,D_r]/\langle I_{lin}+I_{non-lin}\rangle$. The two ideals to be divided out take into account linear and non-linear relations between the divisors. The linear relations have the form $\sum_j\langle m,v_j\rangle D_j$, where $m\in M$ form a set of basis vectors in the M-lattice. The non-linear relations are denoted by $R=\cup R_I$ where the $R_I$ are of the form $R_I=D_{j_1}\cdot\ldots\cdot D_{j_k}=0$. They come from the exceptional set $Z=\cup Z_I$ defined in section \ref{sec-toricdef}, which determines which homogeneous coordinates are not allowed to vanish at the same time. As mentioned before, this is the case when a collection of rays $\rho_{j_1},\ldots,\rho_{j_k}\in N$ is not contained in a single cone. The non-linear relations $R$ generate the ideal $I_{non-lin}$ which is called Stanley-Reisner ideal. Thus, the intersection ring $A_{\ast}(\Sigma)$ of a non-singular toric variety has the following form:
\begin{equation}
\label{intring}
A_{\ast}(X_{\Sigma})=\mathbb{Z}[D_1,\ldots,D_r]/\langle R,\sum_j\langle m,v_j\rangle D_j\rangle
\end{equation}
The definition of the intersection ring holds for non-singular toric varieties but may be generalized to the case where $X_{\Sigma}$ is simplicial projective. This means that the toric variety may be singular but still all the cones of the fan $\Sigma$ are simplicial. Such a situation may occur for example if we choose a non-maximal triangulation of the polytope $\Delta^{\ast}$. In this case the intersection numbers take values in $\mathbb{Q}$. To compute the Stanley-Reisner ideal in the non-singular case one must find a maximal triangulation of the fan $\Sigma$ or the polytope $\Delta^{\ast}$. In order to get intersection numbers we still have to fix a normalization: for a maximal simplicial cone $\sigma\in\Sigma^{(n)}$ spanned by $v_{j_1},\ldots,v_{j_n}$ we fix the intersection numbers of the corresponding divisors to be $D_{j_1}\cdot\ldots D_{j_n}=1/\mathrm{Vol}(\sigma)$, where $\mathrm{Vol}(\sigma)$ is the lattice volume of $\sigma$ (i.e. the geometric volume divided by the volume $1/n!$ of a basic simplex). If $X$ is non-singular the volume is $1$. Using the intersection ring one can compute the total Chern class of the tangent bundle $T_X$ of $X$ which is given by the following formula: $c(T_X)=\prod_{j=1}^r(1+D_j)$.
So far, we have only discussed the intersection ring of the toric variety $X$. However in many applications we rather need the intersection numbers for divisors on a hypersurface given by a divisor $D$ in $X$. Here we can make use of the restriction formula that relates the intersection form on the hypersurface divisor to the intersection form on $X$:
\begin{equation}
D_{j_1}\cdot\ldots\cdot D_{j_{n-1}}|_D=D_{j_1}\cdot\ldots\cdot D_{j_{n-1}}\cdot D|_X
\end{equation}
This allows us to compute the intersection ring of $D$ from the intersection ring of $X$. In (\ref{intring}) restriction to $D$ amounts to computing the ideal quotients of $I_{lin}$ and $I_{non-lin}$ with the ideal generated by $D$. By adjunction the Chern class for the hypersurface specified by $D$ is $c(D)=\prod_{j=1}^r(1+D_j)/(1+D)$.
In order to be able to calculate all the quantities defined in section \ref{sec-fgeom} we miss one more ingredient: the Mori cone. By definition, the Mori cone is the dual of the K\"ahler cone. We need the information about the K\"ahler cone in order to be able to compute the volumes of divisors. By definition the volumes will be positive inside the K\"ahler cone. The Mori cone is generated by $l^{(1)},\ldots,l^{(k)}$, where $k=r-n$ if the fan $\Sigma$ is simplicial. Otherwise the number of Mori generators can be larger. The Mori cone $L$ is then defined as follows: $L=\mathbb{R}_{\geq 0}l^{(1)}+\ldots+\mathbb{R}_{\geq 0}l^{(k)}$. For the calculation of the Mori cone we also require a maximal triangulation of $\Delta^{\ast}$. Given such a triangulation the Mori generators can be determined as follows \cite{Berglund:1995gd}: take every pair of $n$-dimensional simplices $(S_k,S_l)$ which have a common $n-1$-dimensional simplex $s_{kl}=S_k\cap S_l$. Then find the unique linear relation $\sum_il_i^{k,l}v_i=0$ with $v_i\in S_k\cup S_l$ where the $l_i^{k,l}$ are minimal integers and the coefficients of the points in $(S_k\cup S_l)\backslash (S_k\cap S_l)$ are non-negative. The Mori generators are then the minimal integers $l^{(a)}$ by which every $l^{k,l}$ can be expressed as positive integer linear combinations. There is an equivalent algorithm to determine the Mori generators due to Oda and Park \cite{OdaPark} which has been implemented in an unreleased version of PALP \cite{maxnils}. Note that the relations $\sum_{i=1}^rl_i^{(a)}D_i=0$ define the ideal $I_{lin}$ in (\ref{intring}). Assembling the Mori vectors into a $k\times r$-matrix, the columns of the matrix encode inequalities for the values of the K\"ahler parameters. Solving these inequalities yields a basis $K_i$ of the K\"ahler cone such that the K\"ahler form of $X$ can be written as $J=\sum_i r_iK_i$ with $r_i>0$. Note that this prescription computes the K\"ahler cone of the toric variety $X$. It is often assumed that this is a good approximation for the K\"ahler cone of a hypersurface in $X$.
\subsection{Toric Calculations using PALP and other Software}
In string theory and F-theory we deal with compactifications on Calabi-Yau threefolds and fourfolds. In F-theory model building the base manifold $B$ is a hypersurface in a four-dimensional toric ambient space. The fourfolds are complete intersections in a six-dimensional toric space. The associated lattice polytopes live in four- and six-dimensional integer lattices and typically have a large number of points. It is in general not possible to do calculations without computer support. There exist several software packages which are useful for particular aspects in toric geometry. In this section we will mostly focus on the program PALP \cite{Kreuzer:2002uu}. Before that, let us mention some other useful programs: Schubert by Katz and Str\o mme is a Maple package for calculations in intersection theory. TOPCOM \cite{Rambau:TOPCOM-ICMS:2002} computes triangulations of point configurations. Singular \cite{DGPS} is a powerful computer algebra program which is optimized for calculations with polynomial rings, such as the intersection ring. A recent addition is cohomCalg \cite{Blumenhagen:2010pv} which can compute line bundle-valued cohomology classes over toric varieties.
Let us now discuss some features and applications of PALP \cite{Kreuzer:2002uu}, which stands for ``Package for Analyzing Lattice Polytopes''. It consists of several programs.
\begin{itemize}
\item {\tt poly.x} computes the data of a lattice polytope and its dual if the polytope is reflexive. The input can be either a weight matrix or the points of a polytope in the M-lattice or the N-lattice. Apart from the polytope data {\tt poly.x} computes Hodge numbers of the associated Calabi-Yau hypersurfaces, information about fibrations and other data. {\tt poly.x} has been extended with several features that include information about the facets of the polytope, data of Fano varieties and conifold Calabi-Yaus. In \cite{Batyrev:2008rp,Kreuzer:2009is} this extension of PALP has been used to find new Calabi-Yau manifolds with small $h^{1,1}$ which are obtained from known Calabi-Yau threefolds via conifold transitions. The full set of options in PALP can be obtained with {\tt poly.x -h} and {\tt poly.x -x} for extended options.
\item The program {\tt nef.x} can be used for complete intersection Calabi-Yaus. It takes the same input as {\tt poly.x} and computes the polytope data, nef partitions and Hodge numbers as well as information about fibrations. There are several extended options which include most notably the data of the Gorenstein cones (cf. \cite{Batyrev:1994ac} for the definition and construction in toric geometry) in the $M/N$-lattice.
\item {\tt cws.x} creates weight systems and combined weight systems of polytopes of dimension to be specified in the input.
\item {\tt class.x} classifies reflexive polytopes by searching for sub-polytopes of a Newton polytope associated to a combined weight system.
\end{itemize}
Apart from recent applications in F-theory model building, which we will discuss in the next section, PALP has been used in many other contexts. A data base of Calabi-Yau threefolds has been generated by listing all $473\,800\,776$ reflexive polyhedra in four dimensions \cite{Kreuzer:2000xy}. In view of the landscape problem in string theory the statistics of the polytope data is also of interest \cite{Kreuzer:2008nu}. Some of the most recent extensions of PALP which we will mention below have already been used in \cite{Collinucci:2008sq,Chen:2010ts,Knapp:2011wk}.
\subsection{Application to F-theory GUTs}
In this section we make the connection to F-theory model building and discuss how the calculations discussed in section \ref{sec-fgeom} can be carried out explicitly. The approach discussed here is used in \cite{Chen:2010ts,Knapp:2011wk}. Our aim is a systematic construction of a large class of examples of global F-theory models. The first step is the construction of the base manifold $B$. In \cite{Chen:2010ts} we have obtained a set of geometries by systematically constructing weight matrices associated to point and curve blowups on Fano hypersurfaces in $\mathbb{P}_4$. In \cite{Knapp:2011wk} we have considerably extended this class of models by defining hypersurfaces in a subset of the toric ambient spaces described by the $473\,800\,776$ reflexive polyhedra in four dimensions \cite{Kreuzer:2000xy}. Concretely, we have restricted ourselves to configurations where the N-lattice polytopes have at most nine points. As one can check for example at \cite{database}, there are $1088$ such polytopes. We used PALP to recover the toric data of the ambient space and considered all non-negatively curved hypersurfaces in these ambient spaces. In order to be able to perform the calculations outlined in section \ref{sec-fgeom} we must compute the intersection ring and the Mori cone. We have achieved this by using an extended version of {\tt poly.x} \cite{maxnils}. The following additional features have been implemented: processing of non-Calabi-Yau hypersurfaces by specifying the hypersurface degrees as input parameters, a calculation of the maximal triangulations of the $N$-lattice polytope, calculation of the Mori cone and the Stanley-Reisner ideal, and calculation of the intersection ring with the help of Singular. Using this data we can identify del Pezzo divisors, check the existence of a decoupling limit and compute the topological properties of matter curves and Yukawa points. In \cite{Knapp:2011wk} we have analyzed a total number of $569\:674$ base manifolds. The resulting geometries are available at \cite{data}.
The next step in the calculation is to construct the Calabi-Yau fourfold $X_4$ which is an elliptic fibration over the base $B$. The toric data of $X_4$ is obtained by extending the weight matrix of $B$. Schematically, this looks as follows:
\begin{equation}
\begin{array}{cccccc}
3&2&1&0&\cdots&0\\
\ast&\ast&0&w_{11}&\cdots&w_{1n}\\
\ast&\ast&0&\cdots&\cdots&\cdots\\
\ast&\ast&0&w_{m1}&\cdots&w_{mn}\\
\end{array}
\end{equation}
Here the $w_{ij}$ denote the entries of the weight matrix associated to $B$. The $\ast$-entries in the extended weight matrix have to be chosen in such a way that the fiber coordinates $x,y$ are sections of $K_B^{-2}$ and $K_B^{-3}$, respectively. These entries of the fourfold weight matrix contain the information about the hypersurface degrees of the base. Not every extended weight system will lead to a Calabi-Yau fourfold of the form (\ref{cicydef}). The calculations can be done using {\tt nef.x}. Several problems can appear: first, there may be no nef partition and therefore our methods do not work. A second conceptual problem is that the polytope corresponding to the extended weight system is not always reflexive. Many of the combinatorial tools used in PALP are only valid for reflexive polytopes. Even though one might have a perfectly fine Calabi-Yau fourfold we cannot apply our technology to them. The third issue is of a technical nature: due to the complexity of the fourfold polytopes one may reach the software bounds of PALP which results in numerical overflows. For the $569\,674$ extended weight systems discussed in \cite{Knapp:2011wk} we find only $27\,345$ reflexive fourfold polytopes which have at least one nef partition. Furthermore there are $18\,632$ reflexive polytopes without a nef partition, $381\,232$ non-reflexive polytopes and $142\,470$ cases with numerical overflow.
Having found a reflexive fourfold polytope with at least one nef partition is not enough to have a good global F-theory model. If we further demand that the base $B$ has at least one del Pezzo divisor with a mathematical or physical decoupling limit the number of fourfolds decreases significantly. In addition we should also impose some constraints on the regularity of the base. Demanding that $B$ is Cartier leaves us with $16\,011$ good models. Imposing the stronger criterion of base point freedom we are down to $7386$ models. Focusing on these $7\,386$ good geometries we apply the constraint that the nef partition should be compatible with the elliptic fibration. This information can be extracted from the output of {\tt nef.x}. This further reduces the number of geometries to $3978$.
Having found a good Calabi-Yau fourfold, we can construct a GUT model on every (del Pezzo) divisor. A toric description on how to impose a specific GUT group on a Tate model has been given in \cite{Blumenhagen:2009yv}. The Tate form~(\ref{tate}) implies that the $a_n$ appear in the monomials which contain $z^n$. We can isolate these monomials by identifying the vertex $\nu_z$ in $(\nabla_1,\nabla_2)$ that corresponds to the $z$-coordinate in the Tate model. All the monomials that contain $z^r$ are then in the following set:
\begin{equation}
A_r=\{w_k\in\Delta_m\::\:\langle\nu_z,w_k\rangle+1=r\}\qquad \nu_z\in\nabla_m,
\end{equation}
where $\Delta_m$ is the dual of $\nabla_m$, which denotes the polytope containing the $z$-vertex. The polynomials $a_r$ are then given by the following expressions:
\begin{equation}
a_r=\sum_{w_k\in A_r}c_k^{m}\prod_{n=1}^2\prod_{\nu_i\in\nabla_n}y_i^{\langle \nu_i,w_k\rangle+\delta_{mn}}\vert_{x=y=z=1}
\end{equation}
Now we can remove all the monomials in $a_r$ which do not satisfy the factorization constraints (\ref{su5so10tate}) of the Tate algorithm. In order to perform this calculation we have to identify the fiber coordinates $(x,y,z)$ and the GUT coordinate $w$ within the weight matrix of the fourfold. We have applied this procedure to every del Pezzo divisor in the $3978$ ``good'' fourfold geometries. Note that the procedure described above can destroy the reflexivity of the polytope, which happens in about $30\%$ of the examples. For $SU(5)$-models we found $11\,275$ distinct models\footnote{Since the procedure has been applied to all del Pezzo divisors in a given base geometry not all these models may have a decoupling limit.} with reflexive polyhedra, for $SO(10)$ GUTs there are $10\,832$. $U(1)$-restricted GUT models \cite{Grimm:2010ez} can be engineered along the same lines. It turns out that $U(1)$-restriction does not put any further constraints on the reflexivity of the polytope.
\section{Outlook}
In this article we have discussed how toric geometry can be used to construct a large number of geometries that can support global F-theory GUTs. Using this technology we could show that elementary consistency constraints greatly reduce the number of possible models. However, due to computational constraints, we did not quite succeed in systematically listing all possible F-theory models within a class of geometries. Such an endeavor would require substantial changes in the computer programs we are using. It is actually quite remarkable that we could make use of PALP for Calabi-Yau fourfolds and non-Calabi-Yau threefolds, since this goes beyond what it was originally designed for.
Let us present a list of suggestions to extend PALP in order to improve the applicability to the current problems in mathematics and physics and to make it more accessible for users. The original purpose of PALP was to solve a classification problem for polytopes. Over the years it has been adjusted and extended in order to be applied to specific problems. Many of the basic routines that were implemented to tackle some special questions could be used in much more general contexts but cannot be easily accessed. Therefore a better modularization of the software is necessary in order to have flexible access to these basic routines. Another problem of PALP is that one has to specify several parameters and bounds such as the number of points in a polytope in a given dimension at the compilation of the program. It would be practical to have fully dynamical dimensions in order to work with a precision tailored to the problem at hand without recompiling.
A fundamental change would be to step away from the description of polytopes and instead use the ray representation which has the full data of the cones. This is necessary if one wants to deal with non-reflexive polytopes. A further extension which has already been partially implemented is to include triangulations, intersection rings and even the calculation of Picard-Fuchs operators needed for mirror symmetry calculations into PALP. The ultimate goal is to have an efficient and versatile program which can be used for toric calculations of all kinds without having to rely on commercial software. Finally a detailed documentation of all the features of PALP would be helpful \cite{andnils}.
As for the search for F-theory models, an extended version of PALP would hopefully help to overcome the problems of non-reflexivity and overflows we have encountered in \cite{Knapp:2011wk}. Apart from finding new examples for physics applications one might also attempt a partial classification of Calabi-Yau fourfolds. Enumerating all toric Calabi-Yau fourfolds may be out of reach or even impossible but for finding all models of type (\ref{cicydef}) one can at least give a prescription for the construction: take each of the $473\,800\,776$ reflexive polyhedra in four dimensions and put in all non-negatively curved hypersurfaces that are not Calabi-Yau. Then construct fourfolds which are elliptic fibrations over these base manifolds. A rough estimate shows that this procedure would yield $\mathcal{O}(10^{11})$ fourfold geometries.\\\\
{\bf Acknowledgments:} In October 2010 Maximilian Kreuzer asked me to be the co-author of this review article. Sadly, he passed away on November 26, 2010 when this work was still in the early stages. I am grateful for many years of collaboration with Max, as well as for his constant support and encouragement.
I would like to thank my collaborators Ching-Ming Chen, Christoph Mayrhofer and Nils-Ole Walliser for a pleasant and fruitful collaboration on the projects \cite{Chen:2010ts,Knapp:2011wk} this article is based on. Furthermore I thank Emanuel Scheidegger for valuable comments on the manuscript. This work has been supported by World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan.
\addcontentsline{toc}{section}{References}
\providecommand{\href}[2]{#2}\begingroup\raggedright |
\section{Introduction}
Let $\mathbf{P}$ be the optimization problem $f^*= \min \:\{ f(\mathbf{x})\::\:\mathbf{x}\in\mathbf{K}\:\}$,
where
\begin{equation}
\label{setk}
\mathbf{K}:=\{\mathbf{x}\in\mathbb{R}^n\,:\, g_j(\mathbf{x})\geq0,\: j=1,\ldots,m\},
\end{equation}
for some polynomials $f,(g_j)\subset\mathbb{R}[\mathbf{x}]$. This framework is rather general as it encompasses
a large class of important optimization problems, including non convex and
discrete optimization problems.
Problem $\mathbf{P}$ is in general NP-hard and one is often satisfied with a local minimum
which can be obtained by running some local minimization algorithm among those available in the literature.
Typically in such algorithms, at a current iterate (i.e. some feasible solution $\mathbf{y}\in\mathbf{K}$), one checks whether
some optimality conditions (e.g. the Karush-Kuhn-Tucker (KKT) conditions) are satisfied within some
$\epsilon$-tolerance. However, as already mentioned those conditions are only valid
for a local minimum, and in fact, even only for a stationary point of the Lagrangian.
Moreover, in many practical situations
the criterion $f$ to minimize is subject to modeling errors
or is questionable. In such a situation, the practical meaning of a
local (or global) minimum $f^*$ (and local (or global) minimizer) also becomes
questionable. It could well be that the current solution $\mathbf{y}$ is in fact a global
minimizer of an optimization problem $\mathbf{P}'$ with same feasible set as $\mathbf{P}$
but with a different criterion $\tilde{f}$. Therefore, if $\tilde{f}$
is close enough to $f$, one might not be willing to spend
an enormous computing time and effort to find the global (or even local) minimum $f^*$
because one might be already satisfied with the current iterate $\mathbf{y}$ as a global minimizer of $\mathbf{P}'$.
{\bf Inverse Optimization} is precisely concerned with the above issue of determining
a criterion $\tilde{f}$ as close to $f$ as possible, and for which the current solution $\mathbf{y}$ is
an optimal solution of $\mathbf{P}'$ with this new criterion $\tilde{f}$.
Pioneering work in Control dates back to Freeman and Kokotovic
\cite{koko} for optimal stabilization. Whereas it was known that
every value function of an optimal stabilization problem is also a Lyapunov function for the closed-loop system,
in \cite{koko} the authors show the converse, that is, every Lyapunov
function for every stable closed-loop system is also a {\it value function} for a meaningful optimal stabilization problem. In optimization, pioneering works in this direction date back to Burton and Toint and \cite{burton} for shortest path problems,
and Zhang and Liu \cite{zhang1,zhang2}, Huang and Liu \cite{huang}, and
Ahuja and Orlin and \cite{ahuja} for linear programs in the form
$\min \{\mathbf{c}'\mathbf{x}\,:\,\mathbf{A}\mathbf{x}\geq \mathbf{b};\:\mathbf{r} \leq\mathbf{x}\leq \mathbf{s}\}$ (and with the $\ell_1$-norm). For the latter, the inverse problem is again a linear program of the same form. Similar results also hold
for inverse linear programs with the $\ell_\infty$-norm as shown in Ahuja and Orlin \cite{ahuja} while
Zhang et al. \cite{zhangetal} provide a column generation method for the inverse shortest path problem.
In Heuberger \cite{heuberger} the interested reader
will find a nice survey on inverse optimization for linear programming and combinatorial optimization problems.
For integer programming, Schaefer \cite{schaefer} characterizes the feasible set of
cost vectors $c\in\mathbb{R}^n$ that are candidates for inverse optimality. It is the projection on $\mathbb{R}^n$ of
a (lifted) convex polytope obtained from the super-additive dual of integer programs.
Unfortunately and as expected, the dimension of
of the lifted polyhedron (before projection) is exponential in the input size of the problem.
Finally, for linear programs Ahmed and Guan \cite{ahmed} have considered
the variant called {\it inverse optimal value} problem
in which one is interested in finding a linear criterion $c\in C\subset\mathbb{R}^n$ for which
the optimal value is the closest to a desired specified value. Perhaps surprisingly,
they proved that such a problem is NP-hard.\\
As the reader may immediately guess, in inverse optimization the main difficulty
lies in having a tractable characterization of global optimality
for a given current point $\mathbf{y}\in\mathbf{K}$ and some candidate criterion $\tilde{f}$. This is why
most of all the above cited works address linear programs or
combinatorial optimization problems for which some
characterization of global optimality is available and can be (sometimes) effectively
used for practical computation. For instance,
the characterization of global optimality for integer programs described in Schaefer \cite{schaefer} is
via the superadditive dual of Wolsey \cite[\S 2]{wolsey} which is exponential in the
problem size, and so prevents from its use in practice.
This perhaps explains why inverse (non linear) optimization has not attracted much attention in the past, and
it is a pity since inverse optimality could provide an alternative stopping criterion
at a feasible solution $\mathbf{y}$ obtained by a (local) optimization algorithm.
The novelty of the present paper is to provide
a systematic numerical scheme for computing an inverse optimal solution associated with
the polynomial program $\mathbf{P}$ and a given feasible solution $\mathbf{y}\in\mathbf{K}$.
It consists of solving a semidefinite program\footnote{A semidefinite program is a convex (conic)
optimization problem that can be solved efficiently. For instance, up to arbitrary (fixed) precision and using some interior point algorithms,
it can be solved in time polynomial in the input size of the problem. For more details the interested reader is referred to
e.g. Wolkowicz et al. \cite{handbook} and the many references therein.}
whose size can be adapted to
the problem on hand, and so is tractable (at least for moderate size problems
and possibly for larger size problems if sparsity is taken into account).
Moreover, if one uses the $\ell_1$-norm
then the inverse-optimal objective function exhibits a simple and remarkable {\it canonical} (and sparse) form.
{\bf Contribution.} In this paper we investigate the inverse optimization problem for
polynomial optimization problems $\mathbf{P}$ as in (\ref{setk}), i.e., in a rather general context which
includes nonlinear and nonconvex optimization problems and in particular,
0/1 and mixed integer nonlinear programs. Fortunately, in such a context,
Putinar's Positivstellensatz \cite{putinar} provides us with a very
powerful certificate of global optimality that can be adapted to
the actual computational capabilities for a given problem size.
More precisely, and assuming $\mathbf{y}=0$ (possibly after a change of variable $\mathbf{x}'=\mathbf{x}-\mathbf{y}$),
in the methodology that we propose,
one computes the coefficients of a polynomial
$\tilde{f}_d\in\mathbb{R}[\mathbf{x}]$ of same degree $d_0$ as $f$ (or possibly larger degree if desired and/or possibly with some
additional constraints), such that:
\begin{itemize}
\item
$0$ is a global minimizer
of the related problem $\min_\mathbf{x}\{ \tilde{f}_d(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$, with a Putinar's certificate of optimality with degree bound $d$ (to be explained later).
\item $\tilde{f}_d$ minimizes $\Vert \tilde{f}-f\Vert_k$ (where depending on $k$,
$\Vert \cdot\Vert_k$ is the $\ell_1$, $\ell_2$ or $\ell_\infty$-norm of the coefficient vector) over all polynomals $\tilde{f}$ of degree $d_0$, having the previous property.
\end{itemize}
Assuming $\mathbf{K}$ is compact (hence $\mathbf{K}\subseteq [-1,1]^n$ possibly after a change of variable),
it turns out that the optimal value $\rho_d:=\Vert \tilde{f}_d-f\Vert_k$ also measures how close is $f(0)$ to the global optimum
$f^*$ of $\mathbf{P}$, as we also obtain that $f^*\leq f(0)\leq f^*+\rho_d$ if $k=1$ and similarly
$f^*\leq f(0)\leq f^*+\rho_d\,{n+d_0\choose n}$ if $k=\infty$.
In addition, for the $\ell_1$-norm we prove that
$\tilde{f}_d$ has a simple {\it canonical form}, namely
\[\tilde{f}_d\,=\,f+\mathbf{b}'\mathbf{x}+\sum_{i=1}^n\lambda_i\,x_i^2,\]
for some vector $\mathbf{b}\in\mathbb{R}^n$, and nonnegative vector $\lambda\in\mathbb{R}^n$, optimal solutions of a semidefinite program.
(For 0/1 problems it further simplifies to $\tilde{f}_d=f+\mathbf{b}'\mathbf{x}$ only.)
This canonical form is sparse as $\tilde{f}_d$ differs from $f$ in at most $2n$ entries only ($\ll {n+d_0\choose n}$). It illustrates
the sparsity properties of optimal solutions of $\ell_1$-norm minimization problems, already observed in other contexts
(e.g., in some compressed sensing applications).
Importantly, to compute $\tilde{f}_d$, one has to solve a semidefinite program of size parametrized by $d$,
where $d$ is chosen so that the size of semidefinite program associated with Putinar's certificate
(with degree bound $d$) is compatible with current semidefinite solvers available. (Of course,
even if $d$ is relatively small, one is still restricted to problems of relatively modest size.)
Moreover, when $\mathbf{K}$ is compact, generically
$\tilde{f}_d$ is an optimal solution of the ``ideal inverse optimization problem"
provided that $d$ is sufficiently large!
In addition, one may also consider several additional options:
$\bullet$ Instead of looking for a polynomial $\tilde{f}$ of same degree as $f$,
one might allow polynomials of higher degree, and/or restrict
certain coefficients of $\tilde{f}$ to be the same as those of $f$ (e.g. for
structural modeling reasons).
$\bullet$ One may restrict $\tilde{f}$ to a certain class of functions, e.g.,
quadratic polynomials and even convex quadratic polynomials.
In the latter important case and if the $g_j$'s that define $\mathbf{K}$ are concave,
the procedure to compute an optimal solution
$\tilde{f}(\mathbf{x})=\tilde{\mathbf{b}}'\mathbf{x}+\mathbf{x}'\tilde{\mathbf{Q}}\mathbf{x}$ simplifies and reduces to solving separately a linear program
(for computing $\tilde{\mathbf{b}}$) and a semidefinite program (for computing $\tilde{\mathbf{Q}}$).
The paper is organized as follows. In a first introductory section we present the notation, definitions, and
the ideal inverse optimization problem. We then describe how a practical inverse optimization problem
reduces to solving a semidefinite program and exhibit the canonical form of the optimal solution for the $\ell_1$-norm.
We also provide additional results, e.g., an asymptotic analysis when the degree bound in Putinar's certificate
increases and also the particular case where one searches for a convex candidate criterion.
\section{Notation, definitions and preliminaries}
\subsection{Notation and definitions}
Let $\mathbb{R}[\mathbf{x}]$ (resp. $\mathbb{R}[\mathbf{x}]_d$) denote the ring of real polynomials in the variables
$\mathbf{x}=(x_1,\ldots,x_n)$ (resp. polynomials of degree at most $d$), whereas $\Sigma[\mathbf{x}]$ (resp. $\Sigma[\mathbf{x}]_d$) denotes
its subset of sums of squares (s.o.s.) polynomials (resp. of s.o.s. of degree at most $2d$).
For every
$\alpha\in\mathbb{N}^n$ the notation $\mathbf{x}^\alpha$ stands for the monomial $x_1^{\alpha_1}\cdots x_n^{\alpha_n}$ and for every $i\in\mathbb{N}$, let $\mathbb{N}^{p}_d:=\{\beta\in\mathbb{N}^n:\sum_j\beta_j\leq d\}$ whose cardinal is $s(d)={n+d\choose n}$.
A polynomial $f\in\mathbb{R}[\mathbf{x}]$ is written
\[\mathbf{x}\mapsto f(\mathbf{x})\,=\,\sum_{\alpha\in\mathbb{N}^n}\,f_\alpha\,\mathbf{x}^\alpha,\]
and $f$ can be identified with its vector of coefficients $\mathbf{f}=(f_\alpha)$ in the canonical basis $(\mathbf{x}^\alpha)$, $\alpha\in\mathbb{N}^n$. Denote by $\mathcal{S}^t\subset\mathbb{R}^{t\times t}$ the space of real symmetric matrices,
and for any $\mathbf{A}\in\mathcal{S}^t$ the notation $\mathbf{A}\succeq0$ stands for $\mathbf{A}$ is positive semidefinite.
For $f\in\mathbb{R}[\mathbf{x}]_d$, let
\[\Vert f\Vert_k\,=\,\left\{\begin{array}{ll}
\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d}}\vert f_\alpha\vert&\mbox{if $k=1$,}\\
\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d}} f_\alpha^2&\mbox{ if $k=2$,}\\
\displaystyle\max\,\{\vert f_\alpha\vert\,:\:\alpha\in\mathbb{N}^n_{d}\}&\mbox{ if $k=\infty$.}\end{array}\right.\]
A real sequence $\mathbf{z}=(z_\alpha)$, $\alpha\in\mathbb{N}^n$, has a {\it representing measure} if
there exists some finite Borel measure $\mu$ on $\mathbb{R}^n$ such that
\[z_\alpha\,=\,\int_{\mathbb{R}^n}\mathbf{x}^\alpha\,d\mu(\mathbf{x}),\qquad\forall\,\alpha\in\mathbb{N}^n.\]
Given a real sequence $\mathbf{z}=(z_\alpha)$ define the linear functional $L_\mathbf{z}:\mathbb{R}[\mathbf{x}]\to\mathbb{R}$ by:
\[f\:(=\sum_\alpha f_\alpha\mathbf{x}^\alpha)\quad\mapsto L_\mathbf{z}(f)\,=\,\sum_{\alpha}f_\alpha\,z_\alpha,\qquad f\in\mathbb{R}[\mathbf{x}].\]
\subsection*{Moment matrix}
The {\it moment} matrix associated with a sequence
$\mathbf{z}=(z_\alpha)$, $\alpha\in\mathbb{N}^n$, is the real symmetric matrix $\mathbf{M}_d(\mathbf{z})$ with rows and columns indexed by $\mathbb{N}^n_d$, and whose entry $(\alpha,\beta)$ is just $z_{\alpha+\beta}$, for every $\alpha,\beta\in\mathbb{N}^n_d$.
Alternatively, let
$\mathbf{v}_d(\mathbf{x})\in\mathbb{R}^{s(d)}$ be the vector $(\mathbf{x}^\alpha)$, $\alpha\in\mathbb{N}^n_d$, and
define the matrices $(\mathbf{B}_\alpha)\subset\mathcal{S}^{s(d)}$ by
\begin{equation}
\label{balpha}
\mathbf{v}_d(\mathbf{x})\,\mathbf{v}_d(\mathbf{x})^T\,=\,\sum_{\alpha\in\mathbb{N}^n_{2d}}\mathbf{B}_\alpha\,\mathbf{x}^\alpha,\qquad\forall\mathbf{x}\in\mathbb{R}^n.\end{equation}
Then $\mathbf{M}_d(\mathbf{z})=\sum_{\alpha\in\mathbb{N}^n_{2d}}z_\alpha\,\mathbf{B}_\alpha$.
If $\mathbf{z}$ has a representing measure $\mu$ then
$\mathbf{M}_d(\mathbf{z})\succeq0$ because
\[\langle\mathbf{f},\mathbf{M}_d(\mathbf{z})\mathbf{f}\rangle\,=\,\int f^2\,d\mu\,\geq0,\qquad\forall \,\mathbf{f}\,\in\mathbb{R}^{s(d)}.\]
\subsection*{Localizing matrix}
With $\mathbf{z}$ as above and $g\in\mathbb{R}[\mathbf{x}]$ (with $g(\mathbf{x})=\sum_\gamma g_\gamma\mathbf{x}^\gamma$), the {\it localizing} matrix associated with $\mathbf{z}$
and $g$ is the real symmetric matrix $\mathbf{M}_d(g\,\mathbf{z})$ with rows and columns indexed by $\mathbb{N}^n_d$, and whose entry $(\alpha,\beta)$ is just $\sum_{\gamma}g_\gamma z_{\alpha+\beta+\gamma}$, for every $\alpha,\beta\in\mathbb{N}^n_d$.
Alternatively, let $\mathcal{C}_\alpha\in\mathcal{S}^{s(d)}$ be defined by:
\begin{equation}
\label{calpha}
g(\mathbf{x})\,\mathbf{v}_d(\mathbf{x})\,\mathbf{v}_d(\mathbf{x})^T\,=\,\sum_{\alpha\in\mathbb{N}^n_{2d+{\rm deg}\,g}}\mathcal{C}_\alpha\,\mathbf{x}^\alpha,\qquad\forall\mathbf{x}\in\mathbb{R}^n.\end{equation}
Then $\mathbf{M}_d(g\,\mathbf{z})=\sum_{\alpha\in\mathbb{N}^n_{2d+{\rm deg}g}}z_\alpha\,\mathcal{C}_\alpha$.
If $\mathbf{z}$ has a representing measure $\mu$ whose support is
contained in the set $\{\mathbf{x}\,:\,g(\mathbf{x})\geq0\}$ then
$\mathbf{M}_d(g\,\mathbf{z})\succeq0$ because
\[\langle\mathbf{f},\mathbf{M}_d(g\,\mathbf{z})\mathbf{f}\rangle\,=\,\int f^2\,g\,d\mu\,\geq0,\qquad\forall \,\mathbf{f}\,\in\mathbb{R}^{s(d)}.\]
With $\mathbf{K}$ as in (\ref{setk}), let $g_0\in\mathbb{R}[\mathbf{x}]$ be the constant polynomial $\mathbf{x}\mapsto g_0(\mathbf{x})=1$, and
for every $j=0,1,\ldots,m$, let $v_j:=\lceil ({\rm deg}\,g_j)/2\rceil$.
\begin{defn}
With $d,k\in\mathbb{N}$ and $\mathbf{K}$ as in (\ref{setk}), let $Q_k(g)\subset\mathbb{R}[\mathbf{x}]$ and $Q^d_k\subset\mathbb{R}[\mathbf{x}]_d$ be the convex cones:
\begin{eqnarray}
\label{put-suff?}
Q(g)&:=&\left\{\:\sum_{k=0}^m \sigma_j\,g_j\::\: \sigma_j\in\Sigma[\mathbf{x}]\quad j=1,\ldots,m\:\right\}.\\
\label{put-suff}
Q_k(g)&:=&\left\{\:\sum_{k=0}^m \sigma_j\,g_j\::\: \sigma_j\in\Sigma[\mathbf{x}]_{k-v_j},\quad j=1,\ldots,m\:\right\}.\\
\label{puT-suff1}
Q_k^d(g)&:=&Q_k(g)\,\cap\,\mathbb{R}[\mathbf{x}]_d
\end{eqnarray}
We say that every element $h\in Q_k(g)$ has a {\it Putinar's certificate} of nonnegativity on $\mathbf{K}$, with degree bound $k$.
\end{defn}
The cone $Q(g)$ is called the quadratic module associated with the $g_j$'s.
Obviously, if $h\in Q(g)$ the associated s.o.s. polynomials $\sigma_j$'s provide a certificate of nonnegativity of $h$ on $\mathbf{K}$.
The cone $Q(g)$ is said to be {\it Archimedean} if and only if
\begin{equation}
\label{archimedean}
\mathbf{x}\mapsto M-\Vert\mathbf{x}\Vert^2\,\in\,Q(g)\quad\mbox{for some $M>0$.}
\end{equation}
Let ${\rm Psd}_d(\mathbf{K})\subset\mathbb{R}[\mathbf{x}]_d$ be the convex cone of polynomials of degree at most $d$, nonnegative on $\mathbf{K}$.
The name ``Putinar's certificate" is coming from the following Putinar's Positivstellensatz.
\begin{thm}[Putinar's Positivstellensatz \cite{putinar}]
\label{th-put}
Let $\mathbf{K}$ be as in (\ref{setk}) and assume that $Q(g)$ is Archimedean.
Then every polynomial $f\in\mathbb{R}[\mathbf{x}]$ strictly positive on $\mathbf{K}$ belongs to $Q(g)$. In addition,
\begin{equation}
\label{limit}
{\rm cl}\,\left(\bigcup_{k=0}^\infty Q^d_k(g)\right)\,=\,{\rm Psd}_d(\mathbf{K}),\qquad \forall d\in\mathbb{N}.
\end{equation}
\end{thm}
The first statement is just Putinar's Positivstellensatz \cite{putinar} whereas the second statement
is an easy consequence. Indeed let $f\in{\rm Psd}_d(\mathbf{K})$. If $f>0$ on $\mathbf{K}$ then
$f\in Q^d_k(g)$ for some $k$. If $f(\mathbf{x})=0$ for some $\mathbf{x}\in\mathbf{K}$, let $f_n:=f+1/n$, so that
$f_n>0$ on $\mathbf{K}$ for every $n\in\mathbb{N}$. But then $f_n\in\cup_{k=0}^\infty Q^d_k(g)$ and
the result follows because $\Vert f_n-f\Vert_1\to0$ as $n\to\infty$.
In fact, by results from Marshall \cite{marshall2} and more recently Nie \cite{nie},
membership in $Q(g)$ is also {\it generic} for polynomials that are only nonnegative on $\mathbf{K}$.
And so Putinar's Positivstellensatz is particularly useful to {\it certify} and enforce that
a polynomial is nonnegative on $\mathbf{K}$, and in particular the polynomial
$\mathbf{x}\mapsto f(\mathbf{x})-f(\mathbf{y})$ for the inverse optimization problem associated with a feasible solution $\mathbf{y}\in\mathbf{K}$.
Notice that one may also be less demanding and ask $\mathbf{y}$ to be only a global $\epsilon$-minimizer
for some fixed $\epsilon>0$. Again Putinar's Positivstellensatz is exactly what we need
to {\it certify} (global) $\epsilon$-optimality by requiring $f(\cdot)-f(\mathbf{y})+\epsilon\in Q_k(g)$.
\subsection{The ideal inverse problem}
Let $\mathbf{P}$ be the global optimization problem
$f^*=\min_\mathbf{x}\{f(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$ with $\mathbf{K}\subset\mathbb{R}^n$ as in (\ref{setk}),
and $f\in\mathbb{R}[\mathbf{x}]_{d_0}$ where $d_0:={\rm deg}\,f$.
Identifying a polynomial $f\in\mathbb{R}[\mathbf{x}]_{d_0}$ with its vector of coefficients
$\mathbf{f}=(f_\alpha)\in\mathbb{R}^{s(d_0)}$, one may and will identify $\mathbb{R}[\mathbf{x}]_{d_0}$ with the vector space
$\mathbb{R}^{s(d_0)}$, i.e., $\mathbb{R}[\mathbf{x}]_{d_0}\ni f\leftrightarrow \mathbf{f}\in\mathbb{R}^{s(d_0)}$. Similarly,
the convex cone ${\rm Psd}_{d_0}(\mathbf{K})\subset\mathbb{R}[\mathbf{x}]_{d_0}$ can be identified with the convex cone
$\{\mathbf{h}\in\mathbb{R}^{s(d_0)}: \mathbf{h}\leftrightarrow h\in{\rm Psd}_{d_0}(\mathbf{K})\}$ of $\mathbb{R}^{s(d_0)}$. So in the sequel, and unless if necessary, we will not distinguish between $f$ and $\mathbf{f}$.
Next, let $\mathbf{y}\in\mathbf{K}$ and $k\in\{1,2,\infty\}$ both fixed, and consider the following optimization problem $\mathcal{P}$
\begin{equation}
\label{inv-0}
\mathcal{P}:\quad\rho^k=
\displaystyle\min_{\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0}}\:\{\,\Vert \tilde{f}-f\Vert_k\::\:\mathbf{x}\mapsto \tilde{f}(\mathbf{x})-\tilde{f}(\mathbf{y})\,\in\,{\rm Psd}_{d_0}(\mathbf{K})\:\}.
\end{equation}
\begin{thm}
\label{th-ideal}
Let $\mathbf{K}\subset\mathbb{R}^n$ be with nonempty interior.
Problem (\ref{inv-0}) has an optimal solution $\tilde{f}^*\in\mathbb{R}[\mathbf{x}]_{d_0}$. In addition,
$\rho^k=0$ if and only if $\mathbf{y}$ is an optimal solution of $\mathbf{P}$.
\end{thm}
\begin{proof}
Obviously the constant polynomial $\mathbf{x}\mapsto \tilde{f}(\mathbf{x}):=1$ is a feasible solution with associated
value $\delta:=\Vert \tilde{f}-f\Vert_k$.
Moreover the optimal value of (\ref{inv-0}) is bounded below by $0$.
Observe that $\Vert \cdot\Vert_k$ defines a norm on $\mathbb{R}[\mathbf{x}]_{d_0}$.
Consider a minimizing sequence $(\tilde{f}^j)\subset\mathbb{R}[\mathbf{x}]_{d_0}$, $j\in\mathbb{N}$, hence such that
$\Vert \tilde{f}^j-f\Vert_k\to\rho^k$ as $j\to\infty$. As we have
$\Vert \tilde{f}^j-f\Vert_k\leq \delta$ for every $j$, the sequence $(\tilde{f}^j)$ belongs to the $\ell_k$-ball
$\mathbf{B}_k(f):=\{\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0}\,:\,\Vert \tilde{f}-f\Vert_k\leq\delta\}$, a compact set. Therefore, there is an element
$\tilde{f}^*\in\mathbf{B}_k(f)$ and a subsequence $(j_t)$, $t\in\mathbb{N}$, such that
$\tilde{f}^{j_t}\to\tilde{f}^*$ as $t\to\infty$. Let $\mathbf{x}\in\mathbf{K}$ be fixed arbitrary.
Obviously $(0\leq)\, \tilde{f}^{j_t}(\mathbf{x})-\tilde{f}^{j_t}(\mathbf{y})
\to\tilde{f}^*(\mathbf{x})-\tilde{f}^*(\mathbf{y})$ as $t\to\infty$, which implies $\tilde{f}^*(\mathbf{x})-\tilde{f}^*(\mathbf{y})\geq0$,
and so, as $\mathbf{x}\in\mathbf{K}$ was arbitrary, $\tilde{f}^*-\tilde{f}^*(\mathbf{y})\geq0$ on $\mathbf{K}$, i.e., $\tilde{f}^*-\tilde{f}^*(\mathbf{y})\in{\rm Psd}_{d_0}(\mathbf{K})$. Finally, we also obtain the desired result
\[\rho^k\,=\,\lim_{j\to\infty}\Vert \tilde{f}^{j}-f\Vert_k\,=\,\lim_{t\to\infty}\Vert \tilde{f}^{j_t}-f\Vert_k\,=\,\Vert \tilde{f}^*-f\Vert_k.\]
Next, if $\mathbf{y}$ is an optimal solution of $\mathbf{P}$ then $\tilde{f}:=f$ is an optimal solution of
$\mathcal{P}$ with value $\rho^k=0$. Conversely, if $\rho^k=0$ then $\tilde{f}^*=f$, and so by feasibility of $\tilde{f}^*\,(=f)$ for (\ref{inv-0}),
$f(\mathbf{x})\geq f(\mathbf{y})$ for all $\mathbf{x}\in\mathbf{K}$, which shows that $\mathbf{y}$ is an optimal solution of $\mathbf{P}$.
\end{proof}
Theorem \ref{th-ideal} states that the ideal inverse optimization problem is well-defined. However, even though
${\rm Psd}_{d_0}(\mathbf{K})$ is a finite dimensional convex cone, it has no simple and tractable characterization to be used for practical
computation. Therefore one needs an alternative and more tractable version of problem $\mathcal{P}$. Fortunately,
we next show that in the polynomial context such a formulation exists, thanks to the powerful Putinar's Positivstellensatz (Theorem \ref{th-put} above).
\section{Main result}
As the ideal inverse problem is intractable, we here provide tractable formulations
whose size depends on a parameter $d\in\mathbb{N}$. If the polynomial $\tilde{f}^*$ in Theorem \ref{th-ideal} belongs to $Q(g)$ then when $d$
increases the associated optimal value $\rho^k_d$ converges in finitely many steps to the optimal value $\rho^k$ of the ideal problem (\ref{inv-0}),
and $\tilde{f}^*$ can be obtained by solving finitely many semidefinite programs. And in fact this situation is {\it generic}.\\
With no loss of generality, i.e., up to some change of variable $\mathbf{x}'=\mathbf{x}-\mathbf{y}$
we may and will assume that $\mathbf{y}=0\in\mathbf{K}$.
\subsection{A practical inverse problem}
With $d\in\mathbb{N}$ fixed, consider the following optimization problem $\mathbf{P}_d$:
\begin{equation}
\label{inv-1}
\begin{array}{rl}
\mathbf{P}_d:\quad\rho_d^k:=\displaystyle\min_{\tilde{f},\sigma_j\in\mathbb{R}[\mathbf{x}] }&\Vert f-\tilde{f}\Vert_k\\
\mbox{s.t.}& \tilde{f}(\mathbf{x})-\tilde{f}(0)\,=\,\displaystyle\sum_{j=0}^{m} \sigma_j(\mathbf{x})\,g_j(\mathbf{x}),\quad\forall\mathbf{x}\in\mathbb{R}^n\\
&\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0};\:\sigma_j\in\Sigma[\mathbf{x}]_{d-v_j},\quad j=0,1,\ldots,m,
\end{array}
\end{equation}
where $d_0={\rm deg}\,f$, and $v_j=\lceil ({\rm deg}\,g_j)/2\rceil$, $j=1,\ldots,m$.
The parameter $d\in\mathbb{N}$ impacts (\ref{inv-1}) by the maximum degree allowed for
the s.o.s. weights $(\sigma_j)\subset\Sigma[\mathbf{x}]$ in Putinar's certificate for the polynomial
$\mathbf{x}\mapsto \tilde{f}(\mathbf{x})-\tilde{f}(0)$, and so the higher $d$ is, the lower $\rho_d^k$.
Next, observe that in (\ref{inv-1}), the constraint
\[\tilde{f}(\mathbf{x})-\tilde{f}(0)\,=\,\displaystyle\sum_{j=0}^{m} \sigma_j(\mathbf{x})\,g_j(\mathbf{x}),\quad\forall\mathbf{x}\in\mathbb{R}^n,\]
is equivalent to stating that
$\tilde{f}(\mathbf{x})-\tilde{f}(0)\in Q^{d_0}_d(g)$, with $Q^{d_0}_d(g)$ as in (\ref{puT-suff1}).
Therefore, in particular, $\tilde{f}(\mathbf{x})\geq\tilde{f}(0)$ for all $\mathbf{x}\in\mathbf{K}$, and so $0$ is a global minimizer of $\tilde{f}$ on $\mathbf{K}$.
So $\mathbf{P}_d$ is a strengthtening of $\mathcal{P}$ in that one has replaced the constraint
$\tilde{f}-\tilde{f}(0)\in {\rm Psd}_{d_0}(\mathbf{K})$ with the stronger condition $\tilde{f}-\tilde{f}(0)\in Q_d^{d_0}(g)$.
And so $\rho^k\leq\rho^k_d$ for all $d\in\mathbb{N}$. However, as we next see,
$\mathbf{P}_d$ is a tractable optimization problem with nice properties. Indeed,
$\mathbf{P}_d$ is a convex optimization problem and even a semidefinite program.
For instance, if $k=1$ one may rewrite $\mathbf{P}_d$ as:
\begin{equation}
\label{inv-3-k=1}
\begin{array}{rl}
\rho_d^1:=\displaystyle\min_{\lambda_\alpha\geq0,\tilde{f},\mathbf{Z}_j}&\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\lambda_\alpha\\
\mbox{s.t.}& \lambda_\alpha +\tilde{f}_\alpha\geq f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
& \lambda_\alpha -\tilde{f}_\alpha\geq -f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\langle\mathbf{Z}_0,\mathbf{B}_\alpha\rangle+\displaystyle\sum_{j=1}^{m} \langle\mathbf{Z}_j,\mathcal{C}^j_\alpha\rangle\,=\,\left\{\begin{array}{l}
\tilde{f}_\alpha,\:\mbox{ if }0<\vert\alpha\vert\leq\,d_0\\
0,\:\mbox{ if $\alpha=0$ or $\vert\alpha\vert>d_0$}\end{array}\right.\\
&\mathbf{Z}_j\succeq0,\quad j=0,1,\ldots,m.\end{array}
\end{equation}
with $\mathbf{B}_\alpha$ as in (\ref{balpha}) and $\mathcal{C}^j_\alpha$ as in (\ref{calpha}) (with $g_j$ in lieu of $g$).
If $k=\infty$ then one may rewrite $\mathbf{P}_d$ as:
\begin{equation}
\label{inv-3-k=infty}
\begin{array}{rl}
\rho_d^\infty:=\displaystyle\min_{\lambda\geq0,\tilde{f},\mathbf{Z}_j}&\lambda\\
\mbox{s.t.}& \lambda +\tilde{f}_\alpha\geq f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
& \lambda -\tilde{f}_\alpha\geq -f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\langle\mathbf{Z}_0,\mathbf{B}_\alpha\rangle+\displaystyle\sum_{j=1}^{m} \langle\mathbf{Z}_j,\mathcal{C}^j_\alpha\rangle\,=\,\left\{\begin{array}{l}
\tilde{f}_\alpha,\:\mbox{ if }0<\vert\alpha\vert\leq\,d_0\\
0,\:\mbox{ if $\alpha=0$ or $\vert\alpha\vert>d_0$}\end{array}\right.\\
&\mathbf{Z}_j\succeq0,\quad j=0,1,\ldots,m,\end{array}
\end{equation}
and finally, if $k=2$ then one may rewrite $\mathbf{P}_d$ as:
\begin{equation}
\label{inv-3-k=2}
\begin{array}{rl}
\rho_d^2:=\displaystyle\min_{\lambda\geq0,\tilde{f},\mathbf{Z}_j}&\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\lambda_\alpha\\
\mbox{s.t.}&\left[\begin{array}{cc}\lambda_\alpha &\tilde{f}_\alpha-f_\alpha\\
\tilde{f}_\alpha-f_\alpha&1\end{array}\right]\succeq0,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\langle\mathbf{Z}_0,\mathbf{B}_\alpha\rangle+\displaystyle\sum_{j=1}^{m} \langle\mathbf{Z}_j,\mathcal{C}^j_\alpha\rangle\,=\,\left\{\begin{array}{l}
\tilde{f}_\alpha,\:\mbox{ if }0<\vert\alpha\vert\leq\,d_0\\
0,\:\mbox{ if $\alpha=0$ or $\vert\alpha\vert>d_0$}\end{array}\right.\\
&\mathbf{Z}_j\succeq0,\quad j=0,1,\ldots,m.\end{array}
\end{equation}
\begin{rem}
\label{rem-f(0)}
{\rm Observe that in any feasible solution $(\tilde{f},\lambda,\mathbf{Z}_j)$ in all formulations
(\ref{inv-3-k=1})-(\ref{inv-3-k=2}),
$\tilde{f}_0$ plays no role in the constraints of (\ref{inv-1}), but since we minimize
$\Vert \tilde{f}-f\Vert_k$ then it is always optimal to set $\tilde{f}_0=f_0$.
That is, $\tilde{f}_d(0)=\tilde{f}_0=f_0=f(0)$.
}\end{rem}
\subsection*{Sparsity}
The semidefinite program (\ref{inv-1})-(\ref{inv-3-k=1}) has $m+1$ Linear Matrix Inequalities (LMI's)
$\mathbf{Z}_j\succeq0$ of size $O(n^d)$, which limits its application to
problems $\mathbf{P}$ of modest size. However large scale problems usually exhibit
sparsity patterns which sometimes can be exploited. For instance, in \cite{lassparse} we have provided
a specialized ``sparse" version of Theorem \ref{th-put}
for problems with structured sparsity
as described in Waki et al. \cite{waki}. Hence, with this specialized version
of Putinar's Positivstellensatz, one obtains
a {\it sparse} positivity certificate which when substituted in (\ref{inv-1}),
would permit to solve (\ref{inv-1}) for problems of much larger size. Typically, in \cite{waki}
the authors have applied the ``sparse semidefinite relaxations" to problem $\mathbf{P}$ with up
to 1000 variables! Moreover, the {\it running intersection property} that must satisfy the sparsity pattern
for convergence guarantee of such relaxations \cite{lassparse}, is {\it not} needed in the present context of inverse optimization. This is because one imposes $\tilde{f}$ to satisfy this specialized Putinar's Positivstellensatz.
\subsection{Duality}
The semidefinite program dual of (\ref{inv-3-k=1}) reads
\[\left\{\begin{array}{ll}
\displaystyle\max_{\mathbf{u},\mathbf{v}\geq0,\mathbf{z}}&\displaystyle
\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}f_\alpha(u_\alpha-v_\alpha)\:(=L_\mathbf{z}(f(0)-f))\\
\mbox{s.t.}&u_\alpha+v_\alpha\,\leq \,1,\quad\forall\,\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&u_\alpha-v_\alpha+z_\alpha \,=\,0,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\sum_{\alpha\in\mathbb{N}^n_d}z_\alpha\mathbf{B}_\alpha\succeq0\\
&\sum_{\alpha\in\mathbb{N}^n_d}z_\alpha\mathcal{C}^j_\alpha\succeq0,\quad j=1,\ldots,m,
\end{array}\right.\]
which, recalling the respective definitions (\ref{balpha}) and (\ref{calpha}) of the moment and localizing matrix, is the same as
\begin{equation}
\label{dual-k=1}
\left\{\begin{array}{ll}
\displaystyle\max_{\mathbf{u},\mathbf{v}\geq0,\mathbf{z}}&\displaystyle
\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}f_\alpha(u_\alpha-v_\alpha)\:(=L_\mathbf{z}(f(0)-f))\\
\mbox{s.t.}&u_\alpha+v_\alpha\,\leq \,1,\quad\forall\,\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&u_\alpha-v_\alpha+z_\alpha \,=\,0,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\mathbf{M}_d(\mathbf{z}),\,\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\,\succeq\,0,\quad j=1,\ldots,m.
\end{array}\right.
\end{equation}
Similarly, the semidefinite program dual of (\ref{inv-3-k=infty}) reads
\begin{equation}
\label{dual-k=infty}
\left\{\begin{array}{ll}
\displaystyle\max_{\mathbf{u},\mathbf{v}\geq0,\mathbf{z}}&\displaystyle
\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}f_\alpha(u_\alpha-v_\alpha)\:(=L_\mathbf{z}(f(0)-f))\\
\mbox{s.t.}&\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}u_\alpha+v_\alpha\,\leq \,1\\
&u_\alpha-v_\alpha+z_\alpha\,=\,0,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\mathbf{M}_d(\mathbf{z}),\,\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\,\succeq\,0,\quad j=1,\ldots,m,
\end{array}\right.
\end{equation}
and the semidefinite program dual of (\ref{inv-3-k=2}) reads
\begin{equation}
\label{dual-k=2}
\left\{\begin{array}{ll}
\displaystyle\max_{\mathbf{z},\Delta_\alpha}&\displaystyle
\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\left\langle \Delta_\alpha,\left[\begin{array}{cc}0 &f_\alpha\\mathbf{f}_\alpha&-1\end{array}\right]\right\rangle\\
\mbox{s.t.}&\displaystyle
\left\langle \Delta_\alpha,\left[\begin{array}{cc}1 &0\\0&0\end{array}\right]\right\rangle
\leq \,1,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\\
&\left\langle \Delta_\alpha,\left[\begin{array}{cc}0 &1\\1&0\end{array}\right]\right\rangle
+z_\alpha \,=\,0,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\mathbf{M}_d(\mathbf{z}),\,\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\,\succeq\,0,\quad j=1,\ldots,m\\
&\Delta_\alpha\succeq0,\quad \forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}.\\
\end{array}\right.
\end{equation}
One may show that one may replace the criterion in (\ref{dual-k=2})
with the equivalent concave criterion
\[\max_{\mathbf{z}}\:\left\{L_\mathbf{z}(f(0)-f))-\frac{1}{4}\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}z_\alpha^2\:\right\}.\]
\begin{lem}
\label{nodualgap}
Assume that $\mathbf{K}\subset\mathbb{R}^n$ has nonempty interior. Then there is no duality gap between the semidefinite programs (\ref{inv-3-k=1}) and (\ref{dual-k=1}), (\ref{inv-3-k=infty}) and (\ref{dual-k=infty}),
and (\ref{inv-3-k=2}) and (\ref{dual-k=2}). Moreover, all semidefinite programs
(\ref{inv-3-k=1}), (\ref{inv-3-k=infty}) and (\ref{inv-3-k=2}) have an optimal solution $\tilde{f}_d\in\mathbb{R}[\mathbf{x}]_{d_0}$.
\end{lem}
\begin{proof}
The proof is detailed for the case $k=1$ and omitted for the cases $k=2$ and $k=\infty$ because
it is very similar.
Observe that $\rho_d^1\geq0$ and the constant polynomial $\tilde{f}(\mathbf{x})=0$ for all $\mathbf{x}\in\mathbb{R}^n$, is an obviously feasible solution of
(\ref{inv-1}) (hence of (\ref{inv-3-k=1})).
Therefore $\rho_d^1$ being finite, it suffices to prove that Slater's condition\footnote{Slater's condition holds if there exists a strictly feasible solution, and so
for the dual (\ref{dual-k=1}), if there exists $\mathbf{z}$ such that $\mathbf{M}_d(\mathbf{z}),\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\succ0$, $j=1,\ldots,m$, and
$u_\alpha+v_\alpha<1,\,\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}$. Then from a standard result in convex optimization, there is no duality gap between
(\ref{inv-3-k=1}) and (\ref{dual-k=1}), and if the values are bounded then (\ref{inv-3-k=1}) has an optimal solution.}
holds for the dual (\ref{dual-k=1}). Then the conclusion of Lemma
\ref{nodualgap} follows from a standard result of convex optimization.
Let $\mu$ be the finite Borel measure defined by
\[\mu(B)\,:=\, \int_{B\cap\mathbf{K}}{\rm e}^{-\Vert\mathbf{x}\Vert^2}d\mathbf{x},\qquad \forall B\in\mathcal{B}\]
(where $\mathcal{B}$ is the usual Borel $\sigma$-field),
and let $\mathbf{z}=(z_\alpha)$, $\alpha\in\mathbb{N}^n_{2d}$, with
\[z_\alpha\,:=\,\kappa\,\int_\mathbf{K} \mathbf{x}^\alpha \,d\mu(\mathbf{x}),\qquad \alpha\in\mathbb{N}^n_{2d},\]
for some $\kappa>0$ sufficiently small to ensure that
\begin{equation}
\label{cond-1}
\kappa \,\displaystyle\left\vert \int \mathbf{x}^\alpha\,d\mu(\mathbf{x})\right\vert \,<\,1,\qquad\forall \alpha\in\mathbb{N}^n_{2d}\setminus\{0\}.
\end{equation}
Define $u_\alpha=\max[0,-z_\alpha]$ and $v_\alpha=\max[0,z_\alpha]$,
$\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}$, so that
$u_\alpha+v_\alpha< 1$, $\alpha\in\mathbb{N}^n_{2d}\setminus\{0\}$.
Hence $(u_\alpha,v_\alpha,\mathbf{z})$ is a feasible solution
of (\ref{dual-k=1}). In addition, $\mathbf{M}_d(\mathbf{z})\succ0$ and $\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\succ0$, $j=1,\ldots,m$,
because $\mathbf{K}$ has nonempty interior, and so Slater's condition
holds for (\ref{dual-k=1}), the desired result.
If $k=\infty$ one chooses $\mathbf{z}$ such that
\[\kappa \,\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\left\vert \int \mathbf{x}^\alpha\,d\mu(\mathbf{x})\right\vert \,<\,1,\]
and if $k=2$ then one chooses $\mathbf{z}$ as in (\ref{cond-1})
and $\Delta_\alpha:=\left[\begin{array}{cc}1/2&\kappa_\alpha\\\kappa_\alpha &1\end{array}\right]\succ0$, for all
$\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}$, such that
\[2\kappa_\alpha:=\kappa \,\displaystyle\int \mathbf{x}^\alpha\,d\mu(\mathbf{x}),\qquad\forall \alpha\in\mathbb{N}^n_{2d}\setminus\{0\}.\]
\end{proof}
\begin{thm}
\label{th1}
Assume that $\mathbf{K}$ in (\ref{setk}) has nonempty interior, and
let $\mathbf{x}^*\in\mathbf{K}$ be a global minimizer of $\mathbf{P}$ with optimal value $f^*$,
and let $\tilde{f}_d\in\mathbb{R}[\mathbf{x}]_{d_0}$ be an optimal solution of $\mathbf{P}_d$ in (\ref{inv-1}) with optimal value $\rho_d^k$. Then:
{\rm (a)} $0\in\mathbf{K}$ is a global minimizer of the problem $\tilde{f}^*_d=\min_\mathbf{x}\{\tilde{f}_d(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$. In particular,
if $\rho_d^k=0$ then $\tilde{f}_d=f$ and $0$ is a global minimizer of $\mathbf{P}$.\\
{\rm (b)} If $k=1$ then $f^*\,\leq\,f(0)\,\leq\,f^*+\rho_d^1\,\displaystyle\sup_{\alpha\in\mathbb{N}^n_{d_0}} \vert (\mathbf{x}^*)^\alpha\vert$. In particular, if $\mathbf{K}\subseteq [-1,1]^n$ then
$f^*\,\leq\,f(0)\,\leq\,f^*+\rho_d^1$.\\
{\rm (c)} If $k=\infty$ then $f^*\,\leq\,f(0)\,\leq\,f^*+\rho_d^\infty\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}} \vert (\mathbf{x}^*)^\alpha\vert$. In particular if $\mathbf{K}\subseteq [-1,1]^n$ then
$f^*\,\leq\,f(0)\,\leq\,f^*+s(d_0)\,\rho_d^\infty$.
\end{thm}
\begin{proof}
(a) Existence of $\tilde{f}_d$ is guaranteed by Lemma \ref{nodualgap}. From the constraints of (\ref{inv-1}) we have:
$\tilde{f}_d(\mathbf{x})-\tilde{f}(0)=\sum_{j=0}^{m}\sigma_j(\mathbf{x})\,g_j(\mathbf{x})$
which implies that
$\tilde{f}_d(\mathbf{x})\geq \tilde{f}(0)$ for all $\mathbf{x}\in\mathbf{K}$,
and so $0$ is a global minimizer of the optimization problem $\mathbf{P}':\:\min_\mathbf{x}\{\tilde{f}_d(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$.
(b) Let $\mathbf{x}^*\in\mathbf{K}$ be a global minimizer of $\mathbf{P}$. Observe that with $k=1$,
\begin{eqnarray}
\nonumber
f^*=f(\mathbf{x}^*)&=&\underbrace{f(\mathbf{x}^*)-\tilde{f}_d(\mathbf{x}^*)}+\underbrace{\tilde{f}_d(\mathbf{x}^*)-\tilde{f}_d(0)}_{\geq0}+\tilde{f}_d(0)\\
\nonumber
&\geq&\tilde{f}_d(0)-\vert \tilde{f}_d(\mathbf{x}^*)-f(\mathbf{x}^*)\vert\\
\nonumber
&\geq&\tilde{f}_d(0)-\Vert \tilde{f}_d-f\Vert_1\,\times\sup_{\alpha\in\mathbb{N}^n_{d_0}}\vert(\mathbf{x}^*)^\alpha\vert\\
\label{a}
&\geq&f(0)-\rho^1_d\,\sup_{\alpha\in\mathbb{N}^n_{d_0}}\vert(\mathbf{x}^*)^\alpha\vert
\end{eqnarray}
since $\tilde{f}_d(0)=f(0)$; see Remark \ref{rem-f(0)}.
(c) The proof is similar to that of (b) using that with $k=\infty$,
\[\vert\tilde{f}_d(\mathbf{x})-f(\mathbf{x})\vert \,\geq\,\left(\sup_{\alpha\in\mathbb{N}^n_{d_0}}\vert \tilde{f}_{d\alpha}-f_\alpha\vert\right)\times\sum_{\alpha\in\mathbb{N}^n_{d_0}}\vert\mathbf{x}^\alpha\vert.\]
\end{proof}
So not only Theorem \ref{th1} states that $0$ is the global optimum of the optimization problem
$\min\{\tilde{f}_d(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$, but it also states that the optimal value $\rho^k_d$
also measures how far is $f(0)$ from the optimal value $f^*$ of the initial problem $\mathbf{P}$.
Moreover, observe that Theorem \ref{th1} merely requires existence of a minimizer
and nonemptyness of $\mathbf{K}$.
In particular, $\mathbf{K}$ may not be compact.
\subsection{A canonical form for the $\ell_1$-norm}
When $\mathbf{K}$ is compact and if one uses the $\ell_1$-norm then the
optimal solution $\tilde{f}_d\in\mathbb{R}[\mathbf{x}]_{d_0}$ in Theorem \ref{th1} (with $k=1$) takes a particularly
simple {\it canonical} form:
As $\mathbf{K}$ is compact we may and will assume (possibly after some scaling) that $\mathbf{K}\subseteq [-1,1]^n$
and so in the definition of (\ref{setk}) we may and will add the $n$ redundant quadratic constraints $g_{m+i}(\mathbf{x})\geq0$, $i=1,\ldots,n$,
with $\mathbf{x}\mapsto g_{m+i}(\mathbf{x})=1-x_i^2$ for every $i$, that is,
\begin{equation}
\label{setk-1}
\mathbf{K}=\{\mathbf{x}\in\mathbb{R}^n\::\:g_j(\mathbf{x})\geq0,\quad j=1,\ldots,m+n\},\end{equation}
and
\[Q_d(g)\,=\,\left\{\sum_{j=0}^{n+m}\sigma_j\,g_j\::\quad\sigma_j\in\Sigma[\mathbf{x}]_{d-v_j},\:j=0,\ldots,m+n\right\},\]
which is obviously Archimedean.
\begin{thm}
\label{th1-ell1}
Assume that $\mathbf{K}$ in (\ref{setk-1}) has a nonempty interior and
let $\tilde{f}_d\in\mathbb{R}[\mathbf{x}]_{d_0}$ be an optimal solution of $\mathbf{P}_d$ in (\ref{inv-1}) (with $m+n$ instead of $m$) with optimal value $\rho_d^1$
for the $\ell_1$-norm. Then:
(i) $\tilde{f}_d$ is of the form
\begin{equation}
\label{th1-ell1-1}
\tilde{f}_d(\mathbf{x})\,=\,\left\{\begin{array}{ll}f(\mathbf{x})+\mathbf{b}'\mathbf{x}&\mbox{if $d_0=1$}\\
f(\mathbf{x})+\mathbf{b}'\mathbf{x}+\sum_{i=1}^n\lambda_i^*\,x_i^2&\mbox{if $d_0>1$,}\end{array}\right.
\end{equation}
for some vector $\mathbf{b}\in\mathbb{R}^n$ and some
nonnegative vector $\lambda^*\in\mathbb{R}^n$, optimal solution of the semidefinite program:
\begin{equation}
\label{inv-3-ell1}
\begin{array}{rl}
\rho_d^1:=\displaystyle\min_{\lambda,\mathbf{b}}&\displaystyle
\Vert\mathbf{b}\Vert_1+\sum_{i=1}^n\lambda_i\\
\mbox{s.t.}& f-f(0)+\mathbf{b}'\mathbf{x}+\displaystyle\sum_{i=1}^n\lambda_i\,x_i^2\,\in Q_{d}(g),\quad
\lambda\geq0.
\end{array}
\end{equation}
(ii) The vector $\mathbf{b}$ is of the form $-\nabla f(0)+\sum_{j\in J(0)}\theta_j\,\nabla g_j(0)$ for some nonnegative scalars $(\theta_j)$
(where $j\in J(0)$ if and only if $g_j(0)=0$).
\end{thm}
\begin{proof}
(i) Notice that the dual (\ref{dual-k=1}) of (\ref{inv-3-k=1}) is equivalent to:
\begin{equation}
\label{equiv-1}
\left\{\begin{array}{ll}\displaystyle\max_{\mathbf{z}}&L_\mathbf{z}(f(0)-f))\\
\mbox{s.t.}&\mathbf{M}_d(\mathbf{z}),\,\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\,\succeq\,0,\quad j=1,\ldots,m+n,\\
&\vert z_\alpha\vert \leq 1,\quad\forall\,\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}.
\end{array}\right.\end{equation}
Next, since $\mathbf{M}_{d}(\mathbf{z})\succeq0$ one may invoke
same arguments as those used in
Lasserre and Netzer \cite[Lemma 4.1, 4.2]{lass-netzer}, to
obtain that for very $\alpha\in\mathbb{N}^n_{2d}$ with $\vert\alpha\vert>1$,
\[\mathbf{M}_d(\mathbf{z})\succeq0\,\Rightarrow\quad\vert z_\alpha\vert\leq \max_{i=1,\ldots,n}\{\max[L_\mathbf{z}(x_i^2),L_\mathbf{z}(x_i^{2d})]\}.\]
Moreover the constraint $\mathbf{M}_{d-1}(g_{m+i}\,\mathbf{z})\succeq0$ implies
$\mathbf{M}_{d-1}(g_{m+i}\,\mathbf{z})(\ell,\ell)\geq0$ for all $\ell$, and so in particular, one obtains
$L_\mathbf{z}(x_i^{2k-2})\geq L_\mathbf{z}(x_i^{2k})$ for all $k=1,\ldots,d$ and all $i=1,\ldots,n$.
Hence $\vert z_\alpha\vert\leq \max_{i=1,\ldots,n}L_\mathbf{z}(x_i^2)$ for every $\alpha\in\mathbb{N}^n_{2d}$ with $\vert\alpha\vert>1$.
Therefore in (\ref{equiv-1}) one may replace the constraint $\vert z_\alpha\vert\leq 1$ for all $\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}$ with
the $2n$ inequality constraints $\pm L_\mathbf{z}(x_i)\leq1$, $i=1,\ldots,n$, if $d_0=1$ and
the $3n$ inequality constraints:
\begin{equation}
\label{simplify}
\pm L_\mathbf{z}(x_i)\leq 1,\:L_\mathbf{z}(x_i^2)\leq1,\:i=1,\ldots,n\end{equation}
if $d_0>1$.
Consequently, (\ref{equiv-1}) is the same as the semidefinite program
\begin{equation}
\label{equiv-1111}
\left\{\begin{array}{ll}\displaystyle\max_{\mathbf{z}}&L_\mathbf{z}(f(0)-f))\\
\mbox{s.t.}&\mathbf{M}_d(\mathbf{z}),\,\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\,\succeq\,0,\quad j=1,\ldots,m+n,\\
&\pm L_\mathbf{z}(x_i)\leq 1,\:L_\mathbf{z}(x_i^2)\leq1,\:i=1,\ldots,n.
\end{array}\right.\end{equation}
Let $\mathbf{b}^1=(b^1_i)$ (resp. $\mathbf{b}^2=(b^2_i)$) be the nonnegative vector of dual variables associated with the constraints
$L_\mathbf{z}(x_i)\leq 1$ (resp. $-L_\mathbf{z}(x_i)\leq 1$), $i=1,\ldots,n$. Similarly, let $\lambda_i$ be the dual variable associated with the constraint $L_z(x_i^2)\leq 1$. Then the dual of (\ref{equiv-1111}) is the semidefinite program:
\[\left\{\begin{array}{ll}\displaystyle\max_{\mathbf{b}^1,\mathbf{b}^2,\lambda}&\displaystyle\sum_{i=1}^n\left((b_i^1+b_i^2)+\lambda_i\right)\\
\mbox{s.t.}&f-f(0)+(\mathbf{b}^1-\mathbf{b}^2)'\mathbf{x}+\displaystyle\sum_{i=1}^n\lambda_i\, x_i^2\,\in\,Q_{d}(g)\\
&\mathbf{b}^1,\mathbf{b}^2,\lambda\geq0
\end{array}\right.\]
which is equivalent to (\ref{inv-3-ell1}).
(ii) Let $(\mathbf{b},\lambda)$ be an optimal solution of (\ref{inv-3-ell1}), so that
\[f-f(0)+\mathbf{b}'\mathbf{x}+\displaystyle\sum_{i=1}^n\lambda_i\, x_i^2=\sigma_0+\sum_{j=1}^{m+n}\sigma_j\,g_j,\]
for some SOS polynomials $\sigma_j$. Evaluating at $\mathbf{x}=0$ yields
\[\sigma_0(0)=0;\quad \underbrace{\sigma_j(0)}_{\theta_j\geq0}\,g_j(0)=0,\quad j=1,\ldots, m+n.\]
Differentiating and evaluating at $\mathbf{x}=0$ and using that $\sigma_j$ is SOS and $\sigma_j(0)g_j(0)=0$, $j=1,\ldots,n+m$, yields:
\[\nabla f(0)+\mathbf{b}\,=\,\sum_{j=1}^{n+m}\sigma_j(0)\,\nabla g_j(0)\,=\,\sum_{j\in J(0)}\theta_j\,\nabla g_j(0),\]
which is the desired result.
\end{proof}
From the proof of Theorem \ref{th1-ell1}, this special form of $\tilde{f}_d$ is specific to the $\ell_1$-norm,
which yields the constraint $\vert z_\alpha\vert\leq1,\:\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}$ in the dual (\ref{dual-k=1}) and allows its simplification
(\ref{simplify}) thanks to a property of the moment matrix described in \cite{lass-netzer}. Observe that the canonical form (\ref{th1-ell1-1}) of $\tilde{f}_d$ is {\it sparse}
since $\tilde{f}_d$ differs from $f$ in at most $2n$ entries only (recall that $f$ has ${n+d_0\choose n}$ entries). This is another example of sparsity properties
of optimal solutions of $\ell_1$-norm minimization problems, already observed in other contexts (e.g., in some compressed sensing applications). Moreover, it has the following consequence for nonlinear $0/1$ programs.
\begin{cor}
\label{0/1}
Let $\mathbf{K}=\{0,1\}^n$, $f\in\mathbb{R}[\mathbf{x}]_d$ and let $\mathbf{y}\in\mathbf{K}$ with
$I_1:=\{i:y_i=0\}$ and $I_2:=\{i:y_i=1\}$.
Then an optimal solution $\tilde{f}_d\in\mathbb{R}[\mathbf{x}]_{d_0}$ of the inverse problem (\ref{inv-1}) for
the $\ell_1$-norm, is of the form
\begin{equation}
\label{form-01}
\tilde{f}_d(\mathbf{x})\,=\,f(\mathbf{x})+\mathbf{b}^T(\mathbf{x}-\mathbf{y}),\end{equation}
for some coefficient vector $\mathbf{b}\in\mathbb{R}^n$ such that $b_i\geq0$ if $i\in I_1$
and $b_i\leq0$ if $i\in I_2$, $i=1,\ldots,n$. Moreover, $\mathbf{b}$ is an optimal solution of the semidefinite program:
\[\begin{array}{ll}\displaystyle\min_{\mathbf{b},\lambda,\gamma}&\displaystyle\sum_{i\in I_1}b_i-\sum_{i\in I_2}b_i\:(=\vert\mathbf{b}\vert)\\
\mbox{s.t.}&f(\mathbf{x})-f(\mathbf{y})+\mathbf{b}^T (\mathbf{x}-\mathbf{y})=\sigma_0+\displaystyle\sum_{i=1}^n(x_i^2-x_i)\,\sigma_i\\
&b_i\geq0,\:i\in I_1;\:b_i\leq0,\:i\in I_2;\quad\sigma_0\in\Sigma[\mathbf{x}]_d;\:\sigma_i\in\mathbb{R}[\mathbf{x}]_{d-1},\,i=1,\ldots,n.
\end{array}\]
\end{cor}
\begin{proof}
We briefly sketch the proof which
is very similar to that of Theorem \ref{th1-ell1} even though $\mathbf{K}$ does not have a nonempty interior
and $d_0$ is not required to be even. Recall that $\mathbf{y}\in\mathbf{K}$ could be assumed to be $0$ and so if $\mathbf{K}=\{0,1\}^n$
then the new feasible set after the change of variable $\mathbf{u}:=\mathbf{x}-\mathbf{y}$ is now
$\tilde{\mathbf{K}}=\prod_{i\in I_1}(\{0,1\})\prod_{i\in I_2}(\{-1,0\})$. Similarly, let $f'\in\mathbb{R}[\mathbf{x}]_d$ be the polynomial $f$ in the new coordinates $\mathbf{u}$,
i.e., $f'(\mathbf{u})=f(\mathbf{u}+\mathbf{y})$.
So for every $\alpha\in\mathbb{N}^n$, define $\bar{\alpha}\in\{0,1\}^n$ and $\Delta_{\alpha}\subset I_2$ by:
\begin{eqnarray*}
\bar{\alpha}_i&:=&1\mbox{ if $\alpha_i>0$ and $0$ otherwise, $i=1,\ldots,n$}\\
\Delta_{\alpha}&:=&\{i\in I_2:0<\alpha_i\mbox{ is even}\},
\end{eqnarray*}
and let $\vert\Delta_{\alpha}\vert$ denotes the cardinality of $\Delta_{\alpha}$.
Then because of the boolean constraints $u_i^2=u_i$, $i\in I_1$ and $u_i^2=-u_i$, $i\in I_2$,
in the definition of $\tilde{\mathbf{K}}$,
(\ref{equiv-1}) reads
\[\left\{\begin{array}{ll}\displaystyle\max_{\mathbf{z}}&L_\mathbf{z}(f'(0)-f')\\
\mbox{s.t.}&\mathbf{M}_d(\mathbf{z})\succeq\,0;\quad z_{\alpha}=(-1)^{\vert\Delta_{\alpha}\vert}\,z_{\bar{\alpha}},\quad\alpha\in\mathbb{N}^n_{2d}\\
&\vert z_{\alpha}\vert \leq 1,\quad\forall\,\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}.
\end{array}\right.\]
But this combined with $\mathbf{M}_d(\mathbf{z})\succeq0$ implies that $\mathbf{M}_d(\mathbf{z})$
can be simplified to a smaller real symmetric matrix $\overline{\mathbf{M}}_d(\mathbf{z})$ with rows and columns indexed by only the square-free
monomials $\mathbf{x}^{\bar{\alpha}}$, $\alpha\in\mathbb{N}^n_{2d}$.
Indeed, every column $\alpha$ of $\mathbf{M}_d(\mathbf{z})$ is exactly identical to $\pm$ the column of $\mathbf{M}_d(\mathbf{z})$ indexed by $\bar{\alpha}$.
Also, invoking \cite{lass-netzer} and with similar arguments,
$\vert z_{\alpha}\vert\leq \max\left[\max_{i\in I_1}L_\mathbf{z}(u_i),\,\max_{i\in I_2}-L_\mathbf{z}(u_i)\right]$
for all $\alpha$, and so (\ref{equiv-111}) is equivalent to:
\begin{equation}
\label{equiv-111}
\left\{\begin{array}{ll}\displaystyle\max_{\mathbf{z}}&L_\mathbf{z}(f'(0)-f')\\
\mbox{s.t.}&\overline{\mathbf{M}}_d(\mathbf{z})\succeq\,0;\quad z_{\alpha}=(-1)^{\vert\Delta_{\alpha}\vert}\,z_{\bar{\alpha}},\quad\alpha\in\mathbb{N}^n_{2d}\\
&L_\mathbf{z}(\mathbf{x}_i)\,\leq\,1,\qquad i\in I_1\\
&-L_\mathbf{z}(\mathbf{x}_i)\,\leq\,1,\qquad i\in I_2.
\end{array}\right.\end{equation}
Finally, let $\mu$ be a Borel measure with support exactly $\tilde{\mathbf{K}}$, and scaled to satisfy
$\vert\int u_id\mu\vert <1$, $i=1,\ldots, n$. Its associated vector of moment $\mathbf{z}=(\int \mathbf{u}^{\alpha} d\mu)$, $\alpha\in\mathbb{N}^n$, is feasible in
(\ref{equiv-111}) and $\overline{\mathbf{M}}_d(\mathbf{z})\succ0$. Hence Slater's condition holds for (\ref{equiv-111}), which in turn
implies that there is no duality gap with its dual which reads:
\[\begin{array}{ll}\displaystyle\min_{\mathbf{b},\lambda,\gamma}&\displaystyle\sum_{i=1}^nb_i\\
\mbox{s.t.}&f'(\mathbf{u})-f'(0)+\displaystyle\sum_{i\in I_1}b_iu_i-\displaystyle\sum_{i\in I_2}b_iu_i=\sigma_0(\mathbf{u})\\
&+\displaystyle\sum_{i\in I_1}^n\sigma_i(\mathbf{u})\,(u_i^2-u_i)
+\displaystyle\sum_{i\in I_2}^n\sigma_i(\mathbf{u})\,(u_i^2+u_i)\\
&\mathbf{b}\geq0;\quad\sigma_0\in\Sigma[\mathbf{u}]_d;\:\sigma_i\in\mathbb{R}[\mathbf{u}]_{d-1},\,i=1,\ldots,n.
\end{array}\]
Moreover, the dual has an optimal solution because the optimal value
is bounded below by zero. Recalling that $\mathbf{u}=\mathbf{x}-\mathbf{y}$ and $f'(\mathbf{u})=f(\mathbf{x})$, we retrieve the semidefinite program of the Corollary
and so $\tilde{f}_d$ is indeed of the form (\ref{form-01}).
\end{proof}
\subsection{Structural constraints}
\label{structural}
It may happen that the initial criterion $f\in\mathbb{R}[\mathbf{x}]$ has some structure that one wishes to keep
in the inverse problem. For instance, in MAXCUT problems on $\mathbf{K}=\{-1,1\}^n$, $f$ is a quadratic form
$\mathbf{x}\mapsto \mathbf{x}' \mathbf{A}\mathbf{x}$ for some real symmetric matrix $\mathbf{A}$ associated with a graph $(V,E)$, where $\mathbf{A}_{ij}\neq0$ if
and only if $(i,j)\in E$. Therefore, in the inverse optimization problem, one may wish that $\tilde{f}$ in (\ref{inv-1})
is also a quadratic form associated with the same graph $(V,E)$, so that $\tilde{f}(\mathbf{x})=\mathbf{x}'\tilde{\mathbf{A}}\mathbf{x}$ with
$\tilde{\mathbf{A}}_{ij}=0$ for all $(i,j)\not\in E$.
So if $\Delta_f\subset\mathbb{N}^n_{d_0}$ denotes the subset of (structural) multi-indices for which $f$ and $\tilde{f}$ should have same coefficient,
then in (\ref{inv-1}) one includes the additional constraint $\tilde{f}_\alpha=f_\alpha$ for all $\alpha\in\Delta_f$. Notice that $0\in\Delta_f$ because $\tilde{f}_0=f_0$; see Remark \ref{rem-f(0)}.
For instance, with $\mathbf{K}$ as in (\ref{setk-1}) and $k=1$, (\ref{inv-3-k=1}) reads
\begin{equation}
\label{inv-33-k=1}
\begin{array}{rl}
\rho_d^1:=\displaystyle\min_{\tilde{f},\lambda_\alpha,\mathbf{Z}_j}&\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\lambda_\alpha\\
\mbox{s.t.}& \lambda_\alpha +\tilde{f}_\alpha\geq f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\Delta_f\\
& \lambda_\alpha -\tilde{f}_\alpha\geq -f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\Delta_f\\
&\langle\mathbf{Z}_0,\mathbf{B}_\alpha\rangle+\displaystyle\sum_{j=1}^{m+n} \langle\mathbf{Z}_j,\mathcal{C}^j_\alpha\rangle\,=\,\left\{\begin{array}{l}
f_\alpha,\:\mbox{ if }\alpha\in\Delta_f\setminus\{0\}\\
\tilde{f}_\alpha,\:\mbox{ if }\alpha\in\mathbb{N}^n_{d_0}\setminus \Delta_f\\
0,\:\mbox{ if $\alpha=0$ or $\vert\alpha\vert>d_0$}\end{array}\right.\\
&\mathbf{Z}_j\succeq0,\quad j=0,1,\ldots,m+n,\end{array}
\end{equation}
and its dual has the equivalent form,
\begin{equation}
\label{equiv-11}
\left\{\begin{array}{ll}\displaystyle\max_{\mathbf{z}}&L_\mathbf{z}(f(0)-f))\\
\mbox{s.t.}&\mathbf{M}_d(\mathbf{z}),\,\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\,\succeq\,0,\quad j=1,\ldots,m+n,\\
&\vert z_\alpha\vert \leq 1,\quad\forall\,\alpha\in\mathbb{N}^n_{d_0}\setminus(\Delta_f\cup\{0\}).
\end{array}\right.\end{equation}
However now (\ref{inv-33-k=1}) may not have a feasible solution. In problems where $d_0$ is even and $\Delta_f$ does not contain
the monomials $\alpha\in\mathbb{N}^n_{d_0}$ such that $\mathbf{x}^\alpha=x_i^2$, or $\mathbf{x}^\alpha=x_i^{d_0}$, $i=1,\ldots,n$, and if (\ref{inv-33-k=1}) has a feasible solution, then there is an optimal solution $\tilde{f}$ with still the special form described in Theorem \ref{th1-ell1}, but with $b_k=0$ if $\alpha=e_k\in\Delta_f$ (where all entries of
$e_k$ vanish except the one at position $k$).
\subsection{Asymptotics when $d\to\infty$}
We now relate $\mathbf{P}_d$, $d\in\mathbb{N}$, with the ideal inverse problem $\mathcal{P}$ in (\ref{inv-0}) when $d$ increases.
\begin{prop}
\label{prop-asymptotics}
Let $\mathbf{K}$ be as in (\ref{setk}) with nonempty interior.
For every $k=1,2,\infty$, let
$\tilde{f}_d\in\mathbb{R}[\mathbf{x}]_{d_0}$ (resp. $\tilde{f}^*\in\mathbb{R}[\mathbf{x}]_{d_0}$) be an optimal solution of (\ref{inv-1}) (resp. (\ref{inv-0}))
with associated optimal value $\rho^k_d$ (resp. $\rho^k$).
The sequence $(\rho^k_d)$, $d\in\mathbb{N}$, is monotone nonincreasing and converges to $\hat{\rho}^k\geq\rho^k$.
Moreover, every accumulation point $\hat{f}\in\mathbb{R}[\mathbf{x}]_{d_0}$ of the sequence $(\tilde{f}_d)$, $d\in\mathbb{N}$,
is such that $\hat{f}-\hat{f}(0)\in{\rm Psd}_{d_0}(\mathbf{K})$ and
$\Vert \hat{f}-f\Vert_k=\hat{\rho}^k$. Finally, if
$\tilde{f}^*-\tilde{f}^*(0)$ is in $Q(g)$, then
$\rho^k_d=\hat{\rho}^k=\rho^k$ for some $d$.
\end{prop}
\begin{proof}
Observe that the sequence $(\tilde{f}_d)$, $d\in\mathbb{N}$, is contained in the ball
$\{h\,:\,\Vert h-f\Vert_k\leq \rho^k_{d_0}\}\subset\mathbb{R}[\mathbf{x}]_{d_0}$,
for some $d_0\in\mathbb{N}$. So let $\hat{f}$ be an accumulation point of $(\tilde{f}_d)$.
Since $\tilde{f}_d-\tilde{f}(0)\geq0$ on $\mathbf{K}$ for all $d$, a simple continuity argument
yields $\hat{f}-\hat{f}(0)\geq0$ on $\mathbf{K}$, i.e., $\hat{f}-\hat{f}(0)\in{\rm Psd}_{d_0}(\mathbf{K})$. Moreover, the sequence $(\rho^k_d)$ is obviously monotone nonincreasing
and bounded below by zero. Hence $\lim_{d\to\infty}\rho^k_d=:\hat{\rho}^k\geq\rho^k$,
and by continuity $\Vert\hat{f}-f\Vert_k=\hat{\rho}^k$.
Finally, if $\tilde{f}^*-\tilde{f}(0)\in Q(g)$ then $\tilde{f}^*-\tilde{f}(0)\in Q^{d_0}_{d}(g)$ for some $d$, and so
$\tilde{f^*}$ is a feasible solution of (\ref{inv-1}) but with value $\rho^k\leq\rho^k_d$.
Therefore, we conclude that $\tilde{f}^*$ is an optimal solution of (\ref{inv-1}).
\end{proof}
Proposition \ref{prop-asymptotics} relates $\rho^k_d$ and $\rho^k$ in a strong sense when
$\tilde{f}^*-\tilde{f}(0)\in Q(g)$. However, we would like to know how restrictive is the constraint
$\tilde{f}^*-\tilde{f}(0)\in Q(g)$ compared to $\tilde{f}^*-\tilde{f}(0)\in {\rm Psd}_{d_0}(\mathbf{K})$. Indeed,
even though ${\rm Psd}_{d_0}(\mathbf{K})={\rm cl}\,(\cup_{\ell=0}^\infty)Q^{d_0}_\ell$ when $\mathbf{K}$ satisfies the
assumptions of Theorem \ref{limit}, in general
an approximating sequence $(f_\ell)\subset Q(g)$, $\ell\in\mathbb{N}$ (with $\Vert f_\ell-\tilde{f}^*\Vert_k\to 0$), does
not have the property that
$f_\ell(\mathbf{x})-f_\ell(0)\geq0$ for all $\mathbf{x}$ on $\mathbf{K}$.
\subsection{$Q(g)$ versus ${\rm Psd}_{d_0}(\mathbf{K})$}
Therefore the question is: {\em How often a polynomial
nonnegative on $\mathbf{K}$ (and with at least one zero in $\mathbf{K}$) is an element of $Q(g)$?}
This question can be answered in a number of cases which suggest that
$f\geq0$ on $\mathbf{K}$ and $f\not\in Q(g)$ can be true in very specific cases only (at least when $\mathbf{K}$ is compact
and $Q(g)$ is Archimedean).
Indeed $f\geq0$ on $\mathbf{K}$ implies $f\in Q(g)$ whenever:
\begin{itemize}
\item $f$ and $-g_j$ are convex, $j=1,\ldots,n$,
Slater's condition holds and $\nabla^2f(\mathbf{x}^*)\succ0$ at the unique global minimizer $\mathbf{x}^*\in\mathbf{K}$; see e.g.,
de Klerk and Laurent \cite{deklerk-laurent}.
\item $\mathbf{K}\subseteq\{0,1\}^n$, i.e., for 0/1 polynomial programs, and more generally for
all discrete polynomial optimization problems.
\item $Q(g)$ is Archimedean, $f$ has finitely many zeros in $\mathbf{K}$ and the {\em Boundary Hessian Condition} (BHC) holds
at every zero of $f$ in $\mathbf{K}$; see e.g. Marshall \cite{marshall2}. That is, the BHC holds at a zero $\mathbf{x}^*\in\mathbf{K}$
of $f$ if there exists $0\leq k\leq n$, and some $1\leq v_1\leq \cdots\leq v_k\leq m$, such that:
$g_{v_1},\ldots, g_{v_k}$ are part of a system of local parameters at $\mathbf{x}^*$, and the standard sufficient conditions for $\mathbf{x}^*$ to be local
minimizer of $f$ on $\{\mathbf{x}:g_{v_j}(\mathbf{x})\geq0,\:j=1,\ldots,k\}$ hold. Equivalently, if
$(t_1,\ldots,t_n)$ are local parameters at $\mathbf{x}^*$ with $t_j=g_{v_j}$, $j=1,\ldots,k$,
then $f$ can be written as a formal power series $f_0+f_1+f_2\ldots$ in $\mathbb{R}[[t_1,\ldots,t_n]]$, where
each $f_j$ is an homogeneous form of degree $j$,
\[f_1\,=\,a_1\,t_1+\cdots +a_k\,t_k\quad\mbox{ with }a_i>0,\quad i=1,\ldots,k,\]
and the quadratic form $f_2(\underbrace{0,\ldots, 0}_{k \mbox{ times}},t_{k+1},\ldots,t_n)$ is positive definite.
For instance, if all the (finitely many) zeros $\mathbf{x}^*$ of $f$ are in the interior of $\mathbf{K}$, one may take
$k=0$ and $t_i(\mathbf{x})=x_i-x^*_i$ for all $i$. Then $f\in Q(g)$ if $f_2$ is positive definite.
\end{itemize}
It turns out that under a technical condition on the polynomials that define $\mathbf{K}$, the BHC holds generically, i.e.,
the set of polynomials $f\in\mathbb{R}[\mathbf{x}]_d$ for which the BHC holds at every global minimizer on $\mathbf{K}$,
is dense in $\mathbb{R}[\mathbf{x}]_d$; see Marshall \cite[Corollary 4.5]{marshall2}. Finally, and in the same vein,
a recent result of Nie \cite{nie} states that if
the standard constraint qualification, strict complementarity and second-order sufficiency condition hold
at every global minimizer of $f$ on $\mathbf{K}$, then $f-f^*\in Q(g)$. Moreover this property is also generic in the sense that
it does not hold only if the coefficients of the polynomials $f$ and $g_j$, $j=1,\ldots,m$,
satisfy a set of polynomial equations! In other words the three conditions
hold in a Zariski open set; for more details see Nie \cite[Theorem 1.1, Theorem 1.2 and \S 4.2]{nie}.
So in view of Proposition \ref{prop-asymptotics}, one may expect that $\lim_{d\to\infty}\hat{\rho}^k_d=\rho^k$ generically. On the other hand,
given $d\in\mathbb{N}$, identifying whether $\rho^k_d=\rho^k$ (or whether $\vert \rho^k_d-\rho^k\vert<\epsilon$ for some given $\epsilon>0$, fixed) is a open issue. For instance, $\rho^k_d=\rho^k_{d+1}$ (as is the case in Examples \ref{newex1} and \ref{newex2} below) is not a guarantee that $\rho^k_d=\rho^k$.
We will also see in Section \S \ref{section-epsilon} that one may also approach
as closely as desired an optimal solution of the ideal inverse problem (\ref{inv-0}) by asking $\mathbf{y}$ to be only
a global $\epsilon$-minimizer (with $\epsilon>0$ fixed), provided that $\epsilon$ is small enough.
In addition, more can be said by looking at the dual of (\ref{inv-0}).
\subsection*{The dual of the ideal inverse problem $\mathcal{P}$}
We now provide an explicit interpretation of the dual problems $\mathbf{P}^*_d$ in
(\ref{dual-k=1})-(\ref{dual-k=infty}).
Let $M(\mathbf{K})$ be the space of finite Borel measures on $\mathbf{K}$.
Then obviously (\ref{dual-k=1}) is a relaxation of the following problem:
\begin{equation}
\label{newdual-k=1}
\left\{\begin{array}{ll}
\mathbf{r}^1=\displaystyle\max_{\mu\in M(\mathbf{K})}&\displaystyle\int_\mathbf{K} (f(0)-f(\mathbf{x}))\,d\mu(\mathbf{x})\\
\mbox{s.t.}&\pm\displaystyle\int_\mathbf{K} \mathbf{x}^\alpha d\mu(\mathbf{x})\,\le\,1,\quad \forall\,\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\},
\end{array}\right.
\end{equation}
which, denoting by $\delta_0$ the Dirac measure at $\mathbf{x}=0$, and by $P(\mathbf{K})$ the space of Borel probability measures on $\mathbf{K}$,
can be rewritten as
\[\left\{\begin{array}{rl}
\displaystyle\max_{\nu\in P(\mathbf{K}),\gamma\geq0}&\gamma\, (\delta_0(f)-\nu(f))\\
\mbox{s.t.}&
\pm\gamma\,\displaystyle (\nu(\mathbf{x}^\alpha)-\delta_0(\mathbf{x}^\alpha))\,\le\,1,\quad \forall\,\alpha\in\mathbb{N}^n_{d_0};\quad\nu(\mathbf{K})=\gamma.
\end{array}\right.\]
Similarly, (\ref{dual-k=infty}) is a relaxation of the following problem:
\begin{equation}
\label{newdual-k=infty}
\mathbf{r}^\infty=\displaystyle\max_{\mu\in M(\mathbf{K})}\,\left\{\displaystyle\int_\mathbf{K} (f(0)-f(\mathbf{x}))\,d\mu(\mathbf{x})\::\:
\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\left\vert\int_\mathbf{K} \mathbf{x}^\alpha d\mu\right\vert\,\le\,1\,\right\},
\end{equation}
or, again, equivalently,
\[\displaystyle\max_{\nu\in P(\mathbf{K}),\gamma\geq0}\,\left\{\gamma\,(\delta_0(f)-\nu(f))\::\:
\gamma\,\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}}\left\vert\nu(\mathbf{x}^\alpha)-\delta_0(\mathbf{x}^\alpha)\right\vert\,\le\,1;
\quad\nu(\mathbf{K})=\gamma\,\right\}.\]
Hence, in the dual problems (\ref{newdual-k=1}) and (\ref{newdual-k=infty})
one searches for a finite Borel measure $\mu$
which concentrates as much as possible on the set $\{\mathbf{x}\in\mathbf{K}\,:\,f(\mathbf{x})\leq f(0)\}$, and such
that its moments up to order $d_0$ are not too far from those of a measure supported at $\{0\}\in\mathbf{K}$.
In fact, and as one might have expected, (\ref{newdual-k=1}) (resp.
(\ref{newdual-k=infty})) is the dual of $\mathcal{P} $ in (\ref{inv-0}) with $k=1$ (resp. with $k=\infty$).
For instance, with $k=1$, to see that weak duality holds,
let $\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0}$ and $\mu\in M(\mathbf{K})$ be
an arbitrary feasible solution of (\ref{inv-0}) and (\ref{newdual-k=1}), respectively. Then:
\begin{eqnarray*}\displaystyle\int_\mathbf{K} (f(0)-f)\,d\mu&=&
\underbrace{\displaystyle\int_\mathbf{K} (\tilde{f}(0)-\tilde{f})d\mu}_{\leq 0}+\int_\mathbf{K} (\tilde{f}-f)\,d\mu\\
&\leq&\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\vert \tilde{f}_\alpha-f_\alpha\vert\cdot\left\vert\int_\mathbf{K} \mathbf{x}^\alpha\,d\mu(\mathbf{x})\right\vert
\leq\displaystyle\Vert \tilde{f}-f\Vert_1,
\end{eqnarray*}
i.e., weak duality holds and $\mathbf{r}_1\leq\rho^1$. We even have the following:
\begin{lem}
\label{lem-asymp}
Let $\mathbf{K}$ in (\ref{setk}) be with nonempty interior and assume that
$Q(g)$ is Archimedean. Let $\rho^k$ be as in (\ref{inv-0}) with $k=1,\infty$, and
let $\mathbf{z}^d=(z^d_\alpha)\in\mathbb{R}^{s(d)}$ be a nearly optimal solution of (\ref{dual-k=1})
(or (\ref{equiv-1})), e.g., with value
$L_\mathbf{z}(f(0)-f)\geq \rho^k_d-1/d$, for all $d\in\mathbb{N}$.
If $\displaystyle\liminf_{d\to\infty}z^d_0<\infty$
then $\displaystyle\lim_{d\to\infty}\rho^k_d=\rho^k$ and (\ref{newdual-k=1}) has an optimal solution
$\mu^*\in M(\mathbf{K})$ which is supported on the set of global minimizers on $\mathbf{K}$ of the optimal solution $\tilde{f}^*\in\mathbb{R}[\mathbf{x}]_{d_0}$ of
(\ref{inv-0}) (which contains $\{0\}$). Hence either $\rho^k=0$ in which case $0$ is an optimal solution of $\mathbf{P}$, or
$\rho^k>0$ and $\tilde{f}^*$ has a another global minimizer $\tilde{\mathbf{x}}\neq0$ on $\mathbf{K}$ with $f(\tilde{\mathbf{x}})<f(0)$.
\end{lem}
\begin{proof}
The proof for the case $k=\infty$ is omitted as very similar to that of the case $k=1$.
Consider the subsequence $d_i$, $i\in\mathbb{N}$, such that $\displaystyle\liminf_{d\to\infty}z^d_0=\displaystyle\lim_{i\to\infty} z^{d_i}_0 <\infty$. Using the Archimedean property (\ref{archimedean}) of $Q(g)$, we proceed exactly as in the proof of Theorem 3.2 in \cite[p. 57--59 ]{lassjogo}.
There is a infinite sequence $\mathbf{z}^*=(z^*_\alpha)$, $\alpha\in\mathbb{N}^n$, and a subsequence (still denoted
$d_i$ for notational convenience), such that for every $\alpha\in\mathbb{N}^n$, $z^{d_i}_\alpha\to z^*_\alpha$.
Moreover, from the convergence $\mathbf{z}^{d_i}\to\mathbf{z}^*$, $\mathbf{M}_d(g_j\,\mathbf{z}^*)\succeq0$ for every $d\in\mathbb{N}$ and every
$ j=0,1,\ldots,m$; hence by Putinar's Theorem \cite{putinar}, $\mathbf{z}^*$ is the sequence of moments of a finite Borel measure $\mu^*$ supported on $\mathbf{K}$. Moreover,
since $\vert z^{d_i}_\alpha\vert\leq 1$ for all $\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}$, we obtain
\[\vert\int_\mathbf{K}\mathbf{x}^\alpha\,d\mu^*\vert\,=\,\vert z^*_\alpha\vert
=\lim_{i\to\infty}\vert z^{d_i}_\alpha\vert\,\leq \,1\quad\forall \alpha\in\mathbb{N}^n_{d_0}\setminus\{0\},\]
which proves that $\mu^*$ is feasible for (\ref{newdual-k=1}). Finally, by monotonicity of the
sequence $(\rho^1_d)$, $d\in\mathbb{N}$, and using $\rho^1_d\geq\rho^1\geq\mathbf{r}^1$ for all $d$,
\begin{eqnarray*}
\mathbf{r}^1\leq\rho^1\leq\lim_{d\to\infty}\rho^1_{d}&=&\lim_{i\to\infty}\rho^1_{d_i}\,=\,\lim_{i\to\infty}L_{\mathbf{z}^{d_i}}(f(0)-f)\\
&=&L_{\mathbf{z}^*}(f(0)-f)=\int_\mathbf{K}(f(0)-f)d\mu^*,\end{eqnarray*}
which proves that $\mu^*$ is an optimal solution of (\ref{newdual-k=1}), and so
$\mathbf{r}_1=\rho^1$.
Finally, since $\tilde{f}^*(0)=f(0)$,
\begin{eqnarray*}
\rho^1=\displaystyle\int_\mathbf{K} (f(0)-f)\,d\mu^*&=&\underbrace{\displaystyle\int_\mathbf{K} (\tilde{f}^*(0)-\tilde{f}^*)\,d\mu^*}_{\leq 0}
+\int_\mathbf{K} (\tilde{f}^*-f)\,d\mu^*\\
&\leq&\displaystyle\Vert \tilde{f}^*-f\Vert_1=\rho^1,
\end{eqnarray*}
which implies $\mu^*(\{\mathbf{x}:\,\tilde{f}^*(\mathbf{x})-\tilde{f}^*(0)>0\})=0$, that is,
the support of $\mu^*$ is contained in the set of global minimizers of $\tilde{f}^*$ (which contains $\{0\}$).
Therefore, if $\rho^1>0$ then necessarily there is another global minimizer $0\neq\tilde{\mathbf{x}}\in\mathbf{K}$ of $\tilde{f}^*$
with $f(\tilde{\mathbf{x}})<f(0)$, otherwise $\rho^1=\int (f(0)-f)d\mu^*=0$.
\end{proof}
\subsection{Convexity}
One may wish to restrict to search for convex polynomials $\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0}$ (no matter if $f$ itself is convex).
For instance if the $g_j$'s are concave (so that $\mathbf{K}$ is convex) but $f$ is not, one may wish to find the convex optimization problem
whose $\mathbf{y}\in\mathbf{K}$ is an optimal solution and with convex polynomial criterion $\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0}$ closest to $f$.
If $d_0>2$ then in (\ref{inv-1}) it suffices to add the additional Putinar's certificate
\begin{equation}
\label{convex}
(\mathbf{x},\mathbf{u})\mapsto\:\mathbf{u}^T\nabla^2\tilde{f}(\mathbf{x})\,\mathbf{u}\,=\,\sum_{j=0}^m\psi_j(\mathbf{x},\mathbf{u})\,g_j(\mathbf{x})+\psi_{m+1}(\mathbf{x},\mathbf{u})(1-\Vert\mathbf{u}\Vert^2),\end{equation}
with $\psi_{m+1}\in\mathbb{R}[\mathbf{x},\mathbf{u}]$ and
$\psi_j\in\Sigma_{d-v_j}[\mathbf{x},\mathbf{u}]$, for all $j=0,1,\ldots,m$. Indeed, (\ref{convex}) is a Putinar's certificate
of convexity for $\tilde{f}$ on $\mathbf{K}$, with degree bound $d$. As the coefficients of
the polynomial $(\mathbf{x},\mathbf{u})\mapsto\:\mathbf{u}^T\nabla^2\tilde{f}(\mathbf{x})\mathbf{u}$ are linear in the coefficients of $\tilde{f}$,
(\ref{convex}) will translate into additional semidefinite constraints in (\ref{inv-3-k=1}).
If $d_0\leq2$, i.e. if $f(\mathbf{x})=\frac{1}{2}\mathbf{x}^T\mathbf{A}\mathbf{x}+\mathbf{b}^T\mathbf{x}+c$ for some real symmetric matrix $\mathbf{A}\in\mathbb{R}^{n\times n}$, some
vector $\mathbf{b}\in\mathbb{R}^n$ and some scalar $c\in\mathbb{R}$, then
$\tilde{f}(\mathbf{x})=\frac{1}{2}\mathbf{x}^T\tilde{\mathbf{A}}\mathbf{x}+\tilde{\mathbf{b}}^T\mathbf{x}+\tilde{c}$ for some real symmetric matrix
$\tilde{\mathbf{A}}\in\mathbb{R}^{n\times n}$, some $\tilde{\mathbf{b}}\in\mathbb{R}^n$ and some $\tilde{c}\in\mathbb{R}$.
In that case, in (\ref{inv-1}) it suffices to add
constraint $\nabla^2\tilde{f}(\mathbf{x})=\tilde{\mathbf{A}}\succeq0$,
which is just a Linear Matrix Inequality (LMI).
And therefore, again, (\ref{inv-1}) can be rewritten
as a semidefinite program, namely (\ref{inv-3-k=1})-(\ref{inv-3-k=2}) with the additional LMI constraint
$\tilde{\mathbf{A}}\succeq0$.
Notice that for $k=1,2$, it also makes sense to search for
$\tilde{f}\in\mathbb{R}[\mathbf{x}]_2$ even if $f$ has degree $d_0>2$, i.e.,
if $f(\mathbf{x})=c+\mathbf{b}^T\mathbf{x}+\frac{1}{2}\mathbf{x}^T\mathbf{A}\mathbf{x} +h(\mathbf{x})$ where $h\in\mathbb{R}[\mathbf{x}]$ does not contains monomials
of degree smaller than $3$. This means that one searches for the convex program with quadratic cost
closest to $f$.
So for instance, in the case where one searches for $\tilde{f}\in\mathbb{R}[\mathbf{x}]_2$,
and given $\mathbf{y}\in\mathbf{K}$ let
$J(\mathbf{y}):=\{j\in\{1,\ldots,m\}\,:\,g_j(\mathbf{y})=0\}$ be the set of constraints that are active at $\mathbf{y}$.
If the $g_j$'s that define $\mathbf{K}$ are concave
then one may simplify (\ref{inv-1}). Writing $\tilde{f}=\frac{1}{2}\mathbf{x}^T\tilde{\mathbf{A}}\mathbf{x}+\tilde{\mathbf{b}}^T\mathbf{x}+\tilde{c}$, and with $k=1,2$,
(\ref{inv-1}) now reads:
\[\begin{array}{rl}
\rho^k:=\displaystyle\min_{\tilde{\mathbf{A}},\tilde{\mathbf{b}},\lambda}&\Vert f-\tilde{f}\Vert_k\\
\mbox{s.t.}& \tilde{\mathbf{A}}\,\mathbf{y}+\tilde{\mathbf{b}}\,=\,\displaystyle\sum_{j\in J(\mathbf{y})}\lambda_j\,\nabla g_j(\mathbf{y})\\
&\tilde{\mathbf{A}}\succeq0;\:\lambda_j\geq0,\quad j\in J(\mathbf{y}).
\end{array}\]
So, as we did in the previous section, and possibly after the change of variable $\mathbf{x}':=\mathbf{x}-\mathbf{y}$,
with no loss of generality one may and will assume that $\mathbf{y}=0$, in which case (\ref{inv-4}) simplifies to
\begin{equation}
\label{inv-4}
\begin{array}{rl}\rho^k:=\displaystyle\min_{\tilde{\mathbf{A}},\tilde{\mathbf{b}},\lambda}&\Vert f-\tilde{f}\Vert_k\\
\mbox{s.t.}& \tilde{\mathbf{b}}\,=\,\displaystyle\sum_{j\in J(0)}\lambda_j\,\nabla g_j(0)\\
&\tilde{\mathbf{A}}\succeq0;\:\lambda_j\geq0,\quad j\in J(0),
\end{array}\end{equation}
which in turn simplifies to
\begin{equation}
\label{inv-44}
\begin{array}{rcl}
\rho^1&=&\displaystyle\min_{\tilde{\mathbf{A}}\succeq0}\Vert \tilde{\mathbf{A}}-\mathbf{A}\Vert_1+
\displaystyle\min_{\lambda\geq0}\Vert \mathbf{b}-\sum_{j\in J(0)}\lambda_j\nabla g_j(0)\Vert_1\\
\rho^\infty&=&\sup\left[\displaystyle\min_{\tilde{\mathbf{A}}\succeq0}\Vert \tilde{\mathbf{A}}-\mathbf{A}\Vert_\infty\:,\:
\displaystyle\min_{\lambda\geq0}\Vert \mathbf{b}-\sum_{j\in J(0)}\lambda_j\nabla g_j(0)\Vert_\infty\,\right].
\end{array}
\end{equation}
Observe that (\ref{inv-44}) can be solved in two steps. One first
solves the problem $\min_{\lambda\geq0}\Vert \mathbf{b}-\sum_{j\in J(0)}\lambda_j\nabla g_j(0)\Vert_k$,
which is a linear program with finite value, hence with an optimal solution.
One next solves the problem $\min_{\tilde{\mathbf{A}}\succeq0}\Vert \tilde{\mathbf{A}}-\mathbf{A}\Vert_k$
which computes the $\ell_k$-projection of $\mathbf{A}$ onto the closed convex cone of positive semidefinite matrices
(a semidefinite program with an optimal solution).
\begin{lem}
\label{lem-convexity}
Let $\mathbf{K}\subset\mathbb{R}^n$ be as in (\ref{setk}) with $g_j$ being concave for every $j=1,\ldots,m$.
Then (\ref{inv-4}) has an optimal solution
$\tilde{f}^*\in\mathbb{R}[\mathbf{x}]_{2}$ and $0$ is an optimal solution of the convex optimization problem
$\mathbf{P}':\min\{\tilde{f}^*(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$.
\end{lem}
\begin{proof}
Let $(\tilde{f},\lambda)$ (with $\tilde{f}\in\mathbb{R}[\mathbf{x}]_{2}$) be any feasible solution of (\ref{inv-4}).
The constraint in (\ref{inv-4}) states that $\nabla L(0)=0$, where $L\in\mathbb{R}[\mathbf{x}]$
is the Lagrangian polynomial $\mathbf{x}\mapsto L(\mathbf{x}):=\tilde{f}(\mathbf{x})-\sum_{j\in J(0)}\lambda_j\,g_j(\mathbf{x})$, which
is convex on $\mathbf{K}$ because the $g_j$'s are concave, the $\lambda_j$'s are nonnegative, and
$\tilde{f}$ is convex. Therefore $\nabla L(0)=0$ implies that
$0$ is a global minimizer of $L$ on $\mathbb{R}^n$ and a global minimizer of $\tilde{f}$ on $\mathbf{K}$ because
\begin{equation}
\label{lagrange}
\tilde{f}(\mathbf{x})\,\geq\,L(\mathbf{x})\,\geq\,L(0)\,=\,\tilde{f}(0),\qquad\forall\,\mathbf{x}\in\mathbf{K}.\end{equation}
It remains to prove that (\ref{inv-4}) has an optimal solution $\tilde{f}^*$.
But we have seen that (\ref{inv-4}) is equivalent to (\ref{inv-44}) for which an optimal solution can be found
by solving a linear program and a semidefinite program.
\end{proof}
So in this case where the $g_j$'s are concave (hence $\mathbf{K}$ is convex), one obtains
the convex programming problem with quadratic cost, whose criterion is the closest to $f$ for
the $\ell_k$-norm.
\section{Global $\epsilon$-optimality}
\label{section-epsilon}
One may be less demanding and ask $\mathbf{y}\in\mathbf{K}$ (or $0\in\mathbf{K}$ after a change of variable)
to be only a global $\epsilon$-minimizer. That is, one searches for a polynomial $\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0}$ as close as possible to $f$ and such that
$\tilde{f}(\mathbf{x})-\tilde{f}(0)\geq -\epsilon$ for all $\mathbf{x}\in\mathbf{K}$ and some $\epsilon>0$ fixed. Then we will see that one may approach
as closely as desired an optimal solution of the ideal inverse problem (\ref{inv-0}).
With $\mathbf{K}\subset [-1,1]^n$ as in (\ref{setk-1}), the analogue of Problem (\ref{inv-1}) reads:
\begin{equation}
\label{invep-1}
\begin{array}{rl}
\rho_{d\epsilon}^k:=\displaystyle\min_{\tilde{f},\sigma_j\in\mathbb{R}[\mathbf{x}] }&\Vert f-\tilde{f}\Vert_k\\
\mbox{s.t.}& \tilde{f}(\mathbf{x})-\tilde{f}(0)+\epsilon\,=\,\displaystyle\sum_{j=0}^{m+n} \sigma_j(\mathbf{x})\,g_j(\mathbf{x}),\quad\forall\mathbf{x}\in\mathbb{R}^n\\
&\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0};\:\sigma_j\in\Sigma[\mathbf{x}]_{d-v_j},\quad j=0,1,\ldots,m+n.
\end{array}
\end{equation}
For instance, with $k=1$ (\ref{inv-3-k=1}) now reads
\begin{equation}
\label{invep-3-k=1}
\begin{array}{rl}
\rho_{d\epsilon}^1:=\displaystyle\min_{\lambda\geq0,\tilde{f},\mathbf{Z}_j}&\displaystyle\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}\lambda_\alpha\\
\mbox{s.t.}& \lambda_\alpha +\tilde{f}_\alpha\geq f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
& \lambda_\alpha -\tilde{f}_\alpha\geq -f_\alpha,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\langle\mathbf{Z}_0,\mathbf{B}_\alpha\rangle+\displaystyle\sum_{j=1}^{m+n} \langle\mathbf{Z}_j,\mathcal{C}^j_\alpha\rangle\,=\,\left\{\begin{array}{l}
\tilde{f}_\alpha,\:\mbox{ if }0<\vert\alpha\vert\leq\,d_0\\
0,\:\mbox{ if $\vert\alpha\vert>d_0$}\\
\epsilon,\:\mbox{ if $\alpha=0$}\end{array}\right.\\
&\mathbf{Z}_j\succeq0,\quad j=0,1,\ldots,m+n.\end{array}
\end{equation}
while its dual reads
\begin{equation}
\label{epdual-k=1}
\left\{\begin{array}{ll}
(\rho^1_{d\epsilon})^*=\displaystyle\max_{\mathbf{u},\mathbf{v}\geq0,\mathbf{z}}&-\epsilon z_0+\displaystyle
\sum_{\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}}f_\alpha(u_\alpha-v_\alpha)\:(=L_\mathbf{z}(f(0)-f-\epsilon))\\
\mbox{s.t.}&u_\alpha+v_\alpha\,\leq \,1,\quad\forall\,\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&u_\alpha-v_\alpha+z_\alpha \,=\,0,\quad\forall\alpha\in\mathbb{N}^n_{d_0}\setminus\{0\}\\
&\mathbf{M}_d(\mathbf{z}),\,\mathbf{M}_{d-v_j}(g_j\,\mathbf{z})\,\succeq\,0,\quad j=1,\ldots,m+n.
\end{array}\right.
\end{equation}
Again with no loss of generality and possibly after a change of variable, we assume that $\mathbf{y}=0\in\mathbf{K}$.
\begin{lem}
\label{lem-ep}
Let $\mathbf{K}$ be as in (\ref{setk-1}) and let $\rho^k$ be the optimal value of the ideal inverse problem $\mathcal{P}$ in (\ref{inv-0}).
Then for every fixed $\epsilon>0$ there exists $d_\epsilon\in\mathbb{N}$ such that $\rho^k_{d\epsilon}\leq \rho^k$ for
all $d\geq d_\epsilon$.
\end{lem}
\begin{proof}
Let $\tilde{f}^*\in\mathbb{R}[\mathbf{x}]_{d_0}$ be an optimal solution of the ideal inverse problem $\mathcal{P}$ with value $\rho^k=\Vert f-\tilde{f}^*\Vert_k$.
Observe that the polynomial
$\mathbf{x}\mapsto \tilde{f}^*(\mathbf{x})-\tilde{f}^*(0)+\epsilon$ is strictly positive on $\mathbf{K}$ and so by Theorem \ref{th-put}
it belongs to $Q(g)$; and so it belongs to $Q_d(g)$ as soon as $d\geq d_\epsilon$ (for some $d_\epsilon\in\mathbb{N}$).
Hence $\tilde{f}^*$ is a feasible solution of (\ref{invep-1}) which implies the desired result $\rho^k\geq\rho^k_{d\epsilon}$ for all $d\geq d_\epsilon$.
\end{proof}
The following analogue of Theorem \ref{th1}
shows that the optimal value $\rho_{d\epsilon}^k$ of (\ref{invep-1})
is still helpful to bound the quantity $f(0)-f^*$
(where $f^*$ is the global optimum of problem $\mathbf{P}$).
\begin{thm}
\label{th-ep}
Assume that $\mathbf{K}$ in (\ref{setk-1}) has nonempty interior and
let $\mathbf{x}^*\in\mathbf{K}$ be a global minimizer of $\mathbf{P}$ with optimal value $f^*$.
For every $\epsilon>0$ fixed, let $\tilde{f}_{d\epsilon}\in\mathbb{R}[\mathbf{x}]_{d_0}$ be an optimal solution of (\ref{invep-1}) with optimal value $\rho^k_{d\epsilon}$. Then:
{\rm (a)} $0\in\mathbf{K}$ is a global $\epsilon$-minimizer of the problem $\tilde{f}^*_{d\epsilon}=\min_\mathbf{x}\{\tilde{f}_{d\epsilon}(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$. In particular,
if $\rho_{d\epsilon}^k=0$ then $\tilde{f}_{d\epsilon}=f$ and $0$ is a global $\epsilon$-minimizer of $\mathbf{P}$.\\
{\rm (b)} If $k=1$ then $f^*\,\leq\,f(0)\,\leq\,f^*+\epsilon+\rho_{d\epsilon}^1$.\\
{\rm (c)} If $k=\infty$ then $f^*\,\leq\,f(0)\,\leq\,f^*+\epsilon+s(d_0)\,\rho_{d\epsilon}^\infty$.
\end{thm}
The proof is omitted as being a verbatim copy of that of Theorem \ref{th1} and using
\[f^*=f(\mathbf{x}^*)=\underbrace{f(\mathbf{x}^*)-\tilde{f}_d(\mathbf{x}^*)}+\underbrace{\tilde{f}_d(\mathbf{x}^*)-\tilde{f}_d(0)}_{\geq -\epsilon}+\tilde{f}_d(0).\]
Concerning the canonical form associated with the $\ell_1$-norm we also have the following analogue of
Theorem \ref{th1-ell1}.
\begin{thm}
\label{th1ep-ell1}
Assume that $\mathbf{K}$ in (\ref{setk-1}) has a nonempty interior and for every $\epsilon>0$,
let $\tilde{f}_{d\epsilon}\in\mathbb{R}[\mathbf{x}]_{d_0}$ be an optimal solution of (\ref{invep-1}) with optimal value $\rho_{d\epsilon}^1$
for the $\ell_1$-norm. Then $\tilde{f}_{d\epsilon}$ is of the form:
\begin{equation}
\label{th1ep-ell1-1}
\tilde{f}_{d\epsilon}(\mathbf{x})\,=\,f(\mathbf{x})+\mathbf{b}'\mathbf{x}+\sum_{i=1}^n\lambda_i^*\,x_i^2,
\end{equation}
for some vector $\mathbf{b}\in\mathbb{R}^n$ and some
nonnegative vectors $\lambda^*\in\mathbb{R}^n$, optimal solution of the semidefinite program:
\[\begin{array}{rl}
\rho_{d\epsilon}^1:=\displaystyle\min_{\lambda,\gamma,\mathbf{b}}&\displaystyle
\Vert\mathbf{b}\Vert_1+\sum_{i=1}^n\lambda_i\\
\mbox{s.t.}& f-f(0)+\mathbf{b}'\mathbf{x}+\displaystyle\sum_{i=1^n}\lambda_i\,x_i^2+\epsilon\in Q_{d}(g),\quad
\lambda\,\geq0.
\end{array}\]
(And $\lambda^*=0$ if $d_0=1$.)
\end{thm}
We end up with the analogue of Proposition \ref{prop-asymptotics} for the asymptotics as $d\to\infty$.
For every $\epsilon>0$, let us call $\mathcal{P}_\epsilon$ the analogue of problem $\mathcal{P} \,(=\mathcal{P}_0)$, i.e.,
\begin{equation}
\label{invep-0}
\mathcal{P}_\epsilon:\quad\rho^k_\epsilon=
\displaystyle\min_{\tilde{f}\in\mathbb{R}[\mathbf{x}]_{d_0}}\:\{\,\Vert \tilde{f}-f\Vert_k\::\:\mathbf{x}\mapsto \tilde{f}(\mathbf{x})-\tilde{f}(\mathbf{y})+\epsilon\,\in\,{\rm Psd}_{d_0}(\mathbf{K})\:\}.
\end{equation}
\begin{prop}
\label{thep-ideal}
For every $\epsilon>0$ fixed, Problem (\ref{invep-0}) has an optimal solution $\tilde{f}^*_\epsilon\in\mathbb{R}[\mathbf{x}]_{d_0}$,
and $\rho^k_\epsilon=0$ if and only if $\mathbf{y}$ is a global $\epsilon$-minimizer of $\mathbf{P}$.
\end{prop}
The proof is similar to that of Theorem \ref{inv-0}.
Next, interestingly, we are able to relate (\ref{invep-1}) and the ideal inverse problem (\ref{inv-0}) as $\epsilon\to 0$.
\begin{prop}
\label{thep-ep=0}
Let $\mathbf{K}$ in (\ref{setk-1}) be with nonempty interior.
Let $\mathbf{y}=0\in\mathbf{K}$ and $\rho^k$ be the optimal value of the ideal inverse problem $\mathcal{P}$ in (\ref{inv-0})
and let $\tilde{f}_{d\epsilon}\in\mathbb{R}[\mathbf{x}]_{d_0}$ (resp. $\tilde{f}^*_\epsilon\in\mathbb{R}[\mathbf{x}]_{d_0}$) be any optimal solution of
(\ref{invep-1}) (resp. (\ref{invep-0})).
(i) Let $\epsilon_\ell>0$, $\ell\in\mathbb{N}$, be such that $\epsilon_\ell\to0$ as $\ell\to\infty$.
Then
every accumulation point $\hat{f}\in\mathbb{R}[\mathbf{x}]_{d_0}$ of the sequence $(\tilde{f}^*_{\epsilon_\ell})\subset\mathbb{R}[\mathbf{x}]_{d_0}$, $\ell\in\mathbb{N}$,
is an optimal solution of the ideal inverse problem (\ref{inv-0}).
(ii) If for every $\epsilon_\ell>0$, $d_\ell\in\mathbb{N}$ is sufficiently large, every accumulation point of the sequence
$(\tilde{f}_{d_\ell\epsilon_\ell})\subset\mathbb{R}[\mathbf{x}]_{d_0}$ is an optimal solution of the ideal inverse problem (\ref{inv-0}).
\end{prop}
\begin{proof}
(i) As $\Vert \tilde{f}^*_{\epsilon_\ell}-f\Vert_k\leq \rho^k$ for all $\ell$, the sequence
$(\tilde{f}^*_{\epsilon_\ell})\subset\mathbb{R}[\mathbf{x}]_{d_0}$, $\ell\in\mathbb{N}$, has accumulation points. So consider an arbitrary
converging subsequence (still denoted $(\tilde{f}^*_{\epsilon_\ell})$ for notational convenience) $\tilde{f}^*_{\epsilon_\ell}\to\hat{f}\in\mathbb{R}[\mathbf{x}]_{d_0}$, as $\ell\to\infty$.
Observe that $\tilde{f}^*_{\epsilon_\ell}(\mathbf{x})-\tilde{f}^*_{\epsilon_\ell}(0)+\epsilon_\ell\geq 0$ for all $\mathbf{x}\in\mathbf{K}$ and all $\ell\in\mathbb{N}$.
With $\mathbf{x}\in\mathbf{K}$ fixed, arbitrary, letting $\ell\to\infty$ yields $\hat{f}(\mathbf{x})-\hat{f}(0)\geq 0$, and so
$\mathbf{x}\mapsto \hat{f}(\mathbf{x})-\hat{f}(0)\in{\rm Psd}_{d_0}(\mathbf{K})$. Moreover,
$\rho^k_{\epsilon_\ell}\leq\rho^k$ for all $\ell$ yields $\Vert f-\hat{f}\Vert_k\leq \rho^k$, which in turn implies
that $\hat{f}$ is an optimal solution of the inverse problem (\ref{inv-0}).
As the accumulation point $\hat{f}$ was arbitrary the result follows.
The proof of (ii) is similar
if one recalls that by Lemma \ref{lem-ep}, $\rho^k_{d\epsilon_\ell}=\Vert f-\tilde{f}_{d\epsilon_\ell}\Vert_k\leq\rho^k$ for all $\ell$ and
all $d$ sufficiently large, say $d\geq d_{\ell}$.
\end{proof}
Hence, by asking $\mathbf{y}\,(=0)\in\mathbf{K}$ to be only a global $\epsilon$-minimizer, one may obtain
a polynomial $\tilde{f}_{d\epsilon}\in\mathbb{R}[\mathbf{x}]_{d_0}$ as close as desired to an optimal solution of the ideal inverse problem
provided that $\epsilon$ is sufficiently small and $d$ is sufficiently large.
\subsection{Illustrative examples and discussion}
We here provide some simple illustrative examples and
show that the representation of the set $\mathbf{K}$ may be important for getting a Putinar certificate
faster.
\begin{ex}
\label{ex1}
{\rm Let $n=2$ and consider the optimization problem
$\mathbf{P}:\:f^*\,=\,\min_\mathbf{x}\:\{f(\mathbf{x})\,:\,\mathbf{x}\in\mathbf{K}\}$
with $\mathbf{x}\mapsto f(\mathbf{x})=x_1+x_2$, and
\[\mathbf{K}\,=\,\{\mathbf{x}\in\mathbb{R}^2\::\: x_1x_2\geq1;\: 1/2\leq \mathbf{x}\leq 2\:\}.\]
The polynomial $f$ is convex and the set $\mathbf{K}$ is convex as well, but the polynomials that define $\mathbf{K}$ are not all concave.
That is, $\mathbf{P}$ is a convex optimization problem, but not a convex programming problem.
The point $\mathbf{y}=(1,1)\in\mathbf{K}$ is a global minimizer and the KKT conditions at $\mathbf{y}$ are satisfied with
$\lambda=(1,0,0)\in\mathbb{R}^3$, i.e.,
\[\nabla f(\mathbf{x})-\lambda_1\nabla g_1(\mathbf{x})\,=\,0\mbox{ with }\mathbf{x}=(1,1)\mbox{ and }\lambda_1 =1.\]
However, the Lagrangian
\[\mathbf{x}\mapsto L(\mathbf{x})\,:=\,f(\mathbf{x})-f^*-\lambda_1\,g_1(\mathbf{x})\,=\,x_1+x_2-1-x_1x_2,\]
is not convex and $(1,1)$ is not a global minimizer of $L$ on $\mathbb{R}^2$.
This example just illustrates the fact that even in the convex case
where the $g_j$'s are not concave, the KKT conditions do not provide a {\it certificate} of global optimality, contrary to ``convex programming" where since $L$ is now convex, obviously using $L(\mathbf{x})\geq L(\mathbf{y})=0$ (because $\nabla L(\mathbf{y})=0$),
\[f(\mathbf{x})-f^*\geq\,L(\mathbf{x})\,\geq\,L(\mathbf{y})\,=\,0,\]
whenever $\mathbf{x}\in\mathbf{K}$, and so $f(\mathbf{x})\geq f^*$ for all $\mathbf{x}\in\mathbf{K}$, the desired certificate of global optimality.
Next, if we now use the test of inverse optimality with $d=1$,
one searches for a polynomial $\tilde{f}_d$ of degree at most $d_0=1$, and such that
\[\tilde{f}_d(\mathbf{x})-\tilde{f}_d(1,1)\,=\,\sigma_0(\mathbf{x})+\sigma_1(\mathbf{x})(x_1x_2-1)+\sum_{i=1}^2\psi_i(\mathbf{x}) (2-x_i)
+\phi_i(\mathbf{x}) (x_i-1/2),\]
for some s.o.s. polynomials $\sigma_1,\psi_i\,\phi_i\in\Sigma[\mathbf{x}]_0$ and some s.o.s. polynomial $\sigma_0\in\Sigma[\mathbf{x}]_1$.
But then necessarily $\sigma_1=0$ and $\psi_i,\phi_i$ are constant, which in turn implies that
$\sigma_0$ is a constant polynomial. A straightforward calculation shows that $\tilde{f}_1(\mathbf{x})=0$ for all $\mathbf{x}$, and so
$\rho^1_1=2$. And indeed this is confirmed when solving\footnote{To solve
(\ref{dual-k=1}) we have used the GloptiPoly software
of Henrion et al. \cite{gloptipoly}, and dedicated to solving the Generalized Problem of Moments whose
problem (\ref{dual-k=1}) is only a special case.} (\ref{dual-k=1}) with $d=1$. Solving again (\ref{dual-k=1}) with now $d=2$ yields
$\rho^1_2=2$ (no improvement) and with $d=3$ we obtain the desired result $\rho^1_3=0$.
On the other hand, if now $\mathbf{K}$ has the representation:
\[\{\mathbf{x}\::\:x_1x_2-1\,\geq\,0;\quad (x_i-1/2)(2-x_i)\geq\,0;\quad i=1,2\:\},\]
then the situation differs because in fact
\[x_1+x_2-2\,=\,\frac{1}{5}+\frac{2}{5}(x_1-x_2)^2+\frac{4}{5}(x_1x_2-1)+\frac{2}{5}\sum_{i=1}^2(x_i-1/2)(2-x_i),\]
i.e., $f-f^*$ has a Putinar's certificate with degree bound $d=1$. Hence the test of inverse optimality yields $\rho^1_1=0$ with $\tilde{f}_1=f$.
}\end{ex}
The above example illustrates that the representation of $\mathbf{K}$ may be important.
\begin{ex}
\label{newex1}
{\rm Again consider Example \ref{ex1} but now with
$\mathbf{y}=(1.1,1/1.1)\in\mathbf{K}$, which is not a global optimum of $f$ on $\mathbf{K}$ any more. By solving
(\ref{dual-k=1}) with $d=1$ we still find $\rho^1_1=2$ (i.e., $\tilde{f}_1=0$), and with $d=2$ we find $\rho_2\approx 0.1734$ and
$\tilde{f}_2(\mathbf{x})\approx 0.8266\,x_1+x_2$.
And indeed by solving (using GloptiPoly) the new optimization problem with criterion $\tilde{f}_2$
we find the global minimizer $(1.1,0.9091)\approx\mathbf{y}$. With $d=3$ we obtains the same value $\rho^1_3=0.1734$, suggesting (but with no guarantee)
that $\tilde{f}_2$ is already an optimal solution of the ideal inverse problem.
}\end{ex}
\begin{ex}
\label{newex2}
{\rm Consider now the disconnected set $\mathbf{K}:=\{\mathbf{x}\,:\, x_1x_2\geq1;\:x_1^2+x_2^2 <= 3\}$
and the non convex criterion $\mathbf{x}\mapsto f(\mathbf{x}):=-x_1-x_2^2$ for which
$\mathbf{x}^*=(-0.618,-1/0.618)\in\partial\mathbf{K}$ is the unique global minimizer.
Let $\mathbf{y}:=(-0.63,-1/0.63)\in\partial\mathbf{K}$ for which the constraint $x_1x_2\geq1$ is active.
At steps $d=2$ and $d=3$ one finds that
$\tilde{f_d}=0$ and $\rho^1_d=\Vert f\Vert_1$. That is, $\mathbf{y}$ is a global minimizer of the trivial
criterion$\tilde{f}(\mathbf{x})=0$ for all $\mathbf{x}$, and cannot be a global minimizer
of some non trivial polynomial criterion.
Now let $\mathbf{y}=(-0.63,-\sqrt{3-0.63^2})$ so that the constraint $x_1^2+x_2^2<=3$ is active.
With obtain $\rho^1_1=\Vert f\Vert_1$ and $\tilde{f}_1=0$. With $d=2$
we obtain $\tilde{f}_2=1.26\,x_1-x_2^2$. With $d=3$ we obtain the same result, suggesting (but with no guarantee) that
$\tilde{f}_2$ is already an optimal solution of the ideal inverse optimization problem.
}\end{ex}
\begin{ex}
{\rm Consider the MAXCUT problem $\max \{\mathbf{x}'\mathbf{A}\mathbf{x}\,:\,\mathbf{x}_i^2=1,i=1,\ldots,n\}$ where
$\mathbf{A}=\mathbf{A}'\in\mathbb{R}^{n\times n}$ and $\mathbf{A}_{ij}=1/2$ for all $i\neq j$. For $n$ odd,
an optimal solution is $\mathbf{y}=(y_j)$ with $y_j=1$, $j=1,\ldots\lceil n/2\rceil$, and $y_j=-1$ otherwise.
However, the first semidefinite relaxation
\[\displaystyle\max\:\{\lambda \::\: \mathbf{x}'\mathbf{A}\mathbf{x}-\lambda=\sigma+\sum_{j=1}^n\gamma_i(x_i^2-1)
;\:\sigma\in\Sigma[\mathbf{x}]_1;\:\lambda,\gamma\in\mathbb{R}\}\]
provides the lower bound $-n/2$ (with famous Goemans-Williamson ratio guarantee).
So $\mathbf{y}$ cannot be obtained from the first semidefinite relaxation even though it is
an optimal solution. The inverse optimization problem reads:
Find the quadratic form $\mathbf{x}\mapsto\mathbf{x}'\tilde{\mathbf{A}}\mathbf{x}$ such that
$\displaystyle\mathbf{x}'\tilde{\mathbf{A}}\mathbf{x}-\mathbf{y}'\tilde{\mathbf{A}}\mathbf{y}\,=\,\sigma+\sum_{j=1}^n\gamma_i(x_i^2-1)$,
for some $\sigma\in\Sigma[\mathbf{x}]_1,\:\lambda,\gamma\in\mathbb{R}$, and
which minimizes the $\ell_1$-norm $\Vert\mathbf{A}-\tilde{\mathbf{A}}\Vert_1$. This is an inverse optimization problem with structural constraints
as described in Section \ref{structural} (since we search for a quadratic form
and not an arbitrary quadratic polynomial $\tilde{f}_2$). Hence, solving (\ref{inv-33-k=1}) for $n=5$ with $\mathbf{y}$ as above, we find that
\[\tilde{\mathbf{A}}\,=\,\frac{1}{2}\left[\begin{array}{ccccc} 0&2/3&2/3&1&1\\2/3&0&2/3&1&1\\
2/3&2/3&0&1&1\\
1&1&1&0&1\\
1&1&1&1&0\end{array}\right],\]
that is, only the entries $(i,j)\in\{(1,2),(1,3),(2,3)\}$ are modified from $1/2$ to $1/3$.
}\end{ex}
\section*{Conclusion}
We have presented a paradigm for inverse polynomial optimization.
Crucial is Putinar's Positivstellensatz which provides us
with the desired certificate of global optimality for a given
feasible point $\mathbf{y}\in\mathbf{K}$ and a candidate criterion $\tilde{f}$. In addition, to some extent,
the size of the certificate can be adapted to the computational capabilities available. Finally, and remarkably, when using the
$\ell_1$-norm the resulting inverse optimal criterion $\tilde{f}$ has a simple and explicit
{\it canonical} form. We hope that the concept of inverse optimization will receive more attention
from the optimization community as it could even provide an alternative stopping criterion
at the current iterate $\mathbf{y}\in\mathbf{K}$ of any local optimization algorithm for solving the original problem $\mathbf{P}$.
|
\section{Introduction
\label{sec:Intro}}
In iron pnictide superconductors \cite{Hosono},
both spin and orbital degrees of freedom play important roles
on various electronic properties,
such as the high-$T_{\rm c}$ superconductivity, orthorhombic structure
transition, and magnetic transition.
As for the origin of the superconductivity,
fully-gapped sign-reversing $s$-wave ($s_\pm$-wave) state
had been studied based on spin fluctuation theories
\cite{Mazin,Kuroki,hirschfeld,chubukov}.
The origin of the spin fluctuations is the intra-orbital nesting
and the Coulomb interaction.
However, the robustness of $T_{\rm c}$ against randomness in iron pnictides
indicates the absence of sign-reversal in the superconducting (SC) gap
\cite{Onari-impurity,Sato-imp,Nakajima}.
Later, orbital-fluctuation-mediated $s$-wave state
without sign reversal ($s_{++}$-wave) had been proposed
based on the Hubbard-Holstein (HH) model
\cite{Kontani,Saito,Onari}.
The origin of the orbital fluctuations is the inter-orbital nesting
and the electron-phonon ($e$-ph) interactions due to
non-$A_{1g}$ optical phonons.
One of the merits of this scenario is the robustness of
the $s_{++}$-wave state against impurities.
Another merit is that the close relation between $T_{\rm c}$ and the
crystal structure revealed by Lee \cite{Lee},
{\it e.g.}, $T_{\rm c}$ becomes the highest
when the As$_4$ cluster is regular tetrahedron,
is automatically explained \cite{Saito}.
Moreover, orbital-fluctuation-mediated $s_{++}$-wave state scenario is
consistent with the large SC gap on the $z^2$-orbital band
in Ba122 systems \cite{Saito},
observed by bulk-sensitive laser ARPES measurement \cite{Shimo-Science}.
The ``resonance-like'' hump structure in the neutron
inelastic scattering \cite{res-exp}
is frequently explained as the spin-resonance due to
the sign reversal in the SC gap \cite{res-the1,res-the2}.
However, experimental hump structure is well reproduced
in terms of the $s_{++}$-wave SC state, rather than the $s_\pm$-wave SC state,
by taking the suppression
in the inelastic scattering $\gamma(\w)$ for $|\w|\le3\Delta$
in the SC state (dissipationless mechanism)
\cite{Onari-resonance,Yasui}.
To distinguish between both SC states, measurements of phonon spectral
function for $|\w|\lesssim 2\Delta$ would be useful
\cite{Scala}.
In the normal state, prominent
non-Fermi liquid transport phenomena in $\rho$ and $R_{\rm H}$ \cite{Matsuda}
are frequently ascribed to the evidence of spin fluctuations
\cite{Kontani-review}.
However, they are also explained by the development of
antiferro-orbital fluctuations \cite{Onari}.
In order to identify the mechanism of superconductivity,
we have to understand the origin of the ordered state,
although it is still unsolved in iron pnictides.
For example, in many heavy fermion superconductors, the SC phase
appears next to the spin-density-wave (SDW) state, indicating the
occurrence of spin-fluctuation-mediated superconductivity.
In contrast, the ordered state in iron pnictides
is not a simple SDW state: In fact, the tetragonal to orthorhombic
structure transition occurs at $T_S\sim 100$ K,
usually above the SDW transition temperature $T_{\rm N}$ \cite{review}.
In addition, large imbalance of
$xz$- and $yz$-orbitals at the Fermi level and
reconstruction of the Fermi surfaces (FSs) had been observed
by angle resolved photoemission spectroscopy (ARPES) measurements
\cite{Shimojima,ARPES2,ARPES3}.
These results indicate that the ferro orbital-density-wave (ODW)
is the origin of the orthorhombic structure transition.
The SDW state may originate from the in-plane anisotropy
in the exchange interaction ($J_{1a}\ne J_{1b}$) associated with the ODW order
\cite{frustration}.
In addition, prominent symmetry breaking $C_4\rightarrow C_2$ is realized
in detwinned 122 systems even above $T_S$ and $T_{\rm N}$,
under very small uniaxial pressure.
For example, it is recognized as the large
in-plane anisotropy in the resistivity
\cite{rho-nematic,Prosorov}
and the optical conductivity \cite{optical} at $T^*\sim200$K,
which is much higher than $T_S$ and $T_{\rm N}$.
Moreover, the reconstruction of the FSs starts at $T^*$
in detwinned Ba(Fe$_{1-x}$Co$_x$)$_2$As$_2$ \cite{ARPES3}.
The discovery of these ``electronic nematic phase''
would indicate that the the ODW fluctuations deveolop
divergently above $T_S$, and the structure transition
is (almost) the second-order.
Recently, prominent softening of shear modulus
in undoped and under-doped Ba(Fe$_{1-x}$Co$_x$)$_2$As$_2$ had been reported
by acoustic measurements \cite{Fernandes,Yoshizawa}.
Yoshizawa {\it et al.} \cite{Yoshizawa}
observed all shear moduli $C_{44}$, $C_{66}$ and $C_E$.
Recently, they had also performed systematic measurement for
$x=0\sim0.225$, and found that only $C_{66}$ shows the prominent softening
in both under- and over-doped systems \cite{comment4}.
Similar results were reported by Goto {\it et al.} independently \cite{Goto}.
This observation can be explained by the development of
ferro-orbital fluctuations \cite{Saito}
or spin-nematic fluctuations \cite{Fernandes}.
Considering large quadrupole-strain coupling in iron pnictides,
all the observations mentioned above
suggest the importance of orbital physics, and pose a serious
challenge for theories of iron pnictide superconductors.
In fact, Goto {\it et al.} have shown that $C_{66}$ is almost
independent of the magnetic field up to $\sim50$T,
indicating the non-magnetic origin of the softening \cite{Goto}.
Thus, the main features of the iron-pnictide superconductors would be
summarized as (i) the orthorhombic (or nematic) transition
accompanied by remarkable $C_{66}$ softening,
(ii) emergence of high-$T_{\rm c}$ superconductivity
next to the orthorhombic phase, and (iii) the stripe-type
magnetic order induced by the orthorhombicity.
The unified explanation has not been achieved as far as we know.
In this paper, we develop the orbital fluctuation theory
to explain abovementioned features (i)-(iii)
based on the random-phase-approximation (RPA) and beyond RPA.
In the RPA, large $O_{xz}$-antiferro-quadrupole (AFQ) fluctuations
are produced by the $e$-ph interactions,
while they do not produce the softening of $C_{66}$ nor $C_{44}$.
If we go beyond the RPA, however,
we find that the $O_{x^2-y^2}$-ferro-quadrupole (FQ)
fluctuations are brought by the ``two-orbiton process''
near the AFQ quantum-critical-point (QCP).
The induced FQ fluctuations cause the softening of $C_{66}$,
and the commensurate ferro-orbital order is realized at $T=T_S$.
It is predicted that the AFQ-QCP is located
at the FQ-QCP, which is the endpoint of the orthorhombic phase.
Near this multi-orbital QCPs,
the superconductivity is mainly caused by the AFQ fluctuations,
and the orthorhombic transition is brought by the FQ fluctuations.
Moreover, the stripe-type antiferro-magnetic state is induced
in the orbital-ordered state,
since the orbital polarization gives strong
in-plane anisotropy in the spin-nesting.
The present study gives a microscopic justification
for the anisotropic Heisenberg model description
in the SDW state \cite{frustration,Dai}.
There is a long history in the study of superconductivity
due to charge or orbital fluctuations in multiorbital systems,
starting from the exciton-assisted superconductivity
\cite{Little,Ginzburg,Hirsch,Sawa,Tesa}.
In multiorbital systems, on-site Coulomb interaction
is composed of the intra-orbital term $U$,
inter-orbital one $U'$, and Hund's or exchange term $J$.
Since $U$ is usually larger than $U'\approx U-2J$,
spin fluctuations induced by $U$ give non $s$-wave SC states.
However, Takimoto {\it et al} \cite{Takimoto} had shown that
charge (or orbital) fluctuations induced by $U'$ give
conventional $s$-wave SC state
if the relations $U'>U$ and $J\sim0$ are assumed.
Although this idea was applied to pnictides \cite{Ohno1},
the relation $U'\sim0.6U$ is realized in many compounds \cite{Miyake}.
Even if $U'\sim0.6U$ and $J\sim0.2U$,
the present authors had shown that the
orbital-fluctuation-mediated superconductivity is realized by
quadrupole-quadrupole interaction mediated by non-$A_{1g}$ phonons,
which works as the ``negative effective exchange $J_{\rm eff}$ for the
charge sector.'' \cite{Kontani}.
In iron pnictides, orbital fluctuations develop even when the dimensionless
$e$-ph coupling $\lambda$ due to Fe-ion optical phonons is just $\sim 0.2$,
according to the RPA \cite{Saito} and FLEX approximation \cite{Onari}.
As for As-ion $A_{1g}$ mode \cite{Ohno1},
orbital fluctuations develop only when $\lambda_{A_{1g}}\sim1$,
which is unrealistic in iron pnictides.
Recently, Yanagi {\it et al}. \cite{Ohno2}
added the ``orthorhombic phonon''
to the HH model proposed by the present authors \cite{Kontani,Saito,Onari},
and studied the ferro-orbital fluctuations.
However, neither high-$T_{\rm c}$ nor structure transition
are explained by their theory:
First, the ``orthorhombic-phonon'' is acoustic ($\w_\q\propto|\q|$)
although it is treated as optical in Ref. \onlinecite{Ohno2} incorrectly.
Their theory belongs to the cooperative Jahn-Teller structure transition
due to acoustic phonons like manganites; see details in
Appendix \ref{sec:ApB}.
In this case, the energy-scale of ``ferro-orbital fluctuations''
is too low ($\sim\w_\q$) to explain high-$T_{\rm c}$, since experimental
$T_{\rm c}$ is much higher than $\w_\q\sim10$K for $|\q|\sim 0.1\pi$
\cite{comment,Schrieffer}.
Small orthorhombicity $(a-b)/(a+b)\sim0.003$ in iron pnictides
is also inconsistent with the cooperative Jahn-Teller scenerio.
Second, the derived orbital order is ``incommensurate'' \cite{comment5},
which is inconsistent with the orthorhombic structure transition
and the $C_{66}$ softening.
These problems are resolved in the present theory since
both high-energy AFQ and low-energy ``commensurate'' FQ fluctuations
develop at the same time, without the necessity of fine-tuning
model parameters.
The former (latter) fluctuations give the
superconductivity (orthorhombic transition).
\section{Model Hamiltonian
\label{sec:model}}
First, we shortly explain the relation between five-orbital and
ten-orbital models for iron pnictides.
In this study, we set $x$ and $y$ axes parallel to the nearest Fe-Fe bonds,
and represent the $z^2$, $xz$, $yz$, $xy$, and $x^2-y^2$ $d$-orbitals
in the $xyz$-coordinate as 1,2,3,4, and 5, respectively.
The FSs are mainly composed of $t_{2g}$ orbitals ($l=2\sim4$),
although $e_g$ orbitals also play non-negligible roles
in producing orbital fluctuations \cite{Saito}.
Figure \ref{fig:lattice} shows the crystal structure of FeAs-plane.
Since the As-A (As-B) ions form the upper- (lower-) plane,
the unit cell contains Fe-A and Fe-B,
which we call the two-iron unit cell.
Since each Fe-ion contains five-orbitals,
the original tight-binding model for iron-pnictides
is given as the ``ten-orbital model''.
In Refs. \onlinecite{Kuroki,Miyake},
the authors introduced the gauge transformation
$|2,3\rangle \rightarrow -|2,3\rangle$ only for Fe-B sites,
which we call the ``unfold-gauge transformation''.
Due to this gauge transformation, the unit cell of the kinetic term
is halved to become the single-iron unit cell,
shown in Figure \ref{fig:lattice}.
In the obtained ``five-orbital model'',
the kinetic term is given as
\begin{eqnarray}
H_{0}=\sum_{ij;lm;\s}t_{lm}^{ij}c_{i, l\s}^\dagger c_{j, m\s} ,
\end{eqnarray}
where $i,j$ denotes the unit cell,
$l,m=1\sim5$ represent the $d$-orbital, and $\s=\pm1$ is the spin index.
$c_{i, l\s}^\dagger$ is the creation operator of the $d$-electron, and
$t_{lm}^{ij}$ with $i\ne j$ ($i=j$) is the hopping integral (local potential).
This five-orbital model is convenient to study the Eliashberg gap equation
\cite{Kuroki,Kontani}.
In studying the orbital physics, however,
we have to keep the fact in mind that
the sign of quadrupole operators ${\hat O}_{xz}$ and ${\hat O}_{yz}$
at Fe-B sites are reversed by the unfold-gauge transformation.
By taking care of this fact, we study the softening of shear moduli
based on the five-orbital model hereafter.
In Appendix \ref{sec:10}, we calculate the orbital fluctuations
using the original ten-orbital model,
and make comparison between results of two models.
\begin{figure}[!htb]
\includegraphics[width=0.8\linewidth]{fig1.eps}
\caption{
(Color online)
Crystal structure of FeAs-layer, in which the unit cell is
given by the two-iron unit cell composed of Fe-A, Fe-B, As-A, and As-B.
The single-iron unit cell is realized by
applying the ``unfold-gauge transformation''.
}
\label{fig:lattice}
\end{figure}
Figure \ref{fig:fig1} shows the FSs in the (a) five-orbital model
and the (b) ten-orbital model.
The FSs in (a) coincide with the FSs in (b)
if we fold the former FSs into the two-iron Brillouin zone (BZ).
We use the hopping parameters for LaOFeAs given in Ref. \onlinecite{Kuroki}.
The colors correspond to $2$ (green), $3$ (red), and $4$ (blue), respectively.
The inter-orbital nesting between the orbital 2 on FS $\a_2$
and orbital 4 on FS $\b_2$
causes the most divergent AFQ fluctuations.
Next, we introduce the $e$-ph interaction due to Fe-ion Einstein optical modes.
The Hamiltonian given in Eq. (4) of Refs. \onlinecite{Saito}
is simply rewritten as the following bilinear form
in the $xyz$-coordinate:
\begin{eqnarray}
H_{e-\rm{ph}} =\eta \sum_i\left( {\hat O}^i_{yz}u_x^i +{\hat O}^i_{xz}u_y^i
+{\hat O}^i_{xy}u_z^i \right),
\label{eqn:e-ph}
\end{eqnarray}
where $\eta=60e^2 a_d^2/7\sqrt{3}R_{\rm Fe-As}^4$;
$a_d$ is the radius of $d$-orbital
({\it e.g.}, Shannon crystal radius of Fe$^{2+}$ is $0.77$
${\buildrel _{\circ} \over {\mathrm{A}}}$), and
$R_{\rm Fe-As}\approx2.4$ ${\buildrel _{\circ} \over {\mathrm{A}}}$.
${\bm u}^i$ is the displacement vector of the $i$-th Fe-ion, and
${\hat O}^i_{\Gamma}$ ($\Gamma=xz,yz,xy$) is the charge quadrupole
operator given as
\begin{eqnarray}
{\hat O}^i_{\Gamma}\equiv \sum_{lm}^\pm o^{l,m}_\Gamma {\hat m}_{l,m}^i ,
\end{eqnarray}
where ${\hat m}_{l,m}^i \equiv \sum_\s c_{i,l\s}^\dagger c_{i,m\s}$,
and the coefficient is defined as
$o^{l,m}_{xz}= 7\langle l|{\hat x}{\hat z}|m \rangle$
for $\Gamma=xz$, where ${\hat \x}=x/r$ and so on.
The non-zero coefficients are given as
\begin{eqnarray}
&& o^{2,5}_{xz}=o^{3,4}_{xz}=\sqrt{3}o^{1,2}_{xz}=1,
\label{eqn:oxz}\\
&& -o^{3,5}_{yz}=o^{2,4}_{yz}=\sqrt{3}o^{1,3}_{yz}=1,
\label{eqn:oyz}\\
&& o^{2,3}_{xy}=-\sqrt{3}o^{1,4}_{xy}/2=1.
\label{eqn:oxy}
\end{eqnarray}
Be careful not to confuse ${\hat O}_{xz}$ with the $xz$-orbital operator.
Other two quadrupole operators are $O_{z^2}$ and $O_{x^2-y^2}$,
whose coefficients are respectively defined as
$o^{l,m}_{x^2-y^2}= (7/2)\langle l|({\hat x}^2-{\hat y}^2)|m \rangle$
and $o^{l,m}_{z^2}= (7/2\sqrt{3})\langle l|(3{\hat z}^2-1)|m \rangle$.
(They are written as $O_2^0$ and $O_2^2$ in literatures.)
The non-zero coefficients are given as
\begin{eqnarray}
&& o^{2,2}_{x^2-y^2}=-o^{3,3}_{x^2-y^2}=-(\sqrt{3}/2)o^{1,5}_{x^2-y^2}=1,
\label{eqn:ox2y2}\\
&& o^{1,1}_{z^2}=2o^{2,2}_{z^2}=2o^{3,3}_{z^2}=-o^{4,4}_{z^2}=-o^{5,5}_{z^2}=2/\sqrt{3}.
\label{eqn:oz2}
\end{eqnarray}
Expect for $\Gamma=z^2$, all the matrix elements of ${\hat o}_\Gamma$
with respect to the $t_{2g}$-orbital ($2\sim4$) are $\pm 1$.
Here, we derived the $e$-ph interaction based on the point-charge model.
Although $e$-ph interaction is also induced by the change
in the $d$-$p$ hopping, as discussed in Ref. \onlinecite{Yada},
we expect it is small since the weight of $p$-electron
on the Fermi surface is just $\sim5$\% in iron pnictides.
Fortunately, because of the Wigner-Eckart theorem,
the matrix elements of the local
quadrupole-phonon interaction is always given by
the quadrupole operator ${\hat O}^i_{\Gamma}$,
independently of the details of the interaction.
Since the magnitude of the hexadecapole-phonon interaction is
$(a_d/R_{\rm Fe-Fe})^2\sim0.1$ times that of the quadrupole-phonon interaction,
we can safely use Eq. (\ref{eqn:e-ph}).
\begin{figure}[!htb]
\includegraphics[width=0.8\linewidth]{fig2.eps}
\caption{
(Color online)
(a) FSs for $n=6.0$ in the unfolded model.
The colors correspond to $2$ (green), $3$ (red), and $4$ (blue), respectively.
The inter-orbital nesting between FS $\a_2$ (green) and FS $\b_2$ (blue)
causes the AFQ fluctuations.
(b) FSs for the original ten-orbital model.
}
\label{fig:fig1}
\end{figure}
Equation (\ref{eqn:e-ph}) means that the displacement $u_x$
produces the quadrupole potential ${\hat O}_{yz}$,
which causes the scattering of electrons between orbitals 2 and 4.
The $e$-ph interactions in Eq. (\ref{eqn:e-ph}) within $t_{2g}$-orbitals
are shown in Fig. \ref{fig:e-ph}.
Then, the phonon-mediated el-el interaction $V_{\rm{el-el}}^{\rm ph}$
is obtained by taking the contraction of ${\bm u}^i$, which gives
the local phonon Green function
$D(\tau)\equiv \langle T_{\tau} u_{\mu}^i ( \tau ) u_{\mu}^i ( 0 ) \rangle $
$( \mu = x,y,z )$.
By taking the Fourier transformation, we obtain
\begin{equation}
D(\omega_l) = \frac{2 \langle u^2 \rangle_0 \omega_D}{\omega_l^2 + \omega_D^2},
\label{eqn:D}
\end{equation}
where $\w_l=2\pi Tl$ is the Boson Matsubara frequency,
$\omega_D$ is the optical phonon frequency, and
$\sqrt{\langle u^2 \rangle_0} = \sqrt{ 1 / 2M_{\text{Fe}} \omega_D}$ is the
uncertainty in position for Fe ions; $\sqrt{\langle u^2 \rangle_0} = 0.044$
${\buildrel _{\circ} \over {\mathrm{A}}}$ for $\w_{\rm D}=0.02$ eV \cite{Saito}.
Then, $V_{\rm{el-el}}^{\rm ph}$ is expressed as the
following quadrupole-quadrupole interaction:
\begin{eqnarray}
V_{\rm{el-el}}^{\rm ph}=-g(\w_l) \sum_i\left\{
{\hat O}^i_{yz} \cdot{\hat O}^i_{yz} + {\hat O}^i_{xz} \cdot{\hat O}^i_{xz}
+ {\hat O}^i_{xy} \cdot{\hat O}^i_{xy} \right\},
\nonumber \\
\label{eqn:Hint}
\end{eqnarray}
where $g(\w_l)=g \cdot \w_{\rm D}^2/(\w_l^2+\w_{\rm D}^2)$, and
$g$ is the phonon-mediated el-el interaction at zero frequency;
$g=0.34$eV in the present point-charge model \cite{Kontani,Saito}.
\begin{figure}[!htb]
\includegraphics[width=0.8\linewidth]{fig3.eps}
\caption{
(Color online)
Inter-orbital scattering processes due to the
$e$-ph interaction by $u_\mu$ ($\mu=x,y,z$)
within the $t_{2g}$-orbitals ($l=2\sim4$).
}
\label{fig:e-ph}
\end{figure}
\section{Random-phase-approximation
\label{sec:RPA}}
Now, we explain the RPA for the five-orbital HH model \cite{Takimoto}.
The irreducible susceptibility in the five orbital model is given by
\begin{equation}
\chi^0_{ll',mm'} \left( q \right) = - \frac{T}{N} \sum_\k G_{lm}^0
\left( k+ q \right) G_{m' l'}^0 \left( k \right),
\end{equation}
where $\hat{G}^0 ( k ) = [ i \epsilon_n + \mu - \hat{H}^0_{\bm{k}} ]^{-1}$
is the \textit{d} electron Green function in the orbital basis,
$q = ( \bm{q}, \omega_l )$, $k=( \bm{k} , \epsilon_n)$,
and $\epsilon_n = (2n + 1) \pi T$ is the fermion Matsubara frequency.
$\mu$ is the chemical potential, and $\hat{H}^0_{\bm{k}}$ is the kinetic term.
Then, the susceptibilities for spin and charge sectors in the RPA are
given as \cite{Takimoto}
\begin{gather}
\hat{\chi}^{s} \left( q \right) = \frac{\hat{\chi}^0 \left( q \right)}{1 - \hat{\Gamma}^{s} \hat{\chi}^0 \left( q \right)},
\label{eqn:chis}\\
\hat{\chi}^{c} \left( q \right) = \frac{\hat{\chi}^0 \left( q \right)}{1 - \hat{\Gamma}^{c} (\omega_l) \hat{\chi}^0 \left( q \right)},
\end{gather}
where the bare four-point vertices ${\hat \Gamma}^{s,c}$ are
\begin{eqnarray}
\Gamma_{l_{1}l_{2},l_{3}l_{4}}^s = \begin{cases}
U, & l_1=l_2=l_3=l_4 \\
U' , & l_1=l_3 \neq l_2=l_4 \\
J, & l_1=l_2 \neq l_3=l_4 \\
J' , & l_1=l_4 \neq l_2=l_3
\end{cases}
\label{eqn:Gamma-s}
\end{eqnarray}
\begin{equation}
\hat{\Gamma}^c ( \omega_l )= -\hat{C} - 2\hat{V}_{\rm el-el}^{\rm ph}( \w_l ),
\label{eqn:Gc}
\end{equation}
\begin{eqnarray}
C_{l_{1}l_{2},l_{3}l_{4}} = \begin{cases}
U, & l_1=l_2=l_3=l_4 \\
-U'+2J , & l_1=l_3 \neq l_2=l_4 \\
2U' - J, & l_1=l_2 \neq l_3=l_4 \\
J' , & l_1=l_4 \neq l_2=l_3
\end{cases}
\label{eqn:Gamma-c}
\end{eqnarray}
In Eq. (\ref{eqn:Gc}), $(V_{\rm el-el}^{\rm ph})_{l_1,l_2,l_3,l_4}
=-g(\w_l)\sum_\Gamma^{xz,yz,yx}o^{l_1,l_2}_\Gamma o^{l_3,l_4}_\Gamma$.
Here, we neglect the ladder-diagram for phonon-mediated interaction
because of the relation $\omega_D \ll W_{\mathrm{band}}$ \cite{Kontani}.
In the RPA,
the enhancement of the spin susceptibility $\hat{\chi}^{s}$
is mainly caused by the intra-orbital Coulomb interaction $U$,
using the ``intra-orbital nesting''.
On the other hand,
the enhancement of $\hat{\chi}^{c}$ in the present model is caused by
the phonon-induced quadrupole-quadrupole interaction in Eq. (\ref{eqn:Hint}),
utilizing the ``inter-orbital nesting'' in the present model.
The SDW (ODW) state is realized when the spin (charge) Stoner factor $\a_{s(c)}$,
which is the maximum eigenvalue of $\hat{\Gamma}^{s(c)} \hat{\chi}^0(\bm{q},0)$,
is unity.
When $n=6.05$,
the critical value of $U$ is $U_c=1.26$ eV, and
the critical value of $g$ (at $U=0$) is $g_c=0.233$ eV.
Smallness of $g_{\rm c}$ in iron pnictides originates from
the better inter-orbital nesting.
Hereafter, we set the unit of energy as eV unless otherwise noted.
Here, we introduce the diagonal charge quadrupole
susceptibilities in the five-orbital model as
\begin{eqnarray}
\chi_{\Gamma}^{Q}(q)&=&
\sum_{ll'}\sum_{mm'}o_\Gamma^{ll'}\chi_{ll'mm'}^c(q)o_\Gamma^{mm'},
\label{eqn:chiQ}
\end{eqnarray}
for $\Gamma=xz,yz,xy$.
Their momentum dependence at zero frequency
is shown in Fig. \ref{fig:chic} for $\a_c=0.98$.
More generally, the quadrupole susceptibility is defined as
\begin{eqnarray}
\chi_{\Gamma,\Gamma'}^{Q}(q)=
\sum_{ll'}\sum_{mm'}o_\Gamma^{ll'}\chi_{ll'mm'}^c(q)o_{\Gamma'}^{mm'} .
\label{eqn:offdiagonal}
\end{eqnarray}
However, its off-diagonal terms with $\Gamma\ne\Gamma'$
are negligibly small in the present model in the ``$xyz$-coordinate''.
In this approximation, in the absence of Coulomb interaction,
the quadrupole susceptibility in the RPA is given as
\begin{eqnarray}
\chi_{\Gamma}^{Q}(q)\approx
\chi_{\Gamma}^{Q,0}(q)/(1-2g\chi_{\Gamma}^{Q,0}(q))
\label{eqn:chiGU0} ,
\end{eqnarray}
where $\chi_{\Gamma}^{Q,0}$ is the irreducible quadrupole susceptibility.
Considering the fact that $\chi_{ll',mm'}^0(q)$ takes large value for
$l=m$ and $l'=m'$, we obtain $\chi_{xz}^{Q,0}(q)\approx
2\chi_{25,25}^0(q)+2\chi_{34,34}^0(q)+(2/3)\chi_{12,12}^0(q)$.
Since $\chi_{xz}^{Q,0}(q)\approx 2.5$ at $\q=(\pi,0)$,
the critical value of $g$ is $g_c\sim0.2$ in the present model.
We stress that the relations
$\chi_{\Gamma,\Gamma'}^{Q}(q)=\chi_{\Gamma}^{Q}(q)\delta_{\Gamma,\Gamma'}$
and Eq. (\ref{eqn:chiGU0}) holds exactly for $q=({\bm 0},\w_l)$.
We utilize this relation in calculating the shear modulus.
As for the contributions by $t_{2g}$-orbitals ($l=2\sim4$),
$\chi_{xz}^{Q}(q)\propto \chi_{34,34}^c(q)$,
$\chi_{yz}^{Q}(q)\propto \chi_{24,24}^c(q)$, and
$\chi_{xy}^{Q}(q)\propto \chi_{23,23}^c(q)$.
In Fig. \ref{fig:chic} (a), $\chi_{xz}^{Q}(\q)$ has the highest peak
at $\q=(\pi,0)$, which is given by the
inter-orbital nesting between orbital 3 on FS $\a_2$
and orbital 4 on FS $\b_1$ in the five-orbital model
in Fig. \ref{fig:fig1} (a).
Also, $\chi_{yz}^Q(\q)$ has the highest peak at $\q=(0,\pi)$
in Fig. \ref{fig:chic} (b),
due to the inter-orbital nesting between orbital 2 on FS $\a_2$
and the orbital 4 on FS $\b_2$.
We will see that $\chi_{xz(yz)}^Q(\q)$ is modified by the unfolding procedure.
Also, $\chi_{xy}^Q(\q)$ in Fig. \ref{fig:chic} (c)
is given by the inter-orbital nesting between orbital 2 and 3,
due to the out-of-plane oscillations of Fe-ions.
The inter-band and intra-band scattering processes produce the
enhancement of $\chi_{xy}^Q(\q)$ at $\q=(\pi,0), (0,\pi)$
and $\q={\bm 0}$, respectively.
We note that $\chi_{xy}^Q(\q)$ is not affected by the unfolding procedure.
\begin{figure}[!htb]
\includegraphics[width=0.8\linewidth]{fig4.eps}
\caption{
(Color online)
Quadrupole susceptibilities in the five-orbital model for
(a) $\chi_{xz}^{Q}(\q)$, (b) $\chi_{yz}^{Q}(\q)$, and
(c) $\chi_{xy}^{Q}(\q)$, respectively.
The used model parameters are $n=6.05$, $T=0.05$, and $g=0.22$
($\a_c=0.98$).
The correlation length in (a) or (b) is derived as
$\xi=\pi/\Delta q\sim3$, where $\Delta q$ is the half-width of the peak.
Therefore, we obtain the relations $6\xi^2\sim (1-\a_c)^{-1}$
and $c\xi^2\sim 2(1-\a_c)^{-1}\sim 12\xi^2$.
}
\label{fig:chic}
\end{figure}
Figure \ref{fig:chic-ac} shows both
$\chi_{xz}^Q(\q)$ at $\q=(\pi,0)$ and $\chi_{xy}^Q({\bm 0})$
as a function of $\a_c$ for $n=6.0$ and $n=6.05$ given by the RPA.
We see that $\chi_{xz}^Q(\q)$ develops divergently
in proportion to $(1-\a_c)^{-1}\propto (g_c-g)^{-1}$,
while $\chi_{xy}^Q({\bm 0})$ shows an enhanced but saturated value
even at $g=g_{\rm c}$.
\begin{figure}[!htb]
\includegraphics[width=0.8\linewidth]{fig5.eps}
\caption{
(Color online)
Quadrupole susceptibilities as a function of $\a_c$
for $n=6.0$ and $n=6.05$ given by RPA.
Model parameters are $U=0.8$ and $T=0.05$.
We can recognize the relation
$\chi_{xz}^Q({\bm Q})\propto (1-\a_c)^{-1} \propto (g_{\rm c}-g)^{-1}$,
which diverges at $\a_c=1$ or $g=g_{\rm c}$.
}
\label{fig:chic-ac}
\end{figure}
In Appendix \ref{sec:10},
we will calculate $\chi_{\Gamma}^{Q}(q)$ in the ten-orbital model,
and make comparison to Fig. \ref{fig:chic} in the five-orbital model:
Although both results coincide for $\Gamma=xy,z^2,x^2-y^2$,
they are different for $\Gamma=xz,yz$.
The reason is that the signs of $O_{xz/yz}$ at Fe-B sites are changed
by applying the ``unfold-gauge transformation''.
As we will explain in Appendix \ref{sec:10}, the development of
$\chi_{xz/yz}^{Q}({\bm 0},0)$ in Fig. \ref{fig:chic} is the artifact
of the unfold-gauge transformation.
For this reason,
the correct AFQ susceptibility in the ten-orbital model is given as
$\chi_{xz/yz}^{Q}(\q)=\chi_{xz/yz}^{Q,\rm{5-orbital}}(\q+(\pi,\pi))$.
The optical modes that give the enhancements of
$\chi_{yz}^{Q}(\q)$ at $\q=(\pi,0)$ and $(\pi,\pi)$
in the ten-orbital model
($\q=(0,\pi)$ and $(0,0)$ in the five-orbital model)
are caused by the in-plane $u_{x}$
oscillations shown in Fig.\ref{fig:optical} (a).
Also, the enhancements of $\chi_{xy}^{Q}(\q)$ at $\q=(\pi,0)$
and $(0,0)$ are caused by the out-of-plane $u_z$
oscillations in Fig.\ref{fig:optical} (b).
\begin{figure}[!htb]
\includegraphics[width=0.88\linewidth]{fig6.eps}
\caption{
(Color online)
(a) Fe-ion in-plane optical phonons
with momentum $\q=(\pi,0)$ and $\q=(\pi,\pi)$.
(b) Fe-ion out-of-plane optical phonons
with momentum $\q=(\pi,0)$ and $\q=(0,0)$.
}
\label{fig:optical}
\end{figure}
\section{Acoustic phonons
\label{sec:acoustic}}
In previous sections, we studied the $e$-ph interaction due to
optical phonons, and calculated the quadrupole susceptibilities by the RPA.
To obtain the shear modulus,
we also need the knowledge on the $e$-ph interaction due to
acoustic phonons with momentum $\q\approx0$.
In fact, shear modulus is proportional to the square of
the acoustic phonon velocity, which is renormalized
by the electron-acoustic phonon interaction in the presence of strong
quadrupole fluctuations.
In this section, we derive the $e$-ph interaction due to
acoustic phonons with $\q\approx0$.
Hereafter, we use the unit $\hbar=1$, and take
the nearest-neighbor Fe-Fe distance $a_{\rm Fe-Fe}$ as the unit of length.
Figure \ref{fig:accoustic} shows the transverse acoustic modes that are
related to (a) $C_{44}$, (b) $C_E=(C_{11}-C_{12})/2$, and (c) $C_{66}$
\cite{Yoshizawa}.
Now, we calculate the $e$-ph interaction based on the point-charge model,
by following the procedure in Ref. \onlinecite{Kontani} for the optical phonons.
The quadrupole potential energies at Fe-site caused by the
transverse acoustic phonons in Fig. \ref{fig:accoustic} are given as
\begin{eqnarray}
V_{44}&=& -\frac{3e^2}{R_{\rm Fe-As}^4}\frac{8}{\sqrt{3}} xz\cdot
{\tilde u}_{44}
\label{eqn:V44} ,\\
V_{E}&=& -\frac{3e^2}{R_{\rm Fe-As}^4}\frac{8}{\sqrt{3}} xy\cdot
{\tilde u}_{E}
\label{eqn:VE} ,\\
V_{66}&=& \frac{3e^2}{R_{\rm Fe-As}^4}\sqrt{6} (x^2-y^2)\cdot
{\tilde u}_{66}
\label{eqn:V66} ,
\end{eqnarray}
for both Fe-A and Fe-B sites, where
$(x,y,z)$ is the coordinates of $d$-electron.
${\tilde {\bm u}}_{\phi}\equiv {\bm u}_{\phi}-{\bm u}_{\rm Fe}$ ($\phi=44,66,E$)
is the relative displacements
of the nearest As ions from the center Fe ion;
$u_{\phi}$ ($u_{\rm Fe}$) is the displacement vector of
the As- (Fe-) ion we are considering from the original position.
Note that the shear strain tensors are given as
$\e_{44(E)}={\tilde u}_{44(E)}/(a_{\rm Fe-Fe}/2)=2{\tilde u}_{44(E)}$ and
$\e_{66}={\tilde u}_{66}/(a_{\rm Fe-Fe}/\sqrt{2})=\sqrt{2}{\tilde u}_{66}$.
The corresponding operators in the ten-orbital model
are respectively given as
\begin{eqnarray}
{\hat V}_{44}&=& -\frac{3e^2a_d^2}{R_{\rm Fe-As}^4}\frac{8}{7\sqrt{3}}
{\hat O}_{xz}\cdot {\tilde u}_{44}
\label{eqn:V44O} ,\\
{\hat V}_{E}&=& -\frac{3e^2a_d^2}{R_{\rm Fe-As}^4}\frac{8}{7\sqrt{3}}
{\hat O}_{xy}\cdot {\tilde u}_{E}
\label{eqn:VEO} ,\\
{\hat V}_{66}&=& \frac{3e^2a_d^2}{R_{\rm Fe-As}^4}\frac{2\sqrt{6}}{7}
{\hat O}_{x^2-y^2}\cdot {\tilde u}_{66}
\label{eqn:V66O} .
\end{eqnarray}
Therefore, the acoustic modes in Fig. \ref{fig:accoustic} (a)-(c)
couple with the quadrupole susceptibilities at $\q\approx0$;
$\chi_{xz}^Q(0)$, $\chi_{xy}^Q(0)$, and
$\chi_{x^2-y^2}^Q(0)$ for $\phi=44$ $E$, and $66$
in the ten-orbital model, respectively.
To study the softening in the five-orbital models,
we have to perform the ``unfold-gauge transformation''
for Eqs. (\ref{eqn:V44O})-(\ref{eqn:V66O}).
Under the gauge transformation,
Eqs. (\ref{eqn:VEO}) and (\ref{eqn:V66O}) are invariant,
while Eq. (\ref{eqn:V44O}) is changed to
\begin{eqnarray}
{\hat V}'_{44}&=& \mp \frac{3e^2a_d^2}{R_{\rm Fe-As}^4}\frac{8}{7\sqrt{3}}
{\hat O}_{xz}\cdot {\tilde u}_{44} ,
\label{eqn:V44O-2}
\end{eqnarray}
where the $-(+)$ sign corresponds to Fe-A (Fe-B) site.
In the ``five-orbital model'', therefore,
the softening of $C_{E}$ and $C_{66}$ are caused by
$\chi_{xy}^Q({\bm 0},0)$ and $\chi_{x^2-y^2}^Q({\bm 0},0)$, respectively,
while the softening of $C_{44}$ is caused by $\chi_{xz}^Q((\pi,\pi),0)$.
Therefore, the softening in shear modulus ($C_{66}$ and $C_{44}$)
does not occur within the RPA \cite{comment2}.
\begin{figure}[!htb]
\includegraphics[width=0.9\linewidth]{fig7.eps}
\caption{
(Color online)
Displacement vectors ${\bm u}_{\rm As}$ and ${\bm u}_{\rm Fe}$
in the transverse acoustic modes
that couple with (a) $C_{44}$, (b) $C_E$, and (c) $C_{66}$.
$C_{66}$ mode corresponds to the orthorhombic structure transition.
}
\label{fig:accoustic}
\end{figure}
Next, we derive the effective el-el interaction due to $\k\rightarrow0$
transverse acoustic modes.
In the case of $\k=k\cdot(1,1)/\sqrt{2}$ and $k\ll1$
shown in Fig. \ref{fig:accoustic} (c),
the displacement operator for the As site at ${\bm R}_s$ is
\begin{eqnarray}
u_s= \sum_\k \sqrt{\frac{1}{2NM\w_\k}}
\left[ a_\k e^{i\k{\bm R}_s}+a_\k^\dagger e^{-i\k{\bm R}_s} \right] ,
\end{eqnarray}
where
$a_\k$ and $a_\k^\dagger$ satisfy the commutation relation
$[a_\k,a_{\k'}^\dagger]=\delta_{\k,\k'}$, $M$ is the mass of As-ion, and
$\w_\k=v_\k |\k|$; $v_\k$ is the bare acoustic phonon velocity.
The Fourier transformation of $u_s$ is given as
$u_\k= \sqrt{\frac{1}{2M\w_\k}}(a_\k+a_{-\k}^\dagger)$.
Then, the Fourier transformation of the acoustic-phonon Green function
$D_\k(\tau)\equiv \langle T_{\tau} u_\k^i ( \tau ) u_\k^i ( 0 ) \rangle$
is
\begin{eqnarray}
D_\k(\w_n)=\frac{2\w_\k}{\w_n^2+\w_\k^2}\langle{u}_\k^2\rangle_0 ,
\label{eqn:Dq}
\end{eqnarray}
where $\langle {u}_\k^2 \rangle_0 =(1/2M\w_\k)$.
As understood in Fig. \ref{fig:accoustic} (c),
the relative displacement with the origin at the Fe-ion,
${\tilde u}_s$, is given as $\frac12 (u_{s'}-u_s)$, where
${\bm R}_{s'}=(0.5,0.5)$ and ${\bm R}_{s}=(-0.5,-0.5)$
with the origin at the Fe-ion.
Considering that $({\bm R}_{s'}-{\bm R}_{s})\cdot \k=\sqrt{2}k$,
its Fourier transformation is given as
\begin{eqnarray}
{\tilde u}_\k\equiv \sum_s {\tilde u}_s e^{-i \k{\bm R}_s}
\sim \frac{i k}{\sqrt{2}}u_\k .
\end{eqnarray}
To calculate the shear modulus,
we need the quadrupole susceptibility in the ``$k$-limit'',
in which we put $\w=0$ first, and take the limit $k\rightarrow0$ later
\cite{Kontani-VV}.
For this purpose, we derive the effective el-el interaction
due to the transverse acoustic phonon in the $k$-limit.
The el-el interaction due to the phonon in Fig. \ref{fig:accoustic} (c)
is given by the second-order term of Eq. (\ref{eqn:V66O}).
Using the relation $D_\k(0)\k^2=2(1/2M\w_\k)\k^2/\w_\k=1/Mv_\k^2$,
it is given as
\begin{eqnarray}
H_{66}&=&-g_{66}\sum_{\k\k',ll'mm',\s\s'}o_{x^2-y^2}^{ll'}o_{x^2-y^2}^{mm'}
\nonumber\\
& &\times c_{l\k\s}^\dagger c_{l'\k\s}c_{m\k'\s'}^\dagger c_{m'\k'\s'},
\label{eqn:H66} \\
g_{66}&=& \frac{B^2}{R_{Fe-As}^2}\frac{1}{Mv_\k^2},
\end{eqnarray}
where $\displaystyle B\equiv \frac{3e^2}{R_{\rm Fe-As}}
\left(\frac{a_d}{R_{\rm Fe-As}}\right)^2\frac{2\sqrt{3}}{7}$;
$B=0.95$ eV for $a_d=0.77$ ${\buildrel _{\circ} \over {\mathrm{A}}}$
(=Shannon crystal radius of Fe$^{2+}$)
and $R_{\rm Fe-As}=2.4$ ${\buildrel _{\circ} \over {\mathrm{A}}}$.
Therefore,
\begin{eqnarray}
g_{66}= \eta_{66}^2 C_{66,0}^{-1} ,
\end{eqnarray}
where $C_{66,0}\equiv Mv_\k^2$ is the bare shear modulus,
and $\eta_{66}= BR_{\rm Fe-As}^{-1}$ is the quadrupole-strain coupling constant.
If we put $v_\k\sim0.024$ eV ${\buildrel _{\circ} \over {\mathrm{A}}}$
($v_\k\sim0.018$ eV ${\buildrel _{\circ} \over {\mathrm{A}}}$)
according to the first principle study \cite{Boeri},
we obtain $g_{66}=0.12$eV ($g_{66}=0.21$eV).
On the other hand, we obtain $g=0.34$eV for the Fe-ion
optical phonons with $\w_{\rm D}=0.02$eV \cite{Saito}.
Thus, $g_{66}/g=1/2\sim1/3$ in the present point charge model.
In the same way, we also derive the el-el interactions due to the
acoustic phonon with $\k=k\cdot(0,1)$, shown in Fig. \ref{fig:accoustic} (b).
For this mode, the relative displacement ${\tilde u}_\k$ is given as
${\tilde u}_s\equiv \frac12 (u_{s'}-u_s)$
with ${\bm R}_{s'}=(0.5,0.5)$ and ${\bm R}_{s}=(0.5,-0.5)$.
Since $({\bm R}_{s'}-{\bm R}_{s})\cdot\k=k$,
its Fourier transformation is given by
\begin{eqnarray}
{\bar u}_\k\equiv \sum_s {\tilde u}_s e^{-i \k{\bm R}_s}
\sim \frac{i k}{2} u_\k .
\end{eqnarray}
Then, the phonon-mediated el-el interactions
are given by the second-order terms of Eq. (\ref{eqn:VEO}).
As a result,
the el-el interactions due to phonons in Fig. \ref{fig:accoustic} (b)
is given as
\begin{eqnarray}
H_{E}&=&-g_{E}\sum_{\k\k',ll'mm',\s\s'}o_{xy}^{ll'}o_{xy}^{mm'}
\nonumber\\
& &\times c_{l\k\s}^\dagger c_{l'\k\s}c_{m\k'\s'}^\dagger c_{m'\k'\s'},
\end{eqnarray}
where
$\displaystyle g_{E}=\frac{B'^2}{R_{Fe-As}^2}\frac{1}{Mv_\k^2}$,
and $\displaystyle B'\equiv \frac{3e^2}{R_{\rm Fe-As}}
\left(\frac{a_d}{R_{\rm Fe-As}}\right)^2\frac{4}{7\sqrt{3}}$.
In the same way, we obtain $g_{44}=g_{E}$.
Therefore, $g_{\phi}=\eta_{\phi}^2 C_{\phi,0}^{-1}$
and $\eta_{\phi}^2=0.44\eta_{66}^2$ for $\phi=44, E$.
In conclusion, $g_{E}=g_{44}=0.44g_{66}$
if $v_\k$ is equivalent for all modes.
\section{Softening of shear moduli
\label{sec:shear}}
\subsection{Softening due to one-orbiton process; the RPA}
Here, we calculate the shear modulus given by the one-orbiton process
using the RPA.
For this purpose, we introduce the following shear modulus susceptibilities
in the five-orbital model,
in the absence of $e$-ph interaction due to $\q\approx0$ acoustic phonon:
\begin{eqnarray}
\chi_E&=& 2\chi_{xy}^{Q}({\bm 0},0),
\label{eqn:chiE-def} \\
\chi_{44}&=& 2\chi_{xz}^{Q}((\pi,\pi),0),
\label{eqn:chi44-def} \\
\chi_{66}&=& 2\chi_{x^2-y^2}^{Q}({\bm 0},0),
\label{eqn:chi66-def}
\end{eqnarray}
where the factor 2 comes from the spin degeneracy.
They are schematically depicted in Fig. \ref{fig:AL} (a).
Note that $\chi_{44}= 2\chi_{xz}^{Q}({\bm 0},0)$ in the ten-orbital model.
According to Sec. 2 in Ref. \onlinecite{Thal},
the shear modulus is given by the second derivative
of the Free energy with respect to the shear strain tensor:
The expression for the shear modulus $C_\phi$ ($\phi=E,44,66$) is
\cite{Thal,Levy}
\begin{eqnarray}
C_{\phi}=C_{\phi,0} - \eta_\phi^2\chi_\phi ,
\label{eqn:Cx}
\end{eqnarray}
where $C_{\phi,0}=v_\phi^2\rho$ is the bare shear modulus,
where $v_\phi$ is the bare acoustic phonon velocity
and $\rho$ is the mass density.
$\eta_\phi$ is the quadrupole-strain coupling constant
due to the ``acoustic phonon'' given in Sec. \ref{sec:acoustic}.
In Eq. (\ref{eqn:Cx}), the condition for the structure transition,
$C_\phi=0$, is satisfied when $\chi_\phi= g_\phi^{-1}(\gg1)$.
That is, the structure transition
occurs prior to the divergence of $\chi_\phi$.
We can rewrite the expression for $C_{\phi}$ given in Eq. (\ref{eqn:Cx})
as follows:
\begin{eqnarray}
C_{\phi}^{-1}&=&C_{\phi,0}^{-1}[1 + g_\phi {\tilde \chi}_\phi]
\label{eqn:Cphi} ,\\
{\tilde \chi}_\phi&=& \chi_\phi/(1-g_\phi\chi_\phi)
\label{eqn:Cx2} ,
\end{eqnarray}
where $g_\phi\equiv \eta_\phi^2 C_{\phi,0}^{-1}$ is the effective el-el interaction
due to acoustic phonon given in the previous section.
In Eq. (\ref{eqn:Cphi}), the condition $C_\phi=0$
corresponds to the divergence of ${\tilde \chi}_\phi$,
since ${\tilde \chi}_\phi$ is the {\it total susceptibility} including the
$e$-ph interactions due to acoustic phonons.
If we put $U=0$ for simplicity, Eq. (\ref{eqn:Cx2}) is expressed as
\begin{eqnarray}
{\tilde \chi}_{44}&=& \chi_{44}^{0}/(1-(g+g_{44})\chi_{44}^{0}),
\label{eqn:chi44U0}\\
{\tilde \chi}_{E}&=& \chi_{E}^{0}/(1-(g+g_{E})\chi_{E}^{0}),
\label{eqn:chiEU0}\\
{\tilde \chi}_{66}&=& \chi_{66}^{0}/(1-g_{66}\chi_{66}^{0}),
\label{eqn:chi66U0}
\end{eqnarray}
where the suffix $0$ represents the bare susceptibility.
According to ${\tilde \chi}_{44}$ in Eq. (\ref{eqn:chi44U0}),
$g$ in Eq. (\ref{eqn:chiGU0}) for $\chi_{xz}^Q$ is replaced with $g+g_{44}$
when both optical and acoustic phonons are taken into account.
Therefore, we have to reduce $g$ to $g-g_{44}$
to keep the charge Stoner factor $\a_c$ and $\chi_{xz}^Q(\Q)$ unchanged.
Considering the relation $g_{44}\sim g_{E}$ that we derived
in the previous section, we obtain ${\tilde \chi}_{E}\sim \chi_{E}$.
Then, we conclude that (i) $C_{44}\sim C_{44,0}$ since
$\chi_{44}$ is seldom enhanced in the RPA.
Also, (ii) $C_{E}$ softens to some extent
since $\chi_{xy}^Q({\bm 0})$ is weakly enhanced
as shown in Fig. \ref{fig:chic-ac}, although
the relation $C_{E}=0$ will not be satisfied
because of the relation ${\tilde \chi}_{E}\sim \chi_{E}$.
As for $C_{66}$ given in Eq.(\ref{eqn:chi66U0}),
$\chi_{66}^{0}$ in the present model is $\sim 2 {\rm eV}^{-1}$,
while we estimate $g_{66}=0.1\sim0.2{\rm eV}$.
Therefore, we expect (iii) $C_{66}\sim C_{66,0}$ in the RPA,
which is inconsistent with experimentally observed
large softening in $C_{66}$ \cite{Yoshizawa}.
In the RPA, the softening in shear modulus ($C_{66}$, $C_{44}$ and $C_E$)
is small according to Eqs. (\ref{eqn:chiE-def})-(\ref{eqn:chi66-def})
and Figs. \ref{fig:chic} (a)-(c).
In the next subsection, we analyze $\chi_{66}$
by taking account of the two-orbiton process
that is not included in the RPA.
\begin{figure}[!htb]
\includegraphics[width=0.9\linewidth]{fig8.eps}
\caption{
(Color online)
(a) Diagrammatic expression for the shear modulus susceptibilities
$\chi_{\phi}$ ($\phi=44,66,E$) in the RPA.
(b) The second-order term with respect to the third-order anharmonic
phonon-phonon interaction $A_{(3)}u_x^2u_{66}$.
This diagram represents the virtual process in which
an acoustic phonon with $\q=0$ breaks into
two optical phonons conserving the total momentum.
(c) Two-orbiton term for the shear modulus susceptibilities
$\chi_{66}^{\rm TO}$, which is the irreducible susceptibility of
$\chi_{x^2-y^2}({\bm 0},0)$.
This term gives the softening of $C_{66}$.
(d) Dominant contribution for the three-point vertex
$A_{x^2-y^2}(\Gamma,\Gamma;\q)$ for $\Gamma=xz$ and $yz$.
(e) Self-energy correction due to $\chi^Q_{xz}(q)$.
(f) Second-order Maki-Thompson-type vertex correction
with respect to $\chi^Q_{xz}(q)$.
}
\label{fig:AL}
\end{figure}
\subsection{Softening due to two-orbiton process;
the Aslamazov-Larkin-type diagram
\label{sec:AL}}
In the previous subsection,
we studied the softening of shear moduli within the RPA.
However, the obtained softening is very small
since only the AFQ fluctuations develop in the present model.
In this subsection, we analyze $\chi_{66}$
by taking account of the ``two-orbiton process''
that is not included in the RPA.
In usual cases, this higher order process is negligible.
However, it gives divergent increase in $\chi_{66}$,
and creates the FQ-QCP near the AFQ-QCP.
For this reason, the orthorhombic structure transition ($C_{66}=0$)
is induced by the two-orbiton process.
Before calculating the two-orbiton process,
let us consider the third order anharmonic phonon-phonon
coupling $\sim A_{(3)}u_x^2u_{66}$, where $u_{x}$ is the
displacement of Fe-ion optical mode in Eq. (\ref{eqn:e-ph}),
and $u_{66}$ is the displacement of As-ion acoustic node in
Eq. (\ref{eqn:V66}) or (\ref{eqn:V66O}).
Figure \ref{fig:AL} (b) shows the
second-order-term with respect to $A_{(3)}$,
where $D(q)$ is the optical phonon Green function in Eq. (\ref{eqn:D}),
and the factor $1/2$ is introduced to cancel the overcounting
with respect to the upside-down diagrams.
This term gives a self-energy correction for the acoustic phonon
Green function.
This two phonon process would be measurable in Raman spectroscopy.
By considering the $e$-ph interaction,
each $D(q)$ in (b) is replaced with $D(q)+(\eta D(q))^2\cdot\chi_{yz}^Q(q)$.
Here, we study the ferro-quadrupole susceptibility due to
the ``two-orbiton term'' given by Fig. \ref{fig:AL} (c),
and analyze for the softening of $C_{66}$.
Since the coefficient of the anharmonic phonon-phonon coupling $A_{(3)}$
would be small in iron pnictides,
we instead consider the three-point vertex given by
electron Green functions in Fig. \ref{fig:AL} (d).
The two-orbiton term in Fig. \ref{fig:AL} (c) is similar to the
Aslamazov-Larkin (AL) term for the excess conductivities
due to superconducting fluctuations;
such as longitudinal \cite{AL} and Hall \cite{Fuku} conductivities,
and the Nernst coefficient \cite{Uss}.
Figure \ref{fig:AL} (c) is a virtual process in which
one $O_{x^2-y^2}$-type orbiton with $\q=0$ breaks into
two $O_{xz}$-type orbitons with zero total momentum ($\pm \Q$).
After taking the momentum summation, the two-orbiton term (c)
in 2D systems would be strongly enhanced for $\q=0$ since
$\chi_{yz}^Q(\k)$ has a large peak at finite momentum $\k$.
The mathematical expression for the two-orbiton process in
Fig. \ref{fig:AL} (c) for $\chi_{66}$, which we call
$\chi_{66}^{\rm TO}$, is given as
\begin{eqnarray}
\chi_{66}^{\rm TO}&=& \frac12 (2g)^{2}T\sum_{\q,l,\Gamma,\Gamma'}
A_{x^2-y^2}(\Gamma,\Gamma';\q,\w_l)A_{x^2-y^2}(\Gamma,\Gamma';\q,\w_l)
\nonumber\\
& &\times (1+2g\chi_{\Gamma}^{Q}(\q,\w_l))(1+2g\chi_{\Gamma'}^{Q}(\q,\w_l))
\nonumber\\
&\approx& \frac12(2g)^{4}T\sum_{\q,l,\Gamma,\Gamma'}
A_{x^2-y^2}(\Gamma,\Gamma';\q,\w_l)A_{x^2-y^2}(\Gamma,\Gamma';\q,\w_l)
\nonumber\\
& &\times \chi_{\Gamma}^{Q}(\q,\w_l)\chi_{\Gamma'}^{Q}(\q,\w_l)
\label{eqn:chi66} ,
\end{eqnarray}
where $A_{x^2-y^2}(\Gamma,\Gamma';\q,\w_l)$
is the three point vertex with respect to ${\hat V}_{66}$
in Eq. (\ref{eqn:V66O}) and quadrupole operators
${\hat O}_{\Gamma}$ and ${\hat O}_{\Gamma'}$, shown in Fig. \ref{fig:AL} (d).
We used the relation $\chi_\Gamma^Q(q)=\chi_\Gamma^Q(-q)$.
When $U=0$, $A_{x^2-y^2}$ for $\w_l=0$ is given as
\begin{eqnarray}
A_{x^2-y^2}(\Gamma,\Gamma';\q) &=& -2T\sum_{n,\k}{\rm Tr}\left\{
{\hat G}_\k(\e_n){\hat o}_{x^2-y^2} {\hat G}_\k(\e_n) \right.
\nonumber \\
& &\left. \times {\hat o}_\Gamma{\hat G}_{\k+\q}(\e_n){\hat o}_{\Gamma'} \right\}
\label{eqn:Aw0} ,
\end{eqnarray}
where the factor 2 in front of Eq. (\ref{eqn:Aw0})
accounts for the diagrams with reversing three Green functions
${\hat G}_\k(\e_n)\rightarrow {\hat G}_{-\k}(-\e_n)$.
Near the QCP $g\lesssim g_c$,
the most divergent quadrupole susceptibility is $\chi_{xz(yz)}^{Q}$.
Therefore, the dominant contribution for $\chi_{66}^{\rm TO}$
in Eq. (\ref{eqn:chi66}) will be given by the term
with $\Gamma=\Gamma'=xz$ or $yz$.
After the analytic continuation, the functional form of
$\chi_{xz/yz}^{Q}(\q,\w)$ for $\q\approx {\bm Q}_{xz}=(\pi,0)$
or $\q\approx {\bm Q}_{yz}=(0,\pi)$ would be approximately given as
\begin{eqnarray}
\chi_\Gamma^{Q}(\q,\w+i\delta)&=&\frac{c\xi^2}
{1+ \xi^2(\q-\Q_\Gamma)^2-i\w/\w_0},
\label{eqn:chi-analytic}
\end{eqnarray}
for $\Gamma=xz$ or $yz$,
where $\xi$ is the correlation length
and $\w_0$ is the characteristic energy of the fluctuation.
The relation $\xi^2\propto \w_0^{-1}$ holds in the RPA.
Next, we consider the temperature dependence of $\xi$.
In the FLEX approximation \cite{Onari} or SCR theory \cite{Moriya},
the bare susceptibility $\chi_\phi^0$ is approximately suppressed as
$\chi_\phi^0-\a T$ ($\a>0$)
due to the thermal fluctuations, which are described as
the self-energy and Maki-Thompson vertex corrections.
In this case, we obtain
$\chi_{xz}^{Q}({\bm Q},0)\propto (1-g\chi_{xz}^0({\bm Q},0)+g\a T)^{-1}
\propto (T-T_{\rm AFQ})^{-1}$ based on the RPA,
where $T_{\rm AFQ}=-(1-g\chi_{xz}^0({\bm Q},0))/g\a$
is the transition temperature to the AFQ ordered state.
Since $\chi_{xz}^{Q}({\bm Q},0)\propto\xi^2$,
we assume the following relations
\begin{eqnarray}
\xi^2&=&l(T-T_{\rm AFQ})^{-1},
\\
\w_0&=&l'(T-T_{\rm AFQ}),
\label{eqn:w0}
\end{eqnarray}
where $l,l'$ are constants.
Note that $\w_0\xi^2$ is temperature independent \cite{Moriya}.
By carrier doping, $T_{\rm AFQ}$ changes from positive to negative, while
other model parameters ($c$, $l$, and $l'$) would be insensitive to doping.
As shown in Fig. \ref{fig:chic-ac},
$\chi_{xz}^{Q}(\Q,0)\sim 2.4\times(1-\a_c)^{-1}\sim 12\xi^2$.
In the case (i) $T_{\rm AFQ}>0$, the relation $\w_0<T$
is satisfied near $T_{\rm AFQ}$.
In the opposite case (ii) $T_{\rm AFQ}<0$, the relation $\w_0>T$
will hold for wide range of temperatures.
Note that the present phenomenological model
in Eqs. (\ref{eqn:chi-analytic})-(\ref{eqn:w0})
is reproduced by the microscopic calculation by FLEX approximation
\cite{Onari}.
As for the spin propagator in cuprate superconductors,
the relation $\w_0>T$ ($\w_0<T$) holds in over-doped (under-doped) systems.
Here, we comment on the self-energy correction
and the Maki-Thompson-type vertex correction for $\chi_{66}^{\rm TO}$,
shown in Figs. \ref{fig:AL} (e) and (f) respectively.
The former term is included in the FLEX approximation,
and it gives the Curie-Weiss behavior of $\chi^Q_{xz}({\bm Q},0)$
given by Eqs. (\ref{eqn:chi-analytic})-(\ref{eqn:w0}),
as reported in Ref. \onlinecite{Onari} or Ref. \onlinecite{Moriya}.
The latter term would be negligible since its temperature dependence
is smaller than that of the former term.
For this reason, we concentrate on the two-orbiton term
in Fig. \ref{fig:AL} (c) hereafter.
From now on, we perform the numerical calculation of the two-orbiton process
in the case (i), in which the relations $\xi\gg1$ and $\w_0\ll T$
are realized near the orbital-ordered state.
In this case, the dominant contribution in Eq. (\ref{eqn:chi66})
comes form the terms with $\Gamma=\Gamma'=xz$ and $yz$.
Also, we can safely apply the classical approximation,
in which the terms with $\w_l\ne0$ are dropped in Eq. (\ref{eqn:chi66}).
Under these approximations, Eq. (\ref{eqn:chi66}) is simplified as
\begin{eqnarray}
\chi_{66}^{\rm TO} &=& (2g)^{4}T\sum_\q
\{ A_{x^2-y^2}(xz,xz;\q) \chi_{xz}^{Q}(\q,0) \}^2 .
\label{eqn:chi66-highT}
\end{eqnarray}
To calculate $A_{x^2-y^2}(\Gamma,\Gamma';\q)$,
we introduce a uniform FQ potential term
$H'=h\sum_i {\hat O}_{x^2-y^2}^i$, where $h$ is an infinitesimal small constant.
Then, the three point vertex is given as
the following Ward identity \cite{AGD}:
\begin{eqnarray}
A_{x^2-y^2}(\Gamma,\Gamma';\q) =
\frac1h \left[ {\bar \chi}_{\Gamma,\Gamma'}^{Q}(\q,0;h)
-{\bar \chi}_{\Gamma,\Gamma'}^{Q}(\q,0;0) \right]
\label{eqn:Ax2y2-W} ,
\end{eqnarray}
where ${\bar \chi}_{\Gamma,\Gamma'}^{Q}(\q,\w_l;h)$ is the
``irreducible'' quadrupole susceptibility with respect to $g$.
In the numerical calculation, we have to fix $\mu$ against the change in $h$.
Equation (\ref{eqn:Ax2y2-W}) gives the correct
three-point vertex even for $U\ne0$.
In the case of $U=0$, ${\bar \chi}_{\Gamma,\Gamma'}^{Q}$ is simply
given as ${\bar \chi}_{\Gamma,\Gamma'}^{Q}(\q,\w_l;h)
=-T\sum_{\k,n}{\rm Tr}\{ {\hat o}_\Gamma {\hat G}(\k+\q,\e_n+\w_l;h)
{\hat o}_{\Gamma'} {\hat G}(\k,\e_n;h)\}$.
The obtained $A_{x^2-y^2}(xz,xz;\q)$ for $U=0$ is presented
in Fig. \ref{fig:quad} (a).
Similar result is obtained for $U=0.8$eV.
According to the functional form of $\chi_{xz}^{Q}(\q,0)$
in Eq. (\ref{eqn:chi-analytic}),
$\chi_{66}^{\rm TO}/T\propto \sum_\q \{ \chi_{xz}^{Q}(\q,0) \}^2
\propto\xi^2 \propto(T-T_{\rm AFQ})^{-1}$ in two-dimensional systems.
The numerical result for $\chi_{66}^{\rm TO}/T$ given in
Eq. (\ref{eqn:chi66-highT}) is shown in Fig. \ref{fig:quad} (b).
The obtained result follows the relation
$\chi_{66}^{\rm TO}/T \sim 0.1(1-\a_c)^{-1}$.
Since $6\xi^2\sim (1-\a_c)^{-1}$,
the relation $\chi_{66}^{\rm TO}/T\sim 0.6\xi^2$ is verified numerically.
In the same way, two-orbiton processes for other two
shear modulus susceptibilities, $\chi_{44}^{\rm TO}$ and $\chi_{E}^{\rm TO}$,
are proportional to the square of
$A_{\Gamma}(xz,xz;\q)$ for $\Gamma=xz$ and $xy$, respectively.
However, they are four orders of magnitude smaller than
$\chi_{66}^{\rm TO}$, as recognized in TABLE \ref{tab:tab1}:
This selection rule for $A_{\Gamma}$ is understood as follows:
According to the relation
${\rm Tr}\{ {\hat O}_{\Gamma}{\hat O}_{xz}{\hat O}_{xz} \}=0$
for $\Gamma=xz,xy$, we recognize that $A_{xz/xy}$ originates from the
off-diagonal terms of the Green function $G_{l,m}$ ($l\ne m$)
that is much smaller than the diagonal terms.
For this reason, the two-orbiton process
is negligible except for $\chi_{66}^{\rm TO}$.
\begin{table}[h]
\begin{center}
\begin{tabular}{c|c|c|c|c|c}
\hline
$\Gamma$ & & $x^2-y^2$ & $xz$ & $yz$ & $xy$ \\
\hline
$A_\Gamma(xz,xz,{\bm Q})$ & & $-0.60$ &
$1.9\times10^{-3}$ & $-1.0\times10^{-3}$ & $3.2\times10^{-4}$ \\
\hline
\end{tabular}
\caption{\label{tab:tab1}
Three point vertex $A_\Gamma(xz,xz,{\bm Q})$ for
$\Gamma=x^2-y^2$, $xz$, $yz$, and $xy$.
${\bm Q}=(\pi,0)$ corresponds to the peak position of
$\chi_{xz}^Q(\q,0)$ in the five-orbital model.
We recognize that $A_{\Gamma}(xz,xz,{\bm Q})\sim O(1)$ only for $\Gamma=x^2-y^2$;
This selection rule means that $\chi_{44,E}^{\rm TO}\ll1$.
}
\end{center}
\end{table}
Here, we discuss the softening of $C_{66}$
by taking the two-orbiton process into account:
According to Eqs. (\ref{eqn:Cphi}) and (\ref{eqn:Cx2}), we obtain
\begin{eqnarray}
C_{66}^{-1}&=&C_{66,0}^{-1}[1+g_{66}{\tilde \chi}_{66}]
\label{eqn:Cx66} ,\\
{\tilde \chi}_{66}&=& \frac{a_{66}+\chi_{66}^{\rm TO}}
{1-g_{66}(a_{66}+\chi_{66}^{\rm TO})}
\label{eqn:tildeC66} ,
\end{eqnarray}
where $a_{66}\equiv 2\chi_{66}^{0}({\bm 0},0)$.
Now, we consider the case (i) $T_{\rm AFQ}>0$ and $\w_0\ll T$.
As we have obtained the relation $\chi_{66}^{\rm TO}\propto T\xi^2$,
we put $\chi_{66}^{\rm TO}= b_{66}T/(T-T_{\rm AFQ})$.
Since the temperature dependence of $a_{66}$ is small, we obtain
\begin{eqnarray}
{\tilde \chi}_{66}&=&\frac{a_{66}+b_{66}}{1-g_{66}(a_{66}+b_{66})}
\frac{T-(a_{66}/(a_{66}+b_{66}))T_{\rm AFQ}}{T-T_S} ,
\nonumber \\
\label{eqn:tildeC66-2} \\
T_S&=& T_{\rm AFQ}\frac{1-g_{66}a_{66}}{1-g_{66}(a_{66}+b_{66})} \ \ (>T_{\rm AFQ})
\label{eqn:TS} .
\end{eqnarray}
Then, the difference between $T_S$ and $T_{\rm AFQ}$,
which is conventionally denoted as $E_{\rm JT}$ \cite{Thal,Levy},
is given by $E_{\rm JT}=T_S(g_{66}b_{66})/(1-g_{66}a_{66})>0$.
According to Eq. (\ref{eqn:tildeC66-2}),
Eq. (\ref{eqn:Cx66}) is rewritten as
\begin{eqnarray}
C_{66} = C_{66,0}(1-g_{66}(a_{66}+b_{66}))\frac{T-T_S}{T-T_{\rm AFQ}},
\label{eqn:C66-temp}
\end{eqnarray}
Here, $g_{66}=0.1\sim0.2$eV and $a_{66}\sim2{\rm eV}^{-1}$.
We stress that Eqs. (\ref{eqn:tildeC66-2})-(\ref{eqn:C66-temp})
are valid only for $\w_0\ll T$.
\begin{figure}[!htb]
\includegraphics[width=0.7\linewidth]{fig9.eps}
\caption{
(Color online)
(a) Obtained $A_{x^2-y^2}(xz,xz;\q)$ for $U=0$.
We put $n=6.05$ and $T=0.05$.
(b) $\chi_{66}^{\rm TO}/T$ given by Eq. (\ref{eqn:chi66-highT})
for $U=0$ and $U=0.8$ as function of $\a_c$.
Using the relation $6\xi^2\sim(1-\a_c)^{-1}$ given
in the caption of Fig. \ref{fig:chic}, we obtain
$\chi_{66}^{\rm TO}/T\sim0.1(1-\a_c)^{-1}\sim 0.6\xi^2$.
}
\label{fig:quad}
\end{figure}
Finally, we calculate the two-orbiton process
analytically for general value of $\w_0/T$.
Since we cannot apply the classical approximation that
was used to derive Eq. (\ref{eqn:chi66-highT}),
we have to perform the analytic continuation \cite{AGD}
of Eq. (\ref{eqn:chi66}).
The obtained expression including the quantum fluctuation
contribution is given as
\begin{eqnarray}
\chi_{66}^{\rm TO}&=& (2g)^{4}\{A_{x^2-y^2}(xz,xz;{\bm Q})\}^2
\sum_{\q}\int_{-\infty}^\infty \frac{dx}{\pi} n(x)
\nonumber \\
& &\times 2{\rm Im}\chi_{xz}^Q(\q,x+i\delta)
{\rm Re}\chi_{xz}^Q(\q,x+i\delta)
\label{eqn:chi66-analytic} ,
\end{eqnarray}
where $n(x)=(e^{\beta x}-1)^{-1}$ is the Bose distribution function,
and we put $A_{x^2-y^2}$ outside of the $\q$-summation
since its momentum dependence is much smaller than
that of $\chi_{xz}^Q$.
First, we perform the $x$-integration using the following equations:
\begin{eqnarray}
&& \frac1{e^x-1}= \sum_{n=1}^\infty \frac{2x}{(2n\pi)^2+x^2}+\frac1x ,
\\
&& \int_{-\infty}^\infty \frac{x^2}{(a^2+x^2)(b^2+x^2)^2}dx
=\frac{\pi}{2b(a+b)^2} ,
\end{eqnarray}
where $a, b>0$.
Then, the expression after the $x$-integration
in Eq. (\ref{eqn:chi66-analytic}) is given as
\begin{eqnarray}
c^2\xi^4\left( \w_0^2T\sum_{n=1}^\infty (B_\q\w_0+2n\pi T)^{-2}
+ \frac{T}{2B_\q^2} \right),
\label{eqn:chi66-3}
\end{eqnarray}
where $B_\q=1+\xi^2(\q-{\bm Q})^2$.
Next, we take the $\q$-summation
$\sum_\q \approx \frac1{2\pi}\int_0^{\pi}qdq$
under the assumption $\xi^2\gg1$.
Then, the $\q$-summation of the second term in Eq. (\ref{eqn:chi66-3})
is easily obtained as $c^2\xi^2T/8$.
Also, the $\q$-summation of the first term is given as
\begin{eqnarray}
&&\frac{c^2\xi^2\w_0}{8\pi}\sum_{n=1}^{n_{\rm max}}\frac1{n+\frac{\w_0}{2\pi T}}
\nonumber \\
&&=\frac{c^2\xi^2\w_0}{8\pi}\left[ \psi\left(n_{\rm max}+\frac{\w_0}{2\pi T}
+1\right) -\psi\left(\frac{\w_0}{2\pi T}+1\right) \right] ,
\nonumber \\
\end{eqnarray}
where $\psi(x)$ is di-Gamma function, and
the cutoff $n_{\rm max}\equiv(1+\xi^2\pi^2)\w_0/2\pi T$ originates from
the fact that the $\q$-summation is limited to the region $|\q|\le\pi$
in periodic systems.
As a result, the final expression for the two-orbiton term is
\begin{widetext}
\begin{eqnarray}
\chi_{66}^{\rm TO} &=& X \xi^2
\left\{ \frac{\w_0}{\pi} \left[ \psi\left(n_{\rm max}+\frac{\w_0}{2\pi T}
+1\right) -\psi\left(\frac{\w_0}{2\pi T}+1\right) \right] +T\right\}
\label{eqn:chi66-analytic2} ,
\end{eqnarray}
\end{widetext}
where $\displaystyle X\equiv
\frac{(2g)^4 c^2}{4\pi}\{A_{x^2-y^2}(xz,xz;{\bm Q})\}^2$.
Here, we verify Eq. (\ref{eqn:chi66-analytic2}) in the
opposite two limits:
In the case (i) $\w_0\ll T$,
the di-Gamma functions in Eq. (\ref{eqn:chi66-analytic2}) are negligible.
By applying the relation $\xi^2=l/(T-T_{\rm AFQ})$, we obtain
\begin{eqnarray}
\chi_{66}^{\rm TO} &\approx& X \xi^2 T
\nonumber \\
&\approx& b\frac{T}{T-T_{\rm AFQ}}
\label{eqn:chi66-analytic-highT} ,
\end{eqnarray}
where $b=Xl$.
The first line in Eq. (\ref{eqn:chi66-analytic-highT})
coincides with Eq. (\ref{eqn:chi66-highT})
since $\sum_\q\{\chi_{xz}^Q(\q,0)\}^2 = c^2\xi^2/4\pi$.
In the opposite case (ii) $\w_0\gg T$, the term $T$ in the curly brackets
in Eq. (\ref{eqn:chi66-analytic2}) is negligible.
Taking the relations $\psi(x)\approx \log(x)$ for $x\gg1$ and
$\w_0\xi^2\propto\xi^0$ into account, we obtain
\begin{eqnarray}
\chi_{66}^{\rm TO} &\approx& X \xi^2\w_0 \log(2+\pi^2\xi^2)
\nonumber\\
&\approx& b'_{66}\log\left(\frac{\pi^2 l}{T-T_{\rm AFQ}}\right)
\label{eqn:chi66-analytic-lowT} ,
\end{eqnarray}
where $b'=X \xi^2\w_0$.
Therefore, in the case (ii) $T_{\rm AFQ}<0$ and $\w_0\gg T$,
${\tilde \chi}_{66}$ in Eq. (\ref{eqn:Cx66}) is given
by replacing $\chi_{66}^{\rm TO}$ with $b'_{66}\log({\pi^2 l}/(T-T_{\rm AFQ}))$
in Eq. (\ref{eqn:tildeC66}).
In this case, the temperature dependence of
$\chi_{66}^{\rm TO}$ is much moderate.
In Sec. \ref{sec:Dis-C66}, we will discuss the temperature dependence
of $C_{66}$ based on Eq. (\ref{eqn:chi66-analytic2}).
In the above derivation, we have neglected the effect of
mass-enhancement factor brought by the third point vertex.
If we take this effect into account, both
Eqs. (\ref{eqn:chi66-analytic-highT}) and (\ref{eqn:chi66-analytic-lowT})
are multiplied by the factor $(m^*/m)^2=2^2\sim3^2$,
as we will discuss in Sec. \ref{sec:Dis-C66}.
\section{Discussions
\label{sec:discussion}}
\subsection{Why $O_{x^2-y^2}$-FQ fluctuations are the most
divergent for $T_{\rm AFQ}>0$?}
In this paper, we have studied the development of
quadrupole susceptibilities in iron pnictides
based on the RPA and beyond the RPA.
The main fluctuations in the present study is the $O_{xz/yz}$-AFQ
fluctuations, $\chi_{xz/yz}^Q({\bm Q})$,
which are produced by Fe-ion in-plane optical phonons.
The acoustic phonons $\q\sim\Q$ with finite energy also
assist in producing the AFQ fluctuations; see Eq. (\ref{eqn:chi44U0}).
We also find that the $O_{x^2-y^2}$-FQ fluctuations, $\chi_{66}$,
are induced by the ``two-orbiton process'' described by
the AL-type diagram in Fig. \ref{fig:AL} (c).
Here, the anharmonic three-phonon coupling
is produced by the three-point vertex in Eq. (\ref{eqn:Aw0}).
The two-orbiton process is important in iron pnictides
because of the two-dimensionality.
We discuss on the {\it total susceptibility} by including
the electron-acoustic phonon interaction, ${\tilde \chi}_\Gamma^Q(\q)$,
by taking the two-orbiton process into account.
For $\Gamma=xz$ and $x^2-y^2$,
\begin{eqnarray}
{\tilde \chi}_{xz}^Q({\bm Q})
&=& \frac{\chi_{xz}^0({\bm Q})}{1-(g+g_{44})\chi_{xz}^0({\bm Q})},
\\
{\tilde \chi}_{x^2-y^2}({\bm 0})
&=& \frac{\chi_{66}^0({\bm 0})+\chi_{66}^{\rm TO}}
{1-g_{66}(\chi_{66}^0({\bm 0})+\chi_{66}^{\rm TO})},
\end{eqnarray}
where $\chi_{66}^{\rm TO}$ is proportional to the square of the AFQ
correlation length $\xi^2\propto {\tilde \chi}_{xz}^Q({\bm Q})$.
When $g_{66}=0$, therefore,
${\tilde \chi}_{xz}^Q({\bm Q})$ and
${\tilde \chi}_{x^2-y^2}^Q({\bm 0})$
diverges at the same time in proportion to $\xi^2$.
As $g_{66}$ increases from zero, only ${\tilde \chi}_{x^2-y^2}^Q({\bm0})$
is enhanced because of the absence of the two-orbiton process;
$A_\Gamma(xz,xz;\q),A_\Gamma(yz,yz;\q)\ll1$ for $\Gamma=xz$ or $yz$
as shown in TABLE \ref{tab:tab1}.
For this reason, the relation $T_S>T_{\rm AFQ}$ is universally satisfied
even if a fully self-consistent calculation is performed.
As results, both the AFQ fluctuations (=origin of high-$T_{\rm c}$)
and the FQ fluctuations (=origin of shear modulus softening)
develop at the same time, and latter fluctuations overcome
the former near the $T_S$.
The orthorhombic phase transition
in under-doped compounds is brought by the divergence of
the two-orbiton process ${\tilde \chi}_{66}$.
\subsection{Softening of $C_{66}$: comparison between theory and experiment
\label{sec:Dis-C66}}
Is this subsection, we discuss the softening of $C_{66}$
in under- and over-doped iron pnictides
based on the results in Sec. \ref{sec:shear}.
For this purpose, we first estimate the magnitude of the three-point vertex
$A_{x^2-y^2}(xz,xz,\Q)$ based on the Ward identity given
in Eq. (\ref{eqn:Ax2y2-W}).
Since ${\bar \chi}_{xz,xz}^{Q}(\Q,0;0)\sim(2g)^{-1}$,
$|A_{x^2-y^2}|\sim (2g)^{-1}/\delta E$, where $\delta E$ is the
bandwidth of the $xz/yz$ band.
Since $2g\sim0.5{\rm eV}$ and $\delta E\sim 2{\rm eV}$
according to the band calculations \cite{Kuroki},
and considering the effect of band renormalization due to the
mass-enhancement factor $m^*/m\ (=2\sim3)$ \cite{Shishido},
we expect $|A_{x^2-y^2}|\sim 1(m^*/m)\ [{\rm eV}^{-2}]$.
This rough estimation is consistent with the numerical result
in Fig. \ref{fig:quad} (a).
Now, we discus the under-doped case with $T_{\rm AFQ}>0$.
In this case, $\chi_{66}^{\rm TO}/T\approx X\xi^2$
shown in Eq. (\ref{eqn:chi66-analytic-highT}).
Using the relations $c\sim12$ and $6\xi^2\sim (1-\a_c)^{-1}$
as discussed in the caption of Fig \ref{fig:chic}, we obtain
$X\sim 0.7$ and $\chi_{66}^{\rm TO} \sim 0.12(m^*/m)^2(1-\a_c)^{-1}$.
This estimation is consistent with the numerical result
Fig. \ref{fig:quad} (b) if we put $(m^*/m)=1$.
We also discuss the optimum or over-doped systems without
structure transition, in which the relation $\w_0\gg T$ is satisfied.
In this case, $\chi_{66}^{\rm TO}\sim b'\log(\pi^2l/(T-T_{\rm AFQ}))$.
Since the temperature dependence of $\chi_{66}^{\rm TO}$ is moderate,
$\chi_{66}^{\rm TO}$ would be comparable or smaller than $a_{66}$.
Therefore, in over-doped systems,
the softening in $C_{66}$ would be much moderate, showing a
deviation from the Curie-Weiss type form in Eq .(\ref{eqn:C66-temp}).
Now, we analyze the temperature dependence of $\chi_{66}^{\rm TO}$
and $C_{66}/C_{66,0}$ by using Eq. (\ref{eqn:chi66-analytic2}).
We can fix the prefactor
$(2g)^4c^2\xi^2\{A_{x^2-y^2}\}^2/4\pi \equiv X \xi^2$
in front of Eq. (\ref{eqn:chi66-analytic2})
based on the relation $\chi_{66}^{\rm TO}/T = X\xi^2$ for $\w_0\ll T$:
We obtain $X\sim 0.6$ according to Fig. \ref{fig:quad} (b).
Hereafter, we put $X=5.4$ by multiplying the
square of the mass-enhancement factor, ($m^*/m)^2\sim9$.
We also put $\w_0=l'(T-T_{\rm AFQ})$ with $l'=2$,
and $\xi^2=l(T-T_{\rm AFQ})^{-1}$ with $l=0.086\ [{\rm eV}]$,
which means that $\xi\sim2$ when $T-T_{\rm AFQ}=250$K.
Using the obtained $\chi_{66}^{\rm TO}$,
we plot $C_{66}/C_{66,0}$ in Fig. \ref{fig:MATH} (b)
based on Eqs. (\ref{eqn:Cx66}) and (\ref{eqn:tildeC66}).
Here, we set $g_{66}=0.17$eV and $a_{66}=2.5{\rm eV}^{-1}$.
In the case of $T_{\rm AFQ}=100$K, we obtain $E_{\rm JT}\approx27$K.
In the FLEX approximation \cite{Onari},
$T_{\rm AFQ}$ changes from positive to negative by carrier doping, while
other parameters ($X$, $l$, $l'$ and $a_{66}$) are insensitive to the doping.
Similarly to Fig. \ref{fig:MATH} (b), we can fit the recent experimental
data by Yoshizawa {\it et al.} \cite{comment4}
for Ba(Fe$_{1-x}$Co$_x$)$_2$As$_2$ with $x=0\sim0.1$, by choosing $T_{\rm AFQ}$
while other parameters ($X$, $l$, $l'$ and $a_{66}$) are fixed.
This fact is a strong evidence for the success of orbital fluctuation theory
in iron pnictide superconductors.
\begin{figure}[!htb]
\includegraphics[width=0.98\linewidth]{fig10.eps}
\caption{
(Color online)
(a) $\chi_{66}^{\rm TO}$ for $T_{\rm AFQ}=100$K, $50$K, $10$K and $-10$K.
given by Eq. (\ref{eqn:chi66-analytic2}).
(b) $C_{66}/C_{66,0}$ for $T_{\rm AFQ}=100$K, $50$K, $10$K and $-10$K.
given by Eqs. (\ref{eqn:Cx66}) and (\ref{eqn:tildeC66}).
We put $X=5.7$, $l=0.086$, $l'=2$, $g_{66}=0.17$, and $a_{66}=2.5$.
$C_{66}=0$ is realized when $\chi_{66}^{\rm TO}=g_{66}^{-1}-a_{66}$,
which is $3.38$ in the present parameters.
Using these {\it same} parameters, we can fit the recent
experimental data by Yoshizawa {\it et al.} \cite{comment4}
for Ba(Fe$_{1-x}$Co$_x$)$_2$As$_2$ for $x=0\sim0.1$,
just by changing $T_{\rm AFQ}$.
}
\label{fig:MATH}
\end{figure}
In the present paper, we consider that the origin of high-$T_{\rm c}$
is the AFQ fluctuations.
On the other hand, Yanagi {\it et al}. \cite{Ohno2} claimed that
high-$T_{\rm c}$ originates from the FQ fluctuations
that give the softening in $C_{66}$:
In the latter mechanism, a rough estimation of $T_{\rm c}$ is given as
\begin{eqnarray}
T_{\rm c}\sim \w_c \exp(-(1+\beta\lambda)/\beta\lambda),
\label{eqn:BCS}
\end{eqnarray}
where $\w_c$ is
the phonon energy relevant for the orbital fluctuations,
which is just $\sim10$K for $|\q|\sim0.1\pi$.
$\beta\equiv 1+g_{66}{\tilde \chi}_{66}=C_{66,0}/C_{66}$ is the
enhancement factor due to FQ fluctuations \cite{comment}.
However, $C_{66,0}/C_{66}$ observed in optimally-doped Ba(Fe,Co)$_2$As$_2$
is just $\sim1.2$ \cite{Fernandes,comment4}:
Apparently, such small enhancement cannot reproduce
high-$T_{\rm c}$ superconductivity in iron-pnictides.
In the present study, in contrast, weak softening in optimally-doped
sample is ascribed to the change in the scaling of $\chi_{66}^{\rm TO}$,
not to the weakness of AFQ fluctuations.
In fact, the softening is moderate in the case of $T_{\rm AFQ}=-10$K
in Fig. \ref{fig:MATH}, while the AFQ correlation
$\xi^2\approx 1000/(T{\rm [K]}+10)$ is enough to cause the
superconductivity at $T_{\rm c}\sim30$K.
Therefore, moderate softening and high-$T_{\rm c}$ are compatible
in the present study.
\subsection{Quadrupole-ordered state in under-doped compounds}
Here, we consider the orbital or quadrupole ordered state in
under-doped compounds.
In the mean-field approximation for the multiorbital Hubbard model
for iron pnictides \cite{OO-theory-MF,OO-theory-MF2},
stripe-type SDW order occurs for $U>U_{\rm c}$,
and weak orbital polarization ($n_{xz}\ne n_{yz}$) is induced
as the secondary order when the magnetization is large.
However, in real materials, orthorhombic transition occurs
in the paramagnetic state, and the SDW-order is induced
in the orthorhombic phase.
To solve this problem,
we studied the multiorbital HH model beyond the mean-field theory,
and found that the FQ order occurs in the paramagnetic state
due to the two-orbiton process.
Fortunately, this FQ order does produce the experimentally observed
SDW order, as we will explain in the next subsection.
As discussed in Sec. \ref{sec:AL}, the divergence of ${\tilde \chi}_{66}$,
which is the total FQ susceptibility
given by both optical and acoustic phonons,
causes the orthorhombic structure transition when $C_{66}=0$.
The $O_{x^2-y^2}$-FQ order is realized in the orthorhombic phase.
The schematic quadrupole order is shown in Fig. \ref{fig:OO-fig} (a).
Since $O_{x^2-y^2}\approx n_{2}-n_{3}$ according to Eq. (\ref{eqn:ox2y2}),
the order parameter $O_{x^2-y^2}>0$ ($<0$) corresponds
the orbital polarized state with $n_{xz}>n_{yz}$ ($n_{xz}<n_{yz}$).
Figure \ref{fig:OO-fig} (b) shows the AFQ order
brought by the divergence of $\chi^Q_{xz}({\bm Q})$.
Although the FQ order in (a) would occur earlier,
we expect the AFQ order in (b) would coexist with the FQ order
when the structure transition is the weak first order.
In fact, the reconstruction of the FSs above $T_{\rm N}$
in detwinned Ba(Fe$_{1-x}$Co$_x$)$_2$As$_2$ \cite{ARPES3}
would indicate the presence of the AFQ order \cite{comment3}.
\begin{figure}[!htb]
\includegraphics[width=0.9\linewidth]{fig11.eps}
\caption{
(Color online)
(a) $O_{x^2-y^2}$-FQ order given by the divergence of $C_{66}^{-1}$.
Be careful not to confuse ${\hat O}_{x^2-y^2}$ with the
$x^2-y^2$-orbital operator.
(b) $O_{xz}$-AFQ order brought by the divergence of $\chi_{xz}^Q({\bm Q})$.
(c) The correspondence between $O_{x^2-y^2}$-quadrupole order
($O_{xz}$-quadrupole order) and the $d$-orbital
with larger occupation number.
}
\label{fig:OO-fig}
\end{figure}
In Fig. \ref{fig:OO-fig} (c),
we show the correspondence between the quadrupole order
and the $d$-wavefunction with larger electron occupancy.
In the $O_{x^2-y^2}$-type quadrupole order,
the electrons mainly occupy the state $|xz\rangle$ for $O_{x^2-y^2}>0$,
or the state $|yz\rangle$ for $O_{x^2-y^2}<0$.
In the $O_{xz}$-type quadrupole order,
the electrons mainly occupy the state
$|xy\rangle+|xz\rangle$ for $O_{xz}>0$
($|xy\rangle-|xz\rangle$ for $O_{xz}<0$).
Finally, we make comparison between the present study
and the previous work based on the RPA \cite{Ohno2},
which claims that the divergence of Eq. (\ref{eqn:Cx66})
is caused by large $g_{66}a_{66}\lesssim1$ while neglecting $\chi_{66}^{\rm TO}$.
However, the obtained $O_{x^2-y^2}$-quadrupole order is ``incommensurate''
\cite{comment5}.
This result highlights the importance of the two-orbiton process
$\chi_{66}^{\rm TO}$ in order to produce the
``$\q=0$'' orthorhombic structure transition.
\subsection{Stripe magnetic order produced by
$O_{x^2-y^2}$-FQ order}
In under-doped iron pnictides,
the collinear-SDW order is induced in the orthorhombic phase
at $T_{\rm N}$, which is slightly lower temperature than $T_S$.
Various explanations for the origin of this SDW transition
had been proposed previously.
From a strong-coupling scheme,
square-lattice Heisenberg model with in-plane anisotropy,
$J_{1a}$-$J_{1b}$-$J_{2}$ model, had been studied \cite{frustration}.
According to the neutron scattering on CaFe$_2$As$_2$ \cite{Dai},
the high-energy spin-wave dispersion indicates the relation
$J_{1a}\gg J_{1b}$ in the orthorhombic phase.
In this case, experimentally observed staggered spin order
along the $x$-axis ($a$-axis) is expected to be realized.
However, such strong in-plane anisotropy $J_{1a}\gg J_{1b}$
is surprising, considering the small orthorhombicity
$(a-b)/(a+b)\sim0.003$.
These fact would indicate the existence of orbital or quadrupole order
in the orthorhombic phase.
\begin{figure}[!htb]
\includegraphics[width=0.98\linewidth]{fig12.eps}
\caption{
(Color online)
(a) $\chi^s(\q;\Delta E)$ for $\Delta E=0$.
Used model parameters are $n=6.05$, $U=1.1$, $g=0$, and $T=0.05$
(b) $\chi^s(\q;\Delta E)-\chi^s(\q;0)$ for $\Delta E=0.04$.
Model parameters are the same as those in (a).
(c) $U_c$ as function of $\Delta E$ given by the RPA for $m^*/m=1$.
(d) $\Delta n_{2}$ and $\Delta n_{3}
$ as function of $\Delta E$ for $m^*/m=1$.
Note that $\Delta n\equiv \Delta n_{2}- \Delta n_{3}$.
(e) FSs for $\Delta E=-0.04$ for $n=6.05$.
We use the same color coding as in Fig. \ref{fig:fig1}.
(f) FSs for $\Delta E=-0.08$ for $n=6.05$.
In both cases, the best nesting vector is $\q=(\pi,0)$.
}
\label{fig:sdw}
\end{figure}
In this subsection, we study the origin of SDW state
based on the weak-coupling approach.
Hereafter, we assume that $x$-axis corresponds to $a$-axis
(longer lattice constant) in the orthorhombic phase.
In the previous subsection,
we explained that the two-orbital process induces
the $O_{x^2-y^2}$-FQ order in Fig. \ref{fig:OO-fig}.
Note that $O_{x^2-y^2}\approx n_{2}-n_{3}$ according to Eq. (\ref{eqn:ox2y2}).
The corresponding mean-field is given as
\begin{eqnarray}
H'=\Delta E \sum_i \left( |2\rangle\langle 2|
-|3\rangle\langle 3| \right)_i,
\end{eqnarray}
which raises (lowers) the energy-level of orbital 2 (3) by $\Delta E$.
In a similar model, the change in the DOS and FSs by the
orthorhombic potential $\Delta E$ was studied
by Chen {\it et al.} \cite{Chen}.
Here, we study the change in the spin susceptibility
by $\Delta E$ using the RPA.
We calculate the total spin susceptibility
$\chi^s(\q,0)=\sum_{l,m}\chi_{l,l;m,m}^s(\q,0)$
for $U=1.1$ and $g=0$.
Figure \ref{fig:sdw} (a) shows the obtained $\chi^s(\q,0)$
for $\Delta E=0$; the corresponding spin Stoner factor is $\a_S=0.87$.
When $\Delta E$ is finite, the four-hold symmetry in $\chi^s(\q,0)$
disappears quickly.
Figure \ref{fig:sdw} (b) shows the change in the spin susceptibility,
$\chi^s(\q;\Delta E)-\chi^s(\q;0)$, induced by $\Delta E=+0.04$.
We see that $\chi^s(\q,0)$ increases by $+6.5$ at $\q=(0,\pi)$
while decreases by $-4.0$ at $\q=(\pi,0)$.
Therefore, magnetic frustration is resolved
and stripe-SDW order can be induced by small $\Delta E$.
Figure \ref{fig:sdw} (c) shows the $\Delta E$-$U_c$ phase diagram
given by the RPA, that is, by the mean-field-approximation.
It is noteworthy that $U_c$ quickly decreases in proportion to
$|\Delta E|$, because of the degeneracy of orbital 2 and 3.
When $U\lesssim U_{\rm c}$, the experimental SDW order
with momentum $\q=(\pi,0)$ is realized by the negative
potential $\Delta E$ that corresponds to $n_2>n_3$.
Figure \ref{fig:sdw} (d) gives the relation between
$\Delta n=\Delta n_{2}-\Delta n_{3}$ and $\Delta E$:
If we take the band-renormalization effect into account,
we obtain the relation $\Delta n= -0.85(m^*/m)\Delta E$.
According to (c) and (d), we obtain the reduction in $U_c$
due to the FQ order is
\begin{eqnarray}
\Delta U_c= -1.4|\Delta n|= -1.2(m^*/m)|\Delta E|.
\label{eqn:UcDnDe}
\end{eqnarray}
Therefore, only few percent $\Delta n$ can induce large change
in $U_{\rm c}$ that is {\it linear in $|\Delta n|$}.
According to recent ARPES measurement in detwinned BaFe$_2$As$_2$
\cite{ARPES3}, $\Delta E\sim-0.03$eV in the orthorhombic phase,
which corresponds to $\Delta n=+0.026(m^*/m)$
and $\Delta U_c=-0.036(m^*/m)$ in the present five-orbital model.
In this case, the realized SDW order is $\q=(\pi,0)$,
which is consistent with famous strip-type SDW state
in mother compounds \cite{review}.
We can show that the SDW temperature $T_{\rm N}$ also increases
{\it linearly in $|\Delta n|$} based on the Landau theory.
The free energy in the present problem would be given as
\begin{eqnarray}
F(\Delta n)=F(0)+ c\Delta n(m_{(\pi,0)}^2-m_{(0,\pi)}^2),
\end{eqnarray}
where $m_{\Q}$ is the AF order with momentum $\Q$, and
$F(0)=a\cdot (T-T_{\rm N}^0)(m_{(\pi,0)}^2+m_{(0,\pi)}^2)
+b(m_{(\pi,0)}^4+m_{(0,\pi)}^4)/2 +\cdots$ with $a,b>0$.
Then, we obtain $T_{\rm N}=T_{\rm N}^0+ |c\Delta n|/a$.
The present study shows that $c<0$,
which seems consistent with the numerical result
in Ref. \onlinecite{OO-theory-MF2}.
Now, we consider the reason why SDW order is produced by $\Delta E$:
Figure \ref{fig:sdw} (e) and (f) shows the change in the
the FS structure with $\Delta E$.
We can recognize that the intra-orbital (orbital 3) nesting
between FS $\a_2$ and FS $\b_1$ becomes better, compared to the
case of $\Delta E=0$ in Fig. \ref{fig:fig1} (a).
Therefore, the origin of the ``FQ-order-induced stripe-SDW'' is
the ``anisotropy in the intra-orbital nesting''
caused by small $|\Delta E|\sim 0.03$eV,
which corresponds to
a small orbital polarization $|\Delta n|\sim0.026(m^*/m)$.
This result is consistent with the very small orthorhombicity
$(a-b)/(a+b)\lesssim0.003$ in the orthorhombic state \cite{review}.
In the strong-coupling description,
the origin of the stripe-SDW state is the
in-plane anisotropy in the exchange interaction ($J_{1a}\ne J_{1b}$)
\cite{frustration,Dai} brought by two-orbiton process.
If we go beyond the RPA,
the SDW state will be further stabilized by
the reduction in the quasiparticle damping $\gamma$
when the FQ-order is established \cite{Onari}:
In fact, in the FLEX approximation \cite{Onari}, $\chi^s(\q)$ is
suppressed by $\gamma$ due to strong orbital fluctuations in the normal state.
Since the orbital fluctuations is suppressed
when the AFQ-order sets in,
the resultant increment in $\chi^s(\q)$ would stabilize the SDW phase.
\begin{figure}[!htb]
\includegraphics[width=0.95\linewidth]{fig13.eps}
\caption{
(Color online)
The phase-diagram for iron-pnictide superconductors
obtained by the present orbital fluctuation theory.
$T_S$ is the orthorhombic transition temperature
(= FQ order temperature), and $T_{\rm N}$ is SDW transition temperature.
The fact that two QCPs at $T_S=0$ and $T_{\rm AFQ}=0$ almost coincide
means that novel ``multi orbital QCPs'' are realized in iron pnictides.
The left-hand (right-hand) side of the vertical dotted line
corresponds to $T_{\rm AFQ}>0$ ($T_{\rm AFQ}<0$), in which the
two-orbiton process is relevant (irrelevant).
At $T_{\rm AFQ}$, the AFQ-order does not occur
since it is prevented by the FQ-order at $T_S$.
}
\label{fig:phase}
\end{figure}
\subsection{Summary}
In the present paper,
we have studied the realistic five-orbital HH model for iron pnictides.
In the RPA, only the $O_{xz}$-AFQ fluctuations develop
as shown previously \cite{Kontani}, and therefore the softening of
shear moduli ($C_{66}$, $C_{44}$ and $C_E$) cannot be reproduced.
In the present study beyond the RPA, we revealed that
both $O_{xz}$-AFQ fluctuations and $O_{x^2-y^2}$-FQ fluctuations
develop at the same time.
The former and the latter fluctuations are the origins of the
$s_{++}$-wave superconductivity and the orthorhombic structure
transition, respectively.
The commensurate FQ fluctuations are brought by the two-orbiton process
in Fig. \ref{fig:AL} (c) that is dropped in the RPA.
[In the mean-field theory, the orbital order due to large $g_{66}$
is always ``incommensurate'' \cite{comment5}.]
Fluctuation-induced softening occurs only in $C_{66}$ out of three shear moduli
because of the orbital selection rule for the three-point vertex.
The origin of softening would be interpreted as ``virtual
anharmonicity of lattice vibrations'' that is induced by AFQ fluctuations;
see Fig.\ref{fig:AL} (b).
Possible quadrupole orders in the ordered state
are show in Fig. \ref{fig:OO-fig}.
Using the two-orbiton term in Eq. (\ref{eqn:chi66-analytic2}),
we can fit the recent experimental data of $C_{66}$ in
Ba(Fe$_{1-x}$Co$_x$)$_2$As$_2$ \cite{comment4} for wide range of doping,
only by choosing $T_{\rm AFQ}$ while other parameters are fixed.
This fact is a strong evidence for the success of orbital fluctuation theory
in iron pnictide superconductors.
In addition,
we should stress that the stripe-type antiferro-magnetic state
is realized in the orbital-ordered state,
since the small orbital polarization ($\Delta n\lesssim0.05$) can cause
large in-plane anisotropy in the exchange interaction ($J_{1a}\ne J_{1b}$).
Thus, the present study presents a microscopic justification for
the anisotropic Heisenberg model description
for the SDW state \cite{frustration,Dai}.
In Fig. \ref{fig:phase}, we summarize the phase-diagram
of iron-pnictides given by the present orbital fluctuation theory
beyond the RPA.
We stress that $T_{\rm AFQ}$, which is determined experimentally from $C_{66}$,
is positive in the under-doped case ($T_S>0$) while it is negative
in the over-doped case, as recognized from Eq. (\ref{eqn:TS})
obtained in the classical approximation.
Especially, $T_{S}\approx0$ for $T_{\rm AFQ}=0$,
consistently with experiments \cite{Yoshizawa,Goto}.
This result indicats that QCPs for AFQ and FQ orders almost coincide
at the endpoint of the orthorhombic phase.
The emergence of ``multi orbital QCPs'' is favorable to
the orbital-fluctuation-mediated $s_{++}$-wave SC state
\cite{Kontani,Saito,Onari}.
In fact,
$T_{\rm AFQ}$ is derived from experimentally observed $C_{66}$ as follows:
In under-doped systems with $T_S>0$, $T_{\rm AFQ}$ is given
the Weiss temperature of $C_{66}\propto (T-T_S)/(T-\theta)$, and
$T_{\rm AFQ}=\theta$ is indeed positive experimentally.
In over-doped systems, both $T_S$ and $\theta$ are negative ($T_S>\theta$),
and $C_{66}$ starts to deviate from the Curie-Weiss behavior.
These experimental results are the strong evedence for the
realization of the two-orbiton process ($\chi_{66}^{\rm TO}\sim T/(T-T_{\rm AFQ})$)
in iron-pnictides.
In contrast, in the cooperative Jahn-Teller scenario
due to large $g_{66}$ by Yanagi {\it et al.} \cite{Ohno2}, the parameter
$\theta$ is always negative; see Appendix \ref{sec:ApB} in detail.
Finally, we note that the two-orbiton term $\chi_{66}^{\rm TO}$
in the present study is very similar to the bare
nematic susceptibility $\chi_{\rm 0,nem}$ in Ref. \cite{Fernandes},
which is the two-magnon term on different sublattices in our terminology.
In summary, the present study
can explain the superconductivity, orthorhombic transition,
and softening of $C_{66}$ due to FQ and AFQ quantum-criticalities.
The stripe-SDW order is naturally produced
by the ``orthorhombicity'' of the FQ order.
These results are strong evidence for the realization of the
orbital-fluctuation-mediated $s_{++}$-wave superconductivity in iron pnictides.
Finally, we stress that the present study enables us to derive the
important parameters in the orbital fluctuation model in
Eqs. (\ref{eqn:chi-analytic})-(\ref{eqn:w0})
from the experimental data of shear modulus.
\acknowledgements
We are grateful to M. Yoshizawa for valuable discussions
on his interesting experimental results.
We also thank D.S. Hirashima, M. Sato, Y. Matsuda, T. Shibauchi,
and R.M. Fernandes for valuable comments and discussions.
This study has been supported by Grants-in-Aid for Scientific
Research from MEXT of Japan, and by JST, TRIP.
Numerical calculations were performed using the facilities of
the supercomputer center, Institute for Molecular Science.
|
\section{Introduction}
\label{Sec:S_1}
Quite recently there has been a notable increase of experimental activities
aiming at the exploration of properties of thermally generated fluctuating
electromagnetic fields close to the surface of some material, and at detecting
the near-field mediated heat transfer.~\cite{RytovEtAl89,JoulainEtAl05,
VolokitinPersson07} Hu {\em et al.\/} have measured the near-field thermal
radiation between \textmu m-spaced glass plates, and have demonstrated that
the resulting near-field heat transfer exceeds the far-field limit set by
Planck's blackbody radiation law.~\cite{HuEtAl08} Next, Narayanaswamy
{\em et al.\/} and Shen {\em et al.\/} have studied the heat transfer between
microspheres and flat substrates, with emphasis on the coupling of surface
phonon polaritons across the gap between them, and have reported heat
transfer coefficients three orders of magnitude above the blackbody radiation
limit.~\cite{NarayanaswamyEtAl08,ShenEtAl09} Then Rousseau {\em et al.\/}
have carried out precise measurements of the radiative heat transfer between
sodalime glass spheres with diameters of 22 or 40 \textmu m and
borosilicate glass plates for distances ranging from 30~nm to
2.5~\textmu m,~\cite{RousseauEtAl09} and have verified theoretical predictions
based on fluctuational electrodynamics~\cite{RytovEtAl89,PolderVanHove71} with
impressive accuracy. On the other hand, significant progress has been made at
using near-field effects for imaging. Kittel {\em et al\/.} are developing
a device termed Near-Field Scanning Thermal Microscope
(NSThM)~\cite{MuellerHirschEtAl99,KittelEtAl05,WischnathEtAl08} which does not
yet seem capable of highly accurate quantitative measurements of the near-field
heat current between its sensor and the sample, but which lends itself to
nanoscale thermal imaging of structured surfaces.~\cite{KittelEtAl08} Moreover,
De Wilde {\em et al.\/} have reported the successful operation of a Thermal
Radiation Scanning Tunneling Microscope (TRSTM),~\cite{DeWildeEtAl06} providing
images of thermally excited surface plasmons, and giving clear evidence for
spatial coherence effects in near-field thermal emission.
These remarkable developments indicate that thermal near-field physics,
after having been under intense theoretical investigation for some time
already,~\cite{RytovEtAl89,JoulainEtAl05,VolokitinPersson07,PolderVanHove71}
is breaking through to the forefront of experimental research right now.
There are several compelling reasons for this trend: Besides the prospects of
obtaining novel insight into fundamental physics in dielectric matter, and of
developing advanced diagnostic tools for materials science, thermal near-field
effects have great potential for near-field thermo\-photo\-voltaic energy
conversion.~\cite{DiMatteoEtAl01,NarayanaswamyChen03,LarocheEtAl06,
FrancoeurEtAl08,BasuEtAl09}
On the theoretical side, one of the most important quantities
characterizing the fluctuating thermal near field close to a
dielectric surface is its local density of states
(LDOS).~\cite{JoulainEtAl03} In particular, the power~$P$ transferred
between a dielectric sample at temperature~$T_{\rm S}$ and a nanoparticle at
temperature~$T_{\rm P}$, brought into the sample's near field at a position
$\vec{a}$ such that the particle may effectively be treated within the
dipole approximation, and the heat transfer proceeds almost entirely
via evanescent modes, is given by (see, e.g.,
Refs.~\onlinecite{Dorofeyev98,Pendry99,MuletEtAl01,DedkovKyasov07,
Dorofeyev08,ChapuisEtAl08})
\begin{align}
\label{eq:Pdip}
P = & \int\limits_0^{\infty} \! {\rm d} \omega \,
2 \omega \left[ \Theta(\omega, T_{\rm P}) - \Theta(\omega, T_{\rm S}) \right] \\
& \times \left[\alpha_{\rm P}''(\omega) D^{\rm E}(\omega,\vec{a}) +
\mu_{\rm P}''(\omega) D^{\rm H}(\omega,\vec{a}) \right] \; ,
\nonumber
\end{align}
where $D^{\rm E}(\omega,\vec{a})$ is the electric and $D^{\rm H}(\omega,\vec{a})$ is
the magnetic part of the sample's LDOS at the point $\vec{a}$ of effective
interaction; $\alpha_{\rm P}''(\omega)$ and $\mu_{\rm P}''(\omega)$ denote the
imaginary part of the particle's electric and magnetic polarizability,
respectively. Finally,
\begin{equation}
\Theta(\omega, T) = \frac{\hbar\omega}
{\exp(\hbar\omega/k_{\rm B}T) - 1}
\end{equation}
is the Bose-Einstein function; the sign in Eq.~(\ref{eq:Pdip}) is chosen such
that a net energy transfer from the particle to the sample, occurring for
$T_{\rm P} > T_{\rm S}$, gives a positive~$P$. The use of the dipole
approximation underlying Eq.~(\ref{eq:Pdip}) requires that the distance of
the nanoparticle from the sample's surface remains large compared to its linear
size. Those frequencies which significantly contribute to the heat transfer
are limited by the higher temperature $T_{\max}=\max(T_{\rm P},T_{\rm S})$.
Provided the polarizabilities and the LDOS exhibit no resonances in the
accessible frequency regime, requiring in particular the absence of thermally
excitable surface modes, the main contribution to the integral~\reff{eq:Pdip}
merely stems from frequencies in the vicinity of the thermal frequency
$\omega_{\rm th} \approx 2.82 \; k_{\rm B}T/\hbar$, so that
\begin{align}
\label{eq:dipapprox}
P \propto \omega_{\rm th} &
\big[\alpha_{\rm P}''(\omega_{\rm th}) D^{\rm E}(\omega_{\rm th},\vec{a}) \\
& + \mu_{\rm P}''(\omega_{\rm th}) D^{\rm H}(\omega_{\rm th},\vec{a}) \big] \; .
\nonumber
\end{align}
Under suitable conditions, already this simple approximation can give
surprisingly good results when trying to theoretically reconstruct
surface images obtained with the NSThM.~\cite{KittelEtAl08}
The possibility to experimentally assess the LDOS above nanostructured
surfaces demands refined techniques for its calculation. Assuming local
thermal equilibrium, and considering positions~$\vec{r}$ so close to the
sample's surface that the energy density is dominated by evanescent modes
and the contribution of propagating modes can be neglected, the electric
and the magnetic LDOS are related to the imaginary parts of the traces of
the renormalized (or `reflected') electric and magnetic Green's dyadics
$\mathds{G}^{\rm E}_{\rm r}$ and $\mathds{G}^{\rm H}_{\rm r}$ through the
relations~\cite{JoulainEtAl03}
\begin{equation}
D^{\rm E}(\omega,\vec{r}) = \frac{\omega}{\pi c^2}
\Im \, {\rm Tr} \, \mathds{G}^{\rm E}_{\rm r}(\vec{r},\vec{r})
\end{equation}
and
\begin{equation}
D^{\rm H}(\omega,\vec{r}) = \frac{\omega}{\pi c^2}
\Im \, {\rm Tr} \, \mathds{G}^{\rm H}_{\rm r}(\vec{r},\vec{r}) \; .
\end{equation}
In the present paper we exploit this connection for computing the LDOS above
a nano\-structured surface to second order in the surface profile, relying on
an earlier formulation of the perturbation series by Greffet.~\cite{Greffet88}
The method is technically involved, and soon hits practical computational
limits when proceeding to higher orders. Nonetheless, we show that
second-order calculations now are feasible routinely. Our work thus extends
a previous study,~\cite{BiehsEtAl08} which has given first-order results,
and enables us to delineate under which conditions low-order perturbation
theory is sufficient. It also complements a recent investigation by Biehs
and Greffet~\cite{BiehsGreffet09} who have considered rough surfaces
described as stochastic Gaussian processes. In contrast, we focus on surfaces
with deliberately induced, regular nanoprofiles. We proceed as follows:
In Sec.~\ref{Sec:S_2} we sketch the underlying perturbative
scheme,~\cite{Greffet88} and outline a few details of its numerical
implementation, deferring technicalities to the Appendix~\ref{App:A}. We then
present results of our computations in Sec.~\ref{Sec:S_3}, first examining the
relative magnitude of first-and second-order contributions, and then outlining
how to quantify the resolution power of an idealized NSThM. Some conclusions
are drawn in the final Sec.~\ref{Sec:S_4}.
\section{Computation of the Green's dyadics}
\label{Sec:S_2}
In this section we utilize an analytical perturbative approach, originally
developed by Greffet for calculating the scattered electromagnetic waves above
a rough dielectric surface,~\cite{Greffet88} in order to obtain the required
Green's dyadics $\mathds{G}^{\rm E}_{\rm r}$ and $\mathds{G}^{\rm H}_{\rm r}$. Greffet's approach
results in a series of recursively determined contributions in ascending
orders of the surface profile, and thus enables one to systematically assess
the higher-order terms.
\subsection{Calculational scheme}
We assume that the surface is described by an expression $z = hf(x,y)$,
where $f(x,y)$ is a normalized function varying between $+1$ and $-1$; the
scaling parameter~$h$ carries the dimension of a length. The nonmagnetic
dielectric medium, equipped with permittivity $\epsilon(\omega)$, fills the
entire half-space $z<hf(x,y)$. For $z>hf(x,y)$, outside the dielectric,
the total electric field consists of a prescribed incident component
$\vec{E}_{\rm i}(\vec{r})$, and of the so far unknown reflected component
$\vec{E}_{\rm r}(\vec{r})$, while the transmitted field inside the dielectric is
denoted as $\vec{E}_{\rm t}(\vec{r})$. Greffet has given a recursive series
solution for the transmitted and the reflected field,~\cite{Greffet88} invoking
the extinction theorem and the Rayleigh hypothesis.~\cite{FariasMaradudin83}
The extinction theorem amounts to an exact integral formulation of the boundary
condition, whereas the use of the Rayleigh hypothesis means expanding the
transmitted, incident, and reflected fields in plane waves travelling in
$z$-direction,
\begin{equation}
\label{eq:exp_et}
\vec{E}_{\rm t}(\vec{r}) = \int \! {\rm d}^2 \kappa \,
\vec{e}_{\rm t}(\vec{\kappa}){\rm e}^{{\rm i}(\vec{\kappa}\cdot\vec{\rho}-k_z z)} \; ,
\end{equation}
\begin{equation}
\vec{E}_{\rm i}(\vec{r}) = \int \! {\rm d}^2 \kappa \,
\vec{e}_{\rm i}(\vec{\kappa}){\rm e}^{{\rm i}(\vec{\kappa}\cdot\vec{\rho}-k_{z0} z)} \; ,
\end{equation}
and
\begin{equation}
\label{eq:exp_er}
\vec{E}_{\rm r}(\vec{r}) = \int \! {\rm d}^2 \kappa \,
\vec{e}_{\rm r}(\vec{\kappa}){\rm e}^{{\rm i}(\vec{\kappa}\cdot\vec{\rho}+k_{z0} z)} \; .
\end{equation}
Here we write $\vec{r}=[x,y,z]^t$ for the position vector,
$\vec{\rho}=[x,y,0]^t$ for its lateral part, and $\vec{\kappa}=[k_x,k_y,0]^t$ for
the lateral component of the wave vector; moreover,
\begin{equation}
k_z=\sqrt{\epsilon k_0- \kappa}
\end{equation}
and
\begin{equation}
k_{z0}=\sqrt{k_0-\kappa}
\end{equation}
are the $z$-components of the wave vector inside and outside the medium. We
also use the notation $k_0 = \omega/c$ and $\kappa=|\vec{\kappa}|$. These expansions
\reff{eq:exp_et}-\reff{eq:exp_er} assume translational symmetry in the
$x$-$y$-plane and hence are strictly justified outside the structured region,
that is, for $z > h$ and $z < -h$. However, the extinction theorem requires
to evaluate the fields on the very surface of the dielectric, where the
validity of the above expressions cannot be taken for granted. Ignoring
this complication and using the expansions \reff{eq:exp_et}-\reff{eq:exp_er}
nonetheless is a common procedure~\cite{HenkelSandoghdar98,LambrechtEtAl06}
which has been looked into by several authors from the mathematical point of
view;~\cite{BergFokkema79,HugoninEtAl81,Keller00,Ramm02} it appears to work
reliably at least for sufficiently small values of~$h$. For example, in the
case of a sinusoidal grating described by $z = (h/2) \cos(2\pi x/D)$ this
Rayleigh hypothesis holds rigorously~\cite{BergFokkema79,HugoninEtAl81,Keller00}
up to $h_{\max}/D = 0.142\,521$, and may therefore be employed for both the
propagating and the evanescent parts of the field as long as the ratio $h/D$
stays below this boundary.
The field's Fourier components then are expanded in the forms
\begin{equation}
\vec{e}_{\rm t}(\vec{\kappa})=\sum\limits_{m=0}^\infty
\frac{\vec{e}^{(m)}_{\rm t}(\vec{\kappa})}{m!} \; ,
\end{equation}
\begin{equation}
\vec{e}_{\rm r}(\vec{\kappa})=\sum\limits_{m=0}^\infty
\frac{\vec{e}^{(m)}_{\rm r}(\vec{\kappa})}{m!} \; ,
\end{equation}
and
\begin{equation}
\vec{e}_{\rm i}(\vec{\kappa})=\sum\limits_{m=0}^\infty
\frac{\vec{e}^{(m)}_{\rm i}(\vec{\kappa})}{m!} = \vec{e}_{\rm i}^{(0)}(\vec{\kappa}) \; .
\end{equation}
It is useful to split the fields into their s\,- and p\,- components
according to
\begin{equation}
\vec{e}_{\rm t}(\vec{\kappa})=
e_{\rm t,s}(\vec{\kappa})\vec{a}_{\rm s}(\vec{\kappa})+e_{\rm t,p}(\vec{\kappa})\vec{a}_{\rm p,t}^-(\vec{\kappa}) \; ,
\end{equation}
\begin{equation}
\vec{e}_{\rm i}(\vec{\kappa})=
e_{\rm i,s}(\vec{\kappa})\vec{a}_{\rm s}(\vec{\kappa})+e_{\rm i,p}(\vec{\kappa})\vec{a}_{\rm p,0}^-(\vec{\kappa}) \; ,
\end{equation}
and
\begin{equation}
\vec{e}_{\rm r}(\vec{\kappa})=
e_{\rm r,s}(\vec{\kappa})\vec{a}_{\rm s}(\vec{\kappa})+e_{\rm r,p}(\vec{\kappa})\vec{a}_{\rm p,0}^+(\vec{\kappa}) \; ,
\end{equation}
where
\begin{equation}
\vec{a}_{\rm s}(\vec{\kappa})=\frac{1}{\kappa}\left[ \begin{array}{c}
-k_y \\ k_x \\ 0 \end{array} \right] \; ,
\end{equation}
\begin{equation}
\vec{a}_{\rm p,t}^-(\vec{\kappa})=-\frac{1}{n k_0 \kappa}\left[ \begin{array}{c}
k_x k_z \\ k_y k_z \\ \kappa^2 \end{array} \right] \; ,
\end{equation}
\begin{equation}
\vec{a}_{\rm p,0}^-(\vec{\kappa})=-\frac{1}{k_0 \kappa}\left[ \begin{array}{c}
k_x k_{z0} \\ k_y k_{z0} \\ \kappa^2 \end{array} \right] \; ,
\end{equation}
and
\begin{equation}
\vec{a}_{\rm p,0}^+(\vec{\kappa})=\frac{1}{k_0 \kappa}\left[ \begin{array}{c}
k_x k_{z0} \\ k_y k_{z0} \\ -\kappa^2 \end{array} \right] \; ;
\end{equation}
here $n=\sqrt{\epsilon}$ is the index of refraction.
Following Greffet,~\cite{Greffet88} one then obtains the transmitted field
in the recursive form
\begin{align}
\label{eq:rec_et}
\left[\begin{array}{c}
e_{\rm t,s}^{(m)}(\vec{\kappa}) \\ e_{\rm t,p}^{(m)}(\vec{\kappa}) \end{array}\right] & =
\frac{k_z-k_{z0}}{4 \pi^2}\mathds{R}^{-1}(\vec{\kappa},\vec{\kappa})
\int \! {\rm d}^2 \kappa' \, \Bigg\{\mathds{R}(\vec{\kappa},\vec{\kappa}')
\nonumber\\
& \times \sum\limits_{q=1}^{m} \binom{m}{q}\frac{({\rm i} h)^q
\F{q}(\vec{\kappa}'-\vec{\kappa})}{(k_{z0}-k_z')^{1-q}}
\nonumber\\
&\times\left[ \begin{array}{c}
e_{\rm t,s}^{(m-q)}(\vec{\kappa}') \\ e_{\rm t,p}^{(m-q)}(\vec{\kappa}') \end{array} \right]
\Bigg\} \; ,
\end{align}
so that $\vec{e}^{(m)}_{\rm t}(\vec{\kappa})$ is proportional to $h^m$; the according
expression for the reflected field reads
\begin{align}
\label{eq:rec_er}
\left[\begin{array}{c}
e_{\rm r,s}^{(m)}(\vec{\kappa}) \\ e_{\rm r,p}^{(m)}(\vec{\kappa}) \end{array}\right] & =
\frac{\epsilon-1}{8 \pi^2 k_{z0}} \Bigg\{
\int \! {\rm d}^2 \kappa' \, \Bigg( \mathds{P}(\vec{\kappa},\vec{\kappa}')
\nonumber\\
&\hspace{-1 cm}\times \sum\limits_{q=0}^{m-1}\binom{m}{q}
\frac{(-{\rm i} h)^{m-q}\F{m-q}(\vec{\kappa}'-\vec{\kappa})}{(k_{z0}+k_z')^{1+q-m}}
\nonumber\\
&\hspace{-1 cm}\times \left[
\begin{array}{c}
e_{\rm t,s}^{(q)}(\vec{\kappa}') \\ e_{\rm t,p}^{(q)}(\vec{\kappa}') \end{array} \right]\Bigg)
\nonumber\\
&\hspace{-1cm}+\mathds{P}(\vec{\kappa},\vec{\kappa})\frac{4 \pi^2}{k_z+k_{z0}}
\left[\begin{array}{c}
e_{\rm t,s}^{(m)}(\vec{\kappa}) \\ e_{\rm t,p}^{(m)}(\vec{\kappa}) \end{array}
\right]\Bigg\} \; .
\end{align}
In these equations the quantities $\widehat{F}^{(n)}(\vec{\kappa})$ denote the Fourier
transforms of powers of the surface function,
\begin{equation}
\widehat{F}^{(n)}(\vec{\kappa})=\int \! {\rm d}^2 \rho \,
{\rm e}^{{\rm i} \vec{\kappa} \cdot \vec{\rho}} f^n(\vec{\rho}) \; .
\end{equation}
The linear operators $\mathds{R}(\vec{\kappa},\vec{\kappa}')$ and $\mathds{P}(\vec{\kappa},\vec{\kappa}')$
effectuate the double vectorial product with $\vec{k}_{\rm r}^-=[k_x, k_y,-k_{z0}]^t$ and
$\vec{k}_{\rm r}^+=[k_x, k_y, k_{z0}]^t$, respectively; these double products (namely,
$\vec{k}_{\rm r}^- \times \vec{k}_{\rm r}^- \times$ and $\vec{k}_{\rm r}^+ \times \vec{k}_{\rm r}^+ \times$) typically
appear when using the extinction theorem. In the basis chosen, the matrix
forms of these operators are
\begin{equation}
\mathds{R}(\vec{\kappa},\vec{\kappa}')=-\frac{k_0^2}{\kappa \kappa'}
\left[ \begin{array}{c c} \vec{\kappa} \cdot \vec{\kappa}' &
-\frac{k'_z(\vec{\kappa} \times \vec{\kappa}')_z}{n k_0} \\
\frac{k_{z0}(\vec{\kappa} \times \vec{\kappa}')_z}{k_0} &
\frac{\kappa^2 \kappa'^2 + \vec{\kappa} \cdot \vec{\kappa}' k_{z0}k'_z}{n k_0^2}
\end{array}\right]
\end{equation}
and
\begin{equation}
\mathds{P}(\vec{\kappa},\vec{\kappa}')=-\frac{k_0^2}{\kappa \kappa'}
\left[ \begin{array}{c c} \vec{\kappa} \cdot \vec{\kappa}' &
-\frac{k'_z(\vec{\kappa} \times \vec{\kappa}')_z}{n k_0} \\
\frac{-k_{z0}(\vec{\kappa} \times \vec{\kappa}')_z}{k_0} &
\frac{\kappa^2 \kappa'^2 - \vec{\kappa} \cdot \vec{\kappa}' k_{z0}k'_z}{n k_0^2}
\end{array}\right] \; ,
\end{equation}
writing $(\vec{\kappa} \times \vec{\kappa}')_z$ for the $z$-component of the vectorial
product of $\vec{\kappa}$ and $\vec{\kappa}'$.
The above recursions start from the well known half-space results obtained
for a perfectly flat surface, which can be cast into the forms
\begin{equation}
\left[ \begin{array}{c}
e_{\rm t,s}^{(0)}(\vec{\kappa}) \\ e_{\rm t,p}^{(0)}(\vec{\kappa})
\end{array} \right ] = \left[ \begin{array}{cc}
t_{\rm s}(\vec{\kappa}) & 0 \\ 0 & t_{\rm p}(\vec{\kappa})
\end{array} \right] \left[ \begin{array}{c}
e_{\rm i,s}(\vec{\kappa}) \\ e_{\rm i,p}(\vec{\kappa})
\end{array} \right ]
\end{equation}
and
\begin{equation}
\left[ \begin{array}{c}
e_{\rm r,s}^{(0)}(\vec{\kappa}) \\ e_{\rm r,p}^{(0)}(\vec{\kappa})
\end{array} \right ] = \left[ \begin{array}{cc}
r_{\rm s}(\vec{\kappa}) & 0 \\ 0 & r_{\rm p}(\vec{\kappa})
\end{array} \right] \left[ \begin{array}{c}
e_{\rm i,s}(\vec{\kappa}) \\ e_{\rm i,p}(\vec{\kappa})
\end{array} \right ]
\end{equation}
with the Fresnel coefficients
\begin{eqnarray*}
t_{\rm s}(\vec{\kappa}) = \frac{2 k_{z0}}{k_{z0}+k_z} & ; &
t_{\rm p}(\vec{\kappa}) = \frac{2 n k_{z0}}{n^2 k_{z0}+k_z} \; ;
\\\
r_{\rm s}(\vec{\kappa}) = \frac{k_{z0}-k_z}{k_{z0}+k_z} & ; &
r_{\rm p}(\vec{\kappa}) = \frac{n^2k_{z0}-k_z }{n^2k_{z0}+k_z} \; .
\end{eqnarray*}
From these fields~\reff{eq:rec_et}-\reff{eq:rec_er} we now proceed to the
calculation of the Green's dyadics. More precisely, in order to compute the
local density of states we need to determine the `reflected' part of the
Green's dyadics for coinciding source and observation points. To this end, we
take the field of a delta-like source current located at $\vec{\rho}+ z\vec{e}_z$
and pointing into the direction of the unit vector $\vec{j}$ as incident field,
giving
\begin{equation}
\left[ \begin{array}{c}
e_{\rm i,s}(\vec{\kappa}) \\ e_{\rm i,p}(\vec{\kappa})
\end{array} \right] =
-\frac{\omega \mu_0}{2 k_{z0}}{\rm e}^{{\rm i}(\vec{\kappa} \cdot \vec{\rho}+k_{z0} z)}
\left[ \begin{array}{c}
\vec{a}_{\rm s}(\vec{\kappa})\cdot \vec{j} \\ \vec{a}^-_{\rm p,0}(\vec{\kappa}) \cdot \vec{j}
\end{array} \right] \; .
\end{equation}
To zeroth order in $h$, the reflected field then is
\begin{align}
\label{eq:Eref0}
\left[ \begin{array}{c}
e_{\rm r,s}^{(0)}(\vec{\kappa}) \\ e_{\rm r,p}^{(0)}(\vec{\kappa})
\end{array} \right] & =
-\frac{\omega \mu_0}{2 k_{z0}}{\rm e}^{{\rm i}(\vec{\kappa} \cdot \vec{\rho}+k_{z0} z)} \\
&\times \left[ \begin{array}{cc}
r_{\rm s}(\vec{\kappa}) & 0 \\ 0 & r_{\rm p}(\vec{\kappa})
\end{array} \right] \left[ \begin{array}{c}
\vec{a}_{\rm s}(\vec{\kappa})\cdot \vec{j} \\ \vec{a}^+_{\rm p,0}(\vec{\kappa}) \cdot \vec{j}
\end{array} \right] \; .
\nonumber
\end{align}
The directions of the s\,- and p\,-components of the incident field are given
by $\vec{a}_{\rm s}(\vec{\kappa})$ and $\vec{a}^-_{\rm p,0}(\vec{\kappa})$, while the components of the
reflected field are given by $\vec{a}_{\rm s}(\vec{\kappa})$ and $\vec{a}^+_{\rm p,0}(\vec{\kappa})$.
Therefore it is convenient to split the Fourier coefficients of the Green's
dyadics into the four parts that result from taking the dyadic products of
these unit vectors, leading to
\begin{align}
g_{\rm r}^{\rm E,(0)}(\vec{\kappa})&=g^{\rm E,(0)}_{\rm r,ss}(\vec{\kappa})\vec{a}_{\rm s}(\vec{\kappa})\otimes \vec{a}_{\rm s}(\vec{\kappa}) \\
&+g^{\rm E,(0)}_{\rm r,sp}(\vec{\kappa})\vec{a}_{\rm s}(\vec{\kappa})\otimes \vec{a}^-_{\rm p,0}(\vec{\kappa}) \nonumber\\
&+g^{\rm E,(0)}_{\rm r,ps}(\vec{\kappa})\vec{a}^+_{\rm p,0}(\vec{\kappa})\otimes\vec{a}_{\rm s}(\vec{\kappa}) \nonumber\\
&+g^{\rm E,(0)}_{\rm r,pp}(\vec{\kappa})\vec{a}^+_{\rm p,0}(\vec{\kappa})\otimes \vec{a}^-_{\rm p,0}(\vec{\kappa}) \; .
\nonumber
\end{align}
Having employed a delta-like source current the relation between the field
and the electric Green's dyadic simply reads
$\mathds{G}^{\rm E}\cdot \vec{j} = \vec{E} / ({\rm i} \omega \mu_0)$, so that the
coefficients of this dyadic can easily be read off from Eq.~\reff{eq:Eref0}.
By means of an inverse Fourier transform, equating source and observation
point, one then arrives at the familiar result for the reflected Green's
dyadic pertaining to a flat surface,
\begin{align}
\label{eq:green0}
\mathds{G}_{\rm r}^{\rm E,(0)} & = \frac{{\rm i}}{4 \pi^2}\int \! {\rm d}^2\kappa \,
\frac{{\rm e}^{2 {\rm i} k_{z0} z }}{2 k_{z0}}\big[r_{\rm s}(\vec{\kappa}) \vec{a}_{\rm s}(\vec{\kappa})
\otimes \vec{a}_{\rm s}(\vec{\kappa})
\nonumber\\
&+ r_{\rm p}(\vec{\kappa}) \vec{a}^+_{\rm p,0}(\vec{\kappa}) \otimes \vec{a}^-_{\rm p,0}(\vec{\kappa}) \big] \; .
\end{align}
With the help of Eq.~\reff{eq:rec_er} one obtains similar expressions to all
orders in the profile height~$h$. An advantage of this approach lies in the
fact that it is then quite easy to also determine the magnetic Green's dyadic,
which is related to the electric one through~\cite{FelsenMarcuvitz94}
\begin{equation}
\mathds{G}_{\rm r}^{\rm H}(\vec{r},\vec{r}) = -\frac{1}{k_0^2}\nabla \times
\mathds{G}_{\rm r}^{\rm E}(\vec{r},\vec{r}') \times \nabla' |_{\vec{r}'=\vec{r}}
\; .
\end{equation}
In Fourier space the operator $\nabla$ is replaced either by
${\rm i}(\vec{\kappa} + k_{z0} \vec{e}_z)$ or by ${\rm i}(\vec{\kappa} - k_{z0} \vec{e}_z)$, depending
on whether the curl acts on a unit vector belonging to the incident or to the
reflected field. Therefore, the magnetic Green's dyadic is derived from its
electric counterpart by simply replacing $\vec{a}(\vec{\kappa}) \otimes \vec{a}(\vec{\kappa}')$ by the
expression
$\frac{1}{k_0^2}(\vec{\kappa}+ k_{z0}\vec{e}_z) \times \vec{a}(\vec{\kappa}) \otimes \vec{a}(\vec{\kappa}')
\times (\vec{\kappa}'- k'_{z0}\vec{e}_z)$. Here we introduce $\vec{\kappa}'$ and $k'_{z0}$
because it is only to zeroth order that the reflected and the incident field
are characterized by the same wave vector.
Finally, for calculating the trace of the Green's dyadics one just has to
replace the dyadic products by their traces. Hence, the method sketched here
yields a transparent strategy for obtaining the electric and the magnetic
LDOS above a structured surface. In practice, a restriction on the maximum
order achievable is imposed by the available computational resources: To
zeroth order only two-dimensional integrals have to be evaluated, to first
order four-dimensional integrals appear; to second order one already has
to deal with six-dimensional integrals, and so on. Clearly, a good choice
of the numerical tools is decisive here; we therefore present some details
of our implementation.
\subsection{Numerical implementation}
\label{Subsec:IIB}
We exemplarily discuss the calculation of the electric local density of states
to first and second order; its magnetic counterpart, and the higher-order terms,
are determined in a similar way. With the help of Eq.~\reff{eq:rec_er} the
trace of the first-order contribution to the electric reflected Green's dyadic
is computed as
\begin{equation}
\label{eq:trace1}
{\rm Tr}\big\{\mathds{G}_{\rm r}^{\rm E,(1)}\big\} =
-\int\! \frac{{\rm d}^2 q}{ 4 \pi^2} \,
{\rm e}^{{\rm i} \vec{q} \cdot \vec{\rho}} \widehat{F}^{(1)}(-\vec{q}) a_1(\vec{q}) \; ,
\end{equation}
where $a_1(\vec{q})$ is given by the integral
\begin{equation}
\label{eq:S1}
a_1(\vec{q})=\int \! {\rm d}^2 \kappa' \, \vec{S}^{(1)}(\vec{q}+\vec{\kappa}',\vec{\kappa}')\cdot
\vec{A}^{\rm (E)}_{\rm tr}(\vec{q}+\vec{\kappa}',\vec{\kappa}')
\end{equation}
with the four-dimensional vectors $\vec{S}^{(1)}$ and $\vec{A}^{\rm (E)}_{\rm tr}$
specified in Appendix~\ref{App:A}. We either employ an experimentally
determined surface profile,~\cite{KittelEtAl08} or some suitably selected
model function $f(\vec{\rho})$; sample it, and perform a discrete FFT in order
to determine $\F{1}$. Then we compute $a(\vec{q})$ for each required $\vec{q}$ by
numerical integration, and finally take an inverse FFT of $\F{1}(-\vec{q})a(\vec{q})$
to get the trace~(\ref{eq:trace1}) of the first-order Green's dyadic.
To second order one has to deal with two contributions, one containing $\F{2}$,
the other feeding from two factors $\F{1}$. The former contribution has the
same structure as the first-order term,
\begin{align}
\label{eq:trace2a}
{\rm Tr}\big\{\mathds{G}_{\rm r,1}^{\rm E,(2)}\big\}=
\int\! \frac{{\rm d}^2 q}{ 4 \pi^2}\,
{\rm e}^{{\rm i} \vec{q} \cdot \vec{\rho}}\F{2}(-\vec{q})a_2(\vec{q}),
\end{align}
with
\begin{equation}
a_2(\vec{q})= \int \! {\rm d}^2 \kappa'\,
\vec{S}^{(2)}_1(\vec{q}+\vec{\kappa}',\vec{\kappa}') \cdot
\vec{A}^{\rm (E)}_{\rm tr}(\vec{q}+\vec{\kappa}',\vec{\kappa}') \;,
\end{equation}
and is evaluated in the same manner. The other contribution contains
a further integral,
\begin{align}
\label{eq:trace2b}
{\rm Tr}\big\{\mathds{G}_{\rm r,2}^{\rm E,(2)}\big\} & =
\int\! \frac{{\rm d}^2 q}{4 \pi^2} \,
{\rm e}^{{\rm i} \vec{q} \cdot \vec{\rho}} \F{1}(-\vec{q})
\int\! \frac{{\rm d}^2 q'}{4 \pi^2}\\
& \times \, {\rm e}^{{\rm i} \vec{q}' \cdot \vec{\rho}} \F{1}(-\vec{q}') a_3(\vec{q},\vec{q}') \, ,
\nonumber
\end{align}
with integration variables $\vec{q}=\vec{\kappa}-\vec{\kappa}'$ and $\vec{q}'=\vec{\kappa}'-\vec{\kappa}''$,
and with the expression
\begin{align}
\label{eq:S2}
a_3(\vec{q},\vec{q}')& =\int\! {\rm d}^2\kappa'' \,
{\rm e}^{{\rm i}(k_{z0}+k_{z0}'')z} \\
& \times\vec{S}^{(2)}_2(\vec{\kappa},\vec{\kappa}',\vec{\kappa}'')\cdot
\vec{A}^{\rm (E)}_{\rm tr}(\vec{\kappa},\vec{\kappa}'') \; ;
\nonumber
\end{align}
again, $\vec{S}^{(2)}_1$ and $\vec{S}^{(2)}_2$ are stated explicitly in
Appendix~\ref{App:A}. After computing $a_3(\vec{q},\vec{q}')$ on a discrete mesh of
$\vec{q}$ and $\vec{q}'$, a four-dimensional inverse FFT is performed for determining
the resulting contribution to the trace of the Green's dyadic.
The integrals are numerically evaluated using Cuba routines;~\cite{Cuba}
the Fourier transforms are executed by means of the FFTW library.~\cite{FFTW}
\section{Results}
\label{Sec:S_3}
In this section we discuss some numerical results for the local density
of states above example topographies, calculated to second order in the
profile height. We first consider the relative magnitude of the individual
contributions, in order to estimate under which conditions the termination
of the perturbation series can be justified. We then use the second-order
data to discuss the resolution power of an idealized NSThM in a mode of
operation in which the total heat transfer is kept constant while scanning
a sample's surface.
\subsection{Magnitude of second-order contributions}
Our basic model structure is a bar with height~$h$ and smoothed edges placed
on an otherwise perfectly planar surface, infinitely extended in $y$-direction
and possessing the width~$w$ in $x$-direction, as described by the function
\begin{equation}
\label{eq:fermi}
h f_1(x)=h \frac{1}{\exp\big[\zeta(|x|-0.5w) \big]+1} \; .
\end{equation}
Our calculations are done for $h=5\, \text{nm}$, $w=30 \, \text{nm}$, and
inverse smoothing length $\zeta=10^9 \, \text{m}^{-1}$. For comparison, we
also consider the somewhat rounder profile
\begin{equation}
\label{eq:rund}
h f_1(x) = h \exp\!\left(-\frac{1}{1 - (x/v)^2} +1 \right) \; ,
\end{equation}
with~$v$ adjusted such that the respective areas under the two
functions~(\ref{eq:fermi}) and~(\ref{eq:rund}) coincide. The resulting
profile shapes are drawn in Fig.~\ref{Fig:F_1}.
\begin{figure}[tb]
\epsfig{file=Order2_Fig1.eps, width=0.4\textwidth}
\caption{(Color online) Profile function~(\ref{eq:fermi}) with parameters
as employed in our model calculations, height $h=5\, \text{nm}$,
width $w=30 \, \text{nm}$, and inverse smoothing length
$\zeta=10^9 \, \text{m}^{-1}$ (full line), together with the
reference profile~(\ref{eq:rund}) (dashed). In either case, the
dielectric properties of the sample are given by the Drude model
with parameters for gold at $300 \, \text{K}$.}
\label{Fig:F_1}
\end{figure}
Because these profiles depend on only one variable, the Fourier transforms
$\F{n}$ contain delta functions, so that the integrals over $q$ and $q'$
in Eqs.~\reff{eq:trace1}, \reff{eq:trace2a}, and \reff{eq:trace2b} become
effectively one-dimensional, drastically reducing the numerical effort.
The profiles are discretized with a stepsize of $1\, \text{nm}$, covering a
total range of $500\, \text{nm}$; we have checked that the numerical results
thus obtained are stable against further reduction of the grid size to
$0.5\, \text{nm}$. We assume that the dielectric function $\epsilon(\omega)$
of the samples is given by the Drude model~\cite{AshcroftMermin76}
\begin{equation}
\varepsilon(\omega) =
1 - \frac{\omega_{\rm p}^2}{\omega^2 + {\rm i}\gamma\omega}
\end{equation}
with plasma frequency $\omega_{\rm p} = 1.4 \times 10^{16} \, \text{s}^{-1}$
and inverse relaxation time $\gamma = 3.3 \times 10^{13} \, \text{s}^{-1}$,
describing gold at the temperature $T = 300$~K.
\begin{figure}[bt]
\epsfig{file=Order2_Fig2.eps, width=0.4\textwidth}
\caption{(Color online) Zeroth-, first-, and second-order contributions to the
electric~(a) and to the magnetic~(b) part of the LDOS above a gold
surface structured with the nanobar~(\ref{eq:fermi}) as sketched in
Fig.~\ref{Fig:F_1}, for $\omega =10^{14}\, \text{s}^{-1}$ at an
observation height of $10 \, \text{nm}$ above the base substrate
plane. Note the different scales.}
\label{Fig:F_2}
\end{figure}
In Fig.~\ref{Fig:F_2} we depict the zeroth-, first-, and second-order
contributions to the electric and to the magnetic part of the LDOS for
the structure~(\ref{eq:fermi}), evaluated at a constant height of
$10 \, \text{nm}$ above the base plane for $\omega=10^{14}\, \text{s}^{-1}$,
roughly equal to the dominant thermal frequency at $300 \, \text{K}$.
The second-order terms qualitatively show the same behavior as the first-order
ones,~\cite{BiehsEtAl08} but there is a notable difference between the electric
and the magnetic part: The second-order contribution to the magnetic LDOS at
least is smaller than its first-order precessor, although only by a factor
which is not small compared to unity, whereas the magnitude of the second-order
electric contribution remains comparable to that of the first-order one, and
even slightly exceeds it. At the bar's center, the first- and the second-order
electric contributions amount to roughly 1.5 times the zero-order term. This
behavior is not accidental; it can be quantitatively understood with the help
of an elementary consideration: At sufficiently short distances, that is,
for~$z-h$ not too large compared to the profile width~$w$, the local
geometry equals that of a flat surface, shifted by~$h$ against the base plane.
Therefore, the electric LDOS $D^{\rm E}$ at such a point $\vec{r}=[0,0,z]^t$
is approximately given by the LDOS $D^{\rm E}_{\rm fs}$ pertaining to a
perfectly flat surface~\cite{JoulainEtAl03} through the relation
\begin{equation}
D^{\rm{E}}(\vec{r}) \approx D^{\rm E}_{\rm fs}([0,0,z-h]^t) \; .
\label{eq:local}
\end{equation}
Now the distance dependence of $D^{\rm E}_{\rm fs}$ in the near field is
given by
\begin{equation}
D^{\rm E}_{\rm fs}([0,0,z]^t) \sim \frac1{z^3} \; ;
\label{eq:eldos}
\end{equation}
the strong decay of this electric LDOS with the third power of the distance
clearly aids the local approximation~(\ref{eq:local}). Thus, one has
\begin{equation}
D^{\rm{E}}([0,0,z]^t)\sim\frac1{(z-h)^3} \; .
\end{equation}
Expanding in powers of $h$, this yields
\begin{equation}
D^{\rm{E}}([0,0,z]^t) \sim \frac1{z^3} + \frac{3 h}{z^4}
+ \frac{6 h^2}{z^5} + O(h^3) \; ,
\end{equation}
allowing one to estimate the ratios of the contributions appearing in
different orders:
\begin{equation}
\frac{D^{\rm{E},(1)}}{D^{\rm{E},(0)}} \approx 3\frac{h}{z} \; ;
\quad
\frac{D^{\rm{E},(2)}}{D^{\rm{E},(0)}} \approx 6\frac{h^2}{z^2} \; .
\end{equation}
With $h/z = 0.5$, which is the value considered in Fig.~\ref{Fig:F_2}, one
obtains $D^{\rm{E},(1)} / D^{\rm{E},(0)} = D^{\rm{E},(2)} / D^{\rm{E},(0)}
= 1.5$, in quite good agreement with the exact numerical result. Generalizing
this argument to any order~$n$, one observes
\begin{equation}
\frac{D^{{\rm E},(n+1)}}{D^{{\rm E},(n)}} \approx
\frac{n+3}{n+1} \, \frac{h}{z} \; .
\label{eq:este}
\end{equation}
Thus, while the series may still converge for any $z > h$, convergence at
short distances would be rather slow; for $h/z = 0.5$ the magnitude of the
leading successive contributions, normalized to the zeroth-order one, is
$1 : 3/2 : 3/2 : 5/4 : 15/16 : 21/32 : \ldots \; .$ In this example,
terminating the perturbation series at the second order means that one
collects only about half of the exact value.
In the case of the magnetic LDOS, one has
\begin{equation}
D^{\rm H}_{\rm fs}([0,0,z]^t) \sim \frac1{z} \; ,
\end{equation}
so that here a larger area of the sample's surface contributes to the LDOS than
in the electric case~(\ref{eq:eldos}), implying that a local approximation
analogous to Eq.~(\ref{eq:local}) cannot be expected to work as well as before.
Ignoring this restriction and performing the analysis nonetheless, one ends up
with
\begin{equation}
\frac{D^{{\rm H},(n+1)}}{D^{{\rm H},(n)}} \approx
\frac{h}{z} \; ,
\label{eq:esth}
\end{equation}
which, in view of the shaky foundation of the reasoning, still works
satisfactorily, capturing both the correct trends and the orders of magnitude
read off from Fig.~\ref{Fig:F_2}.
It is evident that these general findings do {\em not\/} depend on the
specific dielectric properties of the material. Indeed, when considering
a polar sample with a permittivity described by the Reststrahlen
formula~\cite{Klingshirn05}
\begin{equation}
\epsilon(\omega) = \epsilon_{\infty}
\left(1+\frac{\omega_{\rm L}^2 - \omega_{\rm T}^2}
{\omega_{\rm T}^2 - \omega^2 - {\rm i}\gamma\omega}\right)
\end{equation}
and inserting parameters appropriate for gallium nitride~\cite{Adachi04},
namely, $\epsilon_{\infty} = 5.35$ for the high-frequency permittivity,
$\omega_{\rm L} = 1.41 \times 10^{14}\,\rm{s}^{-1}$ and
$\omega_{\rm T} = 1.06 \times 10^{14}\,\rm{s}^{-1}$ for the frequencies
of the longitudinal and transversal phonons, and
$\gamma=1.51 \times 10^{12}\,\rm{s}^{-1}$ for the relaxation rate, and
again taking the profile~(\ref{eq:fermi}), we obtain Fig.~\ref{Fig:F_3},
which shows the same qualitative features as the previous Fig.~\ref{Fig:F_2}
for the gold sample, although, of course, the scales are quite different;
now the electric contribution dominates. Likewise, the results do not seem
to depend sensitively on the precise form of the structure: The corresponding
data obtained for the reference profile~(\ref{eq:rund}) are remarkably similar
to the previous ones, as shown in Fig.~\ref{Fig:F_4}, and again confirm the
simple estimates (\ref{eq:este}) and (\ref{eq:esth}). For this reason, we only
consider the gold nanobar~(\ref{eq:fermi}) in the following.
\begin{figure}[tb]
\epsfig{file=Order2_Fig3.eps, width=0.4\textwidth}
\caption{(Color online) As Fig.~\ref{Fig:F_2}, but for a sample consisting
of gallium nitride (GaN).}
\label{Fig:F_3}
\end{figure}
\begin{figure}[Hhbt]
\epsfig{file=Order2_Fig4.eps, width=0.4\textwidth}
\caption{(Color online) As Fig.~\ref{Fig:F_2}, for a gold sample with the
reference profile~(\ref{eq:rund}).}
\label{Fig:F_4}
\end{figure}
To conclude the above discussion, in the cases studied so far the
restriction to second-order perturbation theory already seems questionable,
although the strong dominance of the magnetic LDOS for metallic samples might
still mask the problem with the electric one. This is potentially important
for NSThM-applications, where typical probe-sample distances range down to
a few nanometers. On the other hand, low-order perturbation theory may be
expected to work reliably when the profile height~$h$ clearly is the smallest
length scale of the problem; according to the above reasoning, it should
become better when increasing the observation distance~$z$. In order to
estimate the smallest~$z$ at which second-order perturbation theory might
still give quantitatively good results for our model profile~(\ref{eq:fermi}),
we plot in Fig.~\ref{Fig:F_5} the ratios of the various contributions,
evaluated above the bar's center at varying distance; here we also consider
the regime $z - h \gg w$ where the short-distance estimates (\ref{eq:este})
and (\ref{eq:esth}) may no longer be taken for granted. As a rough guideline,
one may accept the truncation of the perturbation series at the second order
if the ratio of the second-order contribution to the zeroth-order one drops
below $10\%$, say. For the dominant magnetic part this criterion is satisfied
for $z > 20 \, \text{nm}$, while the electric part then requires
$z > 41 \, \text{nm}$. (When reducing the acceptance limit to $5\%$, one
gets $z > 26 \, \text{nm}$ and $z > 56 \, \text{nm}$, respectively.) The
potential problem of slow convergence here is expressed by the fact that the
ratio of second- to first-order contribution decreases only rather slowly with
increasing~$z$. Still, in view of the relative smallness of the electric LDOS
above metallic samples this does not seem to be essential.
\begin{figure}[Hhbt]
\epsfig{file=Order2_Fig5.eps, width=0.4\textwidth}
\caption{(Color online) Ratios of the zeroth-, \mbox{first-,} and second-order
contributions to the electric~(a) and to the magnetic~(b) LDOS, for
$\omega =10^{14}\, \text{s}^{-1}$ at an observation point $x = 0$
above the center of the bar~(\ref{eq:fermi}).}
\label{Fig:F_5}
\end{figure}
\subsection{Resolution of an NSThM}
Scanning the LDOS at a constant height above the base plane, as done
numerically in Figs.~\ref{Fig:F_2}, \ref{Fig:F_3}, and \ref{Fig:F_4},
corresponds to the {\it constant height}
mode of operation of an NSThM. This is not an advantageous mode, for two
practical reasons: On the one hand, it is hard to realize with sufficient
accuracy, on the other, it contains the risk of a probe-sample collision
when scanning a surface with an unknown topography, almost always resulting
in irreparable damage to the delicate sensor.~\cite{WischnathEtAl08} A much
more favorable mode avoiding these complications is the {\it constant transfer}
mode, meaning that the sensor height is continuously regulated such that the
detected heat current remains constant when moving the sensor over the
surface; the information about the sample's near field then is embodied in
the recorded sensor height. Note that this latter mode differs from the
{\it constant distance} mode, which uses additional information on the local
distance of the sensor from the structured surface (obtained by electron
tunneling spectroscopy) in order to keep that distance constant.
That constant distance mode was employed experimentally by Kittel
{\it et al.};~\cite{KittelEtAl08} numerical first-order results pertaining
to this mode can be found in Ref.~\onlinecite{BiehsEtAl08}. In contrast,
a major benefit of the constant transfer mode lies in the fact that it
exclusively requires heat-transfer information, so that there is no need
to retain the sensor's ordinary STM-capability. Moreover, it allows one to
assess the resolution limit of the NSThM in a simple manner.
For modeling this constant transfer mode, we select some appropriate fixed
value of the LDOS, and then calculate that observation distance~$a$ at which
this LDOS-value is reached. Only the sum of all contributions is of interest
now. For consistency, we also require that the second- to zeroth-order ratio
remains less than $10\%$ for the magnetic LDOS, as the electric one does not
contribute significantly here. Discussing the NSThM's resolution power first
requires the specification of a norm structure containing the length scale~$s$
to be resolved. Here we take two parallel gold bars of the
form~(\ref{eq:fermi}), as described by the profile
\begin{align}
\label{eq:2bars}
hf_2(x) = &
h\Bigg[\frac{1}{\exp\big[\zeta_1(|x+0.5s|-0.5w_1) \big]+1} \\
& +\frac{1}{\exp\big[\zeta_2(|x-0.5s|-0.5w_2) \big]+1}\Bigg] \; ;
\nonumber
\end{align}
the length scale in question is their separation~$s$. For our
matter-of-principle calculations we again choose $h=5\,\text{nm}$, together
with $w_1=w_2=30 \, \text{nm}$ and $\zeta_1=\zeta_2=10^9\,\text{m}^{-1}$.
\begin{figure}[tb]
\epsfig{file=Order2_Fig6a.eps, width=0.4\textwidth}
\epsfig{file=Order2_Fig6b.eps, width=0.4\textwidth}
\caption{(Color online) (a) Observation distance above the one-bar gold sample
described by Eq.~\reff{eq:fermi} (1 bar), and above the sample with
two parallel bars specified by Eq.~\reff{eq:2bars} (2 bars), determined
such that the second-order LDOS for $\omega=10^{14}\, \text{s}^{-1}$
constantly keeps that value which is attained for the distance
$a_{\rm eff}=20\,\text{nm}$ far away from the bars.
(b) Definition of the quantities $d$ and $\Delta d$ employed for
discussing the resolution of an ideal NSThM.}
\label{Fig:F_6}
\end{figure}
In Fig.~\ref{Fig:F_6}(a) we display second-order results for both the
one-bar geometry, and for the two-bar structure with bar separation
$s = 50\,\text{nm}$. Here the observation distance~$a$ is computed such
that the LDOS remains fixed at the value attained for the distance
$a_{\rm eff} = 20 \, \text{nm}$ above the base plane at positions far away
from the bars, always assuming $\omega=10^{14}\, \text{s}^{-1}$.
In order to discuss the resolution of an idealized NSThM, we make two further
assumptions: First, we propose that the sensor is point-like, so that no
effects due to the real sensor's extension are considered, implying that
we aim at the sensor-independent best possible resolution limit. In reality,
the finite sensor size will lead to a lower resolution. Second, we assume
that the signal recorded by the device is proportional to the LDOS at the
dominant thermal frequency, which actually appears to be quite a good
approximation for metallic samples.~\cite{KittelEtAl08} We then take the
ratio $\Delta d/d$, where $d$ is the maximum difference $a-a_{\rm eff}$
encountered above each of the two identical bars, and $\Delta d$ denotes the
difference between that maximum distance and the minimum distance adopted
between the bars, as illustrated in Fig.~\ref{Fig:F_6}(b). We now stipulate
that the two bars can be resolved if $\Delta d /d \geq r$; this number~$r$
characterizes the sensitivity of the respective experimental set-up.
\begin{figure}[tb]
\epsfig{file=Order2_Fig7.eps, width=0.4\textwidth}
\caption{(Color online) Ratio $\Delta d/d$ for three different values of
$a_{\rm eff}$ ($20 \, \text{nm}$, $30 \, \text{nm}$, and
$40 \, \text{nm}$), as functions of the bar separation~$s$.}
\label{Fig:F_7}
\end{figure}
In Fig.~\ref{Fig:F_7} we plot the ratio $\Delta d/d$ for effective distances
$a_{\rm eff}=20 \, \text{nm}$, $30 \, \text{nm}$, and $40 \, \text{nm}$,
as functions of the bar separation~$s$. If we take $r = 0.2$ for the sake
of discussion, the resolvable minimum distances are
$s_{\min}\approx 50 \, \text{nm}$ for $a_{\rm eff}=20 \, \text{nm}$,
$s_{\min}\approx 65 \, \text{nm}$ for $a_{\rm eff}=30 \, \text{nm}$, and
$s_{\min}\approx 80 \, \text{nm}$ for $a_{\rm eff}=40 \, \text{nm}$.
We emphasize that these figures serve to illustrate the basic principle
and should not be taken at face value. The resolution achievable with an
actual NSThM device will also depend on the type of surface structure under
investigation; a further limit will be imposed by the finite sensor volume.
The key message, however, stands out clearly: When scanning an isothermal,
nanostructured surface with a Near-Field Scanning Thermal Microscope,
one is able to resolve structures with linear extensions which may fall
orders of magnitude below the scale set by the dominant thermal wave
length.~\cite{KittelEtAl08}
\section{Conclusions}
\label{Sec:S_4}
We have demonstrated that a numerical evaluation of Greffet's
perturbation series for the scattered electromagnetic field at rough
surfaces~\cite{Greffet88} is routinely feasible up to second order. This
allows one to evaluate the local density of states above surfaces with
arbitrary profiles, thus lifting the restriction to the limited class of
profiles which can be dealt with analytically.
The convergence properties of this series seem to warrant further analysis.
While one may reasonably guess that low-order perturbation theory should be
sufficient when the profile height~$h$ is by far the smallest length scale
of the problem, the slow decrease of the successive contributions to the
electric LDOS depicted in Fig.~\ref{Fig:F_5}(a) for the smoothed, but still
quite steep metallic model structure sketched in Fig.~\ref{Fig:F_1}, together
with the elementary estimates based on Eq.~(\ref{eq:local}), are warning signs.
While our results have been obtained for specific model profiles, it would be
quite helpful to have mathematically rigorous and sharp upper bounds on the
higher-order contributions for any given surface structure.
We did not discuss possible effects due to the nonlocal dielectric response
of the sample, which might come into play at distances of a few
nanometers.~\cite{HenkelJoulain06} The question whether such effects would
be detectable with a Near-Field Scanning Thermal Microscope (NSThM)
deserves further investigations.
With respect to NSThM surface imaging, we have shown how to estimate the
best possible resolution limit, attained for a point-like sensor. Here we
have introduced a mode of operation characterized by constant heat transfer,
giving access to isolines of the LDOS. It is now a major task to extend the
preliminary studies reported in Ref.~\onlinecite{KittelEtAl08}, considering
surfaces with both regular and random nanostructures, and to further explore
the concept of near-field thermal imaging.
\begin{acknowledgments}
We wish to thank Achim Kittel, Uli F.\ Wisch\-nath, David Hellmann,
Lars Hoelzel, Ludwig Worbes, and J\"urgen Parisi for continuing
discussions of their experiments.
This work was supported by the Deutsche Forschungs\-gemeinschaft through
Grant No.\ KI 438/8-1.
Computer power was obtained from the GOLEM~I cluster of the Universit\"at
Oldenburg.
S.-A.~B.\ gratefully acknowledges support from the Deutsche Akademie der
Naturforscher Leopoldina under Grant No.\ LPDS 2009-7.
\end{acknowledgments}
\begin{appendix}
\begin{widetext}
\section{Calculational details}
\label{App:A}
In this Appendix we state the precise forms of the expressions which have
been used in Subsection~\ref{Subsec:IIB}. The vector $\vec{S}^{(1)}$
introduced in Eq.~(\ref{eq:S1}), required for computing the first-order
electric contribution~(\ref{eq:trace1}), contains the products of the
transmission coefficients $t_{\rm s}(\vec{\kappa})=(2 k_{z0})(k_{z0}+k_z)^{-1}$ and
$t_{\rm p}(\vec{\kappa})=(2 \sqrt{\epsilon} k_{z0})(\epsilon k_{z0}+k_z)^{-1}$, together
with a convenient prefactor:
\begin{equation}
\vec{S}^{(1)}(\vec{\kappa},\vec{\kappa}') =
\frac{k_0^2}{\kappa \kappa'}\frac{\epsilon-1}{16 \pi^2}
\frac{{\rm e}^{-{\rm i}(k_{z0}+k_{z0}')z}}{k_{z0}k_{z0}'}
\left[ \begin{array}{c}
t_{\rm s}(\vec{\kappa}) t_{\rm s}(\vec{\kappa}') \vec{\kappa} \cdot \vec{\kappa}' \\
-t_{\rm s}(\vec{\kappa}) t_{\rm p}(\vec{\kappa}') \frac{k_z'}{n k_0}(\vec{\kappa} \times \vec{\kappa}')_z \\
-t_{\rm p}(\vec{\kappa})t_{\rm s}(\vec{\kappa}')\frac{k_z}{n k_0}(\vec{\kappa} \times \vec{\kappa}')_z \\
\frac{t_{\rm p}(\vec{\kappa})t_{\rm p}(\vec{\kappa}')}{n^2 k_0^2}
[n^2 \kappa^2 \kappa'^2-k_z k_z' (\vec{\kappa} \cdot \vec{\kappa}')]
\end{array} \right] \; .
\end{equation}
The other vector $\vec{A}^{\rm (E)}_{\rm tr}$ appearing in Eq.~(\ref{eq:S1})
contains the traces of the dyadic products,
\begin{equation}
\vec{A}^{\rm (E)}_{\rm tr}(\vec{\kappa}, \vec{\kappa}') = \frac{1}{\kappa \kappa'}
\left[ \begin{array}{c}
\vec{\kappa} \cdot \vec{\kappa}'\\
-\frac{k_{z0}'}{k_0}(\vk \times \vk')_z \\
-\frac{k_{z0}}{k_0}(\vk \times \vk')_z\\
\frac{1}{k_0^2}(\kappa^2\kappa'^2-k_{z0}k_{z0}'(\vk \cdot \vk'))
\end{array} \right] \; .
\end{equation}
For computing the magnetic contribution, this vector has to be replaced by
\begin{equation}
\vec{A}^{\rm (H)}_{\rm tr}(\vec{\kappa}, \vec{\kappa}') = \frac{1}{\kappa \kappa'}
\left[ \begin{array}{c}
\frac{1}{k_0^2}(\kappa^2\kappa'^2-k_{z0}k_{z0}'(\vk \cdot \vk'))\\
\frac{k_{z0}}{k_0}(\vk \times \vk')_z \\
\frac{k_{z0}'}{k_0}(\vk \times \vk')_z\\
\vec{\kappa} \cdot \vec{\kappa}'
\end{array} \right] \; .
\end{equation}
The vector $\vec{S}^{(2)}_1(\vec{\kappa},\vec{\kappa}')$ determining the second-order
term~(\ref{eq:trace2a}) is given by
\begin{equation}
\vec{S}^{(2)}_1(\vec{\kappa},\vec{\kappa}') =
{\rm i} \frac{k_0^2(\epsilon-1)}{16 \pi^2 \kappa \kappa'}
\frac{{\rm e}^{-{\rm i}(k_{z0}+k_{z0}')z}}{k_{z0}k_{z0}'}
\left[ \begin{array}{c}
t_{\rm s}(\vk) t_{\rm s}(\vk') (k_z + k_z') (\vk \cdot \vk')\\
\\
-t_{\rm s}(\vk) t_{\rm p}(\vk') \frac{k_z'}{n k_0}(k_z+k_z')(\vk \times \vk')_z\\
\\
-t_{\rm p}(\vk) t_{\rm s}(\vk') \frac{1}{nk_0}(k_z k_z' +n^2 k_{z0}^2)(\vk \times \vk')_z\\
\\
t_{\rm p}(\vk) t_{\rm p}(\vk') \frac{1}{n^2 k_0^2}\bigg[ \kappa^2 \kappa'^2 (n^2 k_z'+k_z)\\
- (\vk \cdot \vk') k_z' (k_zk_z' + n^2 k_{z0}^2) \bigg]
\end{array} \right] \; ,
\end{equation}
whereas the vector $\vec{S}^{(2)}_2(\vec{\kappa},\vec{\kappa}',\vec{\kappa}'')$ entering into the
expression~(\ref{eq:S2}), and thus into the other second-order
contribution~(\ref{eq:trace2b}), takes the form
\begin{equation}
\vec{S}^{(2)}_2(\vec{\kappa},\vec{\kappa}',\vec{\kappa}'') =
-{\rm i} \frac{k_0^2 (\epsilon-1){\rm e}^{-{\rm i}(k_{z0}+k_{z0}'')z}}{8 \pi^2\kappa \kappa'^2 \kappa''}
\frac{k_z'-k_{z0}'}{k_{z0}''k_{z0}}
\left[ \begin{array}{c}
t_{\rm s}(\vk) t_{\rm s}(\vk'') \bigg[(\vk \cdot \vk') (\vk' \cdot \vk'') \\
-\frac{k_z' k_{z0}'}{\kappa'^2+k_z' k_{z0}'}(\vk \times \vk')_z (\vk' \times \vk'')_z \bigg]\\
\\
-t_{\rm s}(\vk) t_{\rm p}(\vk'') \bigg[\frac{k_z''}{n k_0} (\vk \cdot \vk') (\vk' \times \vk'')_z \\
+ \frac{k_z'}{n k_0} \frac{\kappa'\kappa''+k_{z0}'k_z''(\vk' \cdot \vk'')}
{\kappa'^2+k_{z0}'k_z'}(\vk \times \vk')_z \bigg]\\
\\
-t_{\rm p}(\vk) t_{\rm s}(\vk'') \bigg[ \frac{k_z}{n k_0} (\vk \times \vk')_z (\vk' \cdot \vk'')\\
+ \frac{k_{z0}' (\vk' \times \vk'')_z}{n k_0}
\frac{n^2 \kappa^2 \kappa'^2-k_z k_z'(\vk \cdot \vk')}{\kappa'^2+k_{z0}'k_z'}
\bigg] \\
\\
t_{\rm p}(\vk) t_{\rm p}(\vk'') \bigg[ \frac{k_z'' k_z}{n^2 k_0^2}(\vk \times \vk')_z (\vk' \times \vk'')_z\\
+\frac{1}{n^2 k_0^2} \frac{\kappa'^2 \kappa''^2 + k_{z0}'k_z''(\vk' \cdot \vk'')}
{\kappa'^2+k_{z0}'k_z'}\big[ n^2 \kappa^2 \kappa'^2 - k_z k_z' (\vk \cdot \vk')
\big] \bigg] \end{array} \right] \; .
\end{equation}
\end{widetext}
\end{appendix}
|
\section{Introduction}
With the sizes of online music and media databases growing to millions
and billions of items, users need tools for searching and browsing
these items in intuitive ways. One approach that has proven to be
popular is the use of social tags \cite{lamere08}, short descriptions
applied by users to items. Users can search and browse through a
collection using the tags that they or others have applied. This
system works well for popular items that have been tagged by many
users, but fails for items that are new or niche, this is the so-called cold
start problem \cite{ScheinEtAl2002}.
One promising way to overcome the cold start is through content-based
analysis and tagging of the items in the collection, known as
autotagging. Researchers have investigated a number of autotagging
techniques for music over the last decade \cite{whitman02, eck08,
tingle10}. While a few autotagging techniques attempt to capture
the relationship between tags (e.g.\ \cite{bertinmahieux08}), many
treat each tag as a separate classification or ranking problem (e.g.\
\cite{mandel08b}). The problem of predicting the presence or
relevance of multiple tags simultaneously is known as the
multi-label classification problem \cite{trohidis+etal:2008}.
This paper explores techniques for autotagging music that incorporate
the relationships between tags. We approach this problem in two ways,
both of which are based on conditional restricted Boltzmann machines (RBMs)
described in Section~\ref{sec:rbms}. The first approach, described in
Section~\ref{sec:smoothing}, is a novel model
trained to predict the tags that a user will apply to music based on
the tags other users have applied to it. It is a purely textual model
in that it does not utilize the audio at all to make
predictions. These predicted tags, which we call ``smoothed'' tags,
are then used to train different types of classifiers that do utilize
audio.
The second approach, described in
Section~\ref{sec:drbm}, is a discriminative RBM
\cite{Larochelle+Bengio-2008}, which learns to jointly predict tags
from features extracted from the audio. We extend the discriminative
RBM to perform multi-label classification instead of the
winner-take-all classification performed by previous discriminative
RBMs. This new model requires a new training algorithm. We explore
four techniques for approximating the gradient of the model parameters, namely maximum
likelihood using contrastive divergence, maximum
pseudo-likelihood,
mean-field contrastive divergence, and loopy belief propagation
approximations.
Section~\ref{sec:experiments} investigates the performance of these
two methods separately and together on three different datasets, two
of which have been previously described in the literature, and one of
which (the largest of the three) is new and has not been used to train or test
autotaggers before.
\section{Restricted Boltzmann machines} \label{sec:rbms}
This section describes the restricted Boltzmann machine (RBM)
\cite{smolensky86}, its conditional variant the conditional RBM, and
one particular type of conditional RBM, the discriminative RBM.
The RBM is an
undirected graphical model that generatively models a set of input
variables $\mathbf{y}=(y_1,\ldots,y_C)^T$ with a set of hidden variables
$\mathbf{h}=(h_1,\ldots,h_n)^T$. Both $\mathbf{y}$ and $\mathbf{h}$ are typically binary,
although other distributions are possible. The model is
``restricted'' in that the dependency between the hidden and visible
variables is bipartite, meaning that the hidden variables are
independent when conditioned on the visible variables and vice versa.
The joint probability density function is
\begin{equation}
p(\mathbf{y},\mathbf{h}) = \frac{1}{Z} e^{-E(\mathbf{y},\mathbf{h})}
\end{equation}
where
\begin{equation}
E(\mathbf{y},\mathbf{h}) = -\mathbf{h}^T U \mathbf{y} - c^T \mathbf{h} - d^T \mathbf{y},
\end{equation}
\begin{equation}
Z = \sum_{\mathbf{y},\mathbf{h}} e^{-E(\mathbf{y},\mathbf{h})}, \label{eq:partition}
\end{equation}
$U$ is a matrix of real numbers, and $c$ and $d$ are vectors of real
numbers.
The computation of $Z$, known as the partition function, is
intractable due to the number of terms being exponential in
the number of units.
The marginal of $y$, however, is $p(\mathbf{y}) = e^{-\mathcal{F}(\mathbf{y})}/Z$, where
$\mathcal{F}(\mathbf{y})$ is the free energy of $\mathbf{y}$ and can be easily
computed as
\begin{equation}
\mathcal{F}(\mathbf{y}) = - \log \sum_\mathbf{h} e^{-E(\mathbf{y},\mathbf{h})} = -d^T \mathbf{y} - \sum_i \log
(1+e^{c_i + U_i \mathbf{y}}).
\end{equation}
The parameters
of the model can be optimized using gradient descent to minimize the
negative log likelihood of data $\{\mathbf{y}_t\}$ under this model
\begin{equation}
\frac{\partial}{\partial \theta} p(\mathbf{y}_t) = -\mathbb{E}_{\mathbf{h}
\,|\, \mathbf{y}_t} \left[ \frac{\partial}{\partial \theta}
E(\mathbf{y}_t,\mathbf{h}) \right] + \mathbb{E}_{\mathbf{y}, \mathbf{h}} \left[
\frac{\partial}{\partial \theta} E(\mathbf{y},\mathbf{h})
\right]. \label{eq:joint_grad}
\end{equation}
The first expectation in this expression is easy to compute, but the
second is intractable and must be approximated. One popular
approximation for it is contrastive divergence \cite{hinton02}, which
uses a small number of Gibbs sampling steps \emph{starting from the
observed example} to sample from $p(\mathbf{y},\mathbf{h})$.
\begin{figure*}[t]
\centering
\makebox[\textwidth]{%
\begin{tabular}{ccc}
\includegraphics[height=0.93in]{crbm} &
\includegraphics[height=0.93in]{crbm_other} &
\includegraphics[height=0.93in]{drbm} \\
(a) & (b) & (c) \\
\end{tabular}}
\caption{Schematic diagrams of the various restricted Boltzmann
machines under investigation. (a) RBM for tag smoothing conditioned
on just auxiliary information: user, track, clips identity, (b) RBM
for tag smoothing conditioned on auxiliary information and tags of
other users, (c) discriminative RBM for audio classification.
Filled circles show variables that are always observed, open circles
show variables that are inferred at test
time.} \label{fig:crbms}
\end{figure*}
RBMs can be conditioned on other variables~\cite{Taylor+2007}. In
general, as shown in Figure~\ref{fig:crbms}(b), both the hidden
and visible units can be conditioned on other variables,
$\mathbf{u}=(v_1,\ldots,v_d)^T$ and $\mathbf{a}=(a_1,\ldots,a_A)^T$, respectively.
Including these interactions, the energy function becomes
\begin{equation}
E(\mathbf{y},\mathbf{h},\mathbf{u},\mathbf{a}) = -\mathbf{h}^T U \mathbf{y} - \mathbf{h}^T W \mathbf{u} - \mathbf{y}^T V \mathbf{a} - d^T \mathbf{y} - c^T \mathbf{h} \label{eqn:energy}
\end{equation}
and $p(\mathbf{y},\mathbf{h} \,|\, \mathbf{u}, \mathbf{a}) \propto e^{-E(\mathbf{y},\mathbf{h},\mathbf{u},\mathbf{a})}$, where
$W$ and $V$ are real matrices. The
vectors $V \mathbf{a}$ and $W \mathbf{u}$ act like additional biases on $\mathbf{y}$ and
$\mathbf{h}$. By setting the appropriate $W$ or $V$ matrix or conditioning
vector $\mathbf{u}$ or $\mathbf{a}$ to 0, the conditioning
can apply to only the visible units, as in
Figure~\ref{fig:crbms}(a), or only the hidden units, as in
Figure~\ref{fig:crbms}(c). For an observed data point
$\mathbf{y}_t,\mathbf{u}_t,\mathbf{a}_t$, the gradient of the log likelihood
with respect to a parameter $\theta$ becomes
\begin{multline}
\frac{\partial}{\partial \theta}\log p(\mathbf{y}_t \,|\, \mathbf{u}_t, \mathbf{a}_t) =
-\mathbb{E}_{\mathbf{h} \,|\, \mathbf{y}_t, \mathbf{u}_t, \mathbf{a}_t} \left[
\frac{\partial}{\partial \theta} E(\mathbf{y}_t,\mathbf{h},\mathbf{u}_t,\mathbf{a}_t) \right] \\
+ \mathbb{E}_{\mathbf{y}, \mathbf{h} \,|\, \mathbf{u}_t, \mathbf{a}_t}
\left[\frac{\partial}{\partial \theta} E(\mathbf{y},\mathbf{h},\mathbf{u}_t,\mathbf{a}_t)
\right]. \label{eq:cond_grad}
\end{multline}
\cite{salakhutdinov07} describes a conditional RBM used for
collaborative filtering in which only the hidden variables are
conditioned on other variables.
\subsection{Conditional RBMs for tag smoothing} \label{sec:smoothing}
We first employ conditional RBMs to learn relationships
among tags and between tags and users, tracks, and audio clips of tracks.
This model is purely textual, meaning that it only operates on the
tags and not on the audio.
All of the datasets used in this paper were collected by open endedly
soliciting tags from users to describe audio clips. This means that the
tags that they contain are most
likely relevant, but the tags that are not present are not
necessarily irrelevant. Thus there is a need to distinguish tags
omitted but still relevant from those that do not apply, as well as
tags that were included erroneously from those that truly apply. As
shown in \cite{mandel10b},
the co-occurrences of tags can be used to predict both of these cases.
For example, if the tags \tg{rap} and \tg{hip-hop}
frequently co-occur and a clip has been tagged \tg{hip-hop} but not
\tg{rap}, it would be reasonable to increase the likelihood of
\tg{rap} being relevant to that clip, although perhaps not as much as
if it had actually been applied by a user. Similarly, it might be
reasonable to decrease the likelihood of \tg{hip-hop} being relevant
as it was not corroborated by an application of \tg{rap}.
We use the {\em doubly conditional RBM} shown in Figure~\ref{fig:crbms}(b)
for this sort of tag ``smoothing'' as we call it.
The binary visible units
represent the tags that one user has applied to a clip and the hidden
units capture second order relationships between these tags. The
visible units are conditioned on auxiliary variables $\mathbf{a}$ which
represent as one-hot vectors the user, track, and clip from which a
vector of tags is observed. The hidden units are conditioned on
auxiliary variables $\mathbf{u}$, which represent the tags that other users
have applied to the same clip.
The vectors $\mathbf{y}$ and $\mathbf{u}$ are the same size, but whereas $\mathbf{y}$ is a
binary vector representing which of the fixed vocabulary of tags the
target user applied to the target clip, $\mathbf{u}$ is a vector of the
average of these binary vectors for all of the other users who have
seen the target clip. Thus the values in $\mathbf{u}$ are still between 0
and 1, but are continuous-valued. At test time, $\mathbf{u}$ is set to the
average tag vector of all of the users and predicts the tags that a
new user would likely apply to the clip given the tags that other
users have already applied.
The weights $V$ and $W$ are penalized with an $\ell_1$ cost to
encourage them to only capture dependencies that depend on specific
settings of the auxiliary variables and push into $U$ the dependencies
that exist independently of the auxiliary variables. This means that
$V$ should ideally only capture tag information relevant to a
particular user, clip, or track, $W$ should capture information about
the relationships between other user's tags and the current user's
tags, and $U$ should capture information about the co-occurrences of
tags in general.
Compare this to the singly
conditional RBM shown in Figure~\ref{fig:crbms}(a) and
described in \cite{mandel10b}. This CRBM also includes the conditioning of
the visible units on
the user, clip, and track information, but does not include the
conditioning on other users' tags. While the doubly conditional RBM
can use its modeling power to learn to predict specific user's tags
from general tag patterns, this singly conditional model must predict
both the general tag patterns and specific user's tags, a harder
problem. We found that the doubly-conditional RBM's smoothing trains
better SVMs on a validation experiment, and so we did not include the
singly-conditional RBM in our experiments.
\section{Discriminative RBMs} \label{sec:drbm}
One useful variant of the conditional RBM is the discriminative RBM
\cite{Larochelle+Bengio-2008}, shown in Figure~\ref{fig:crbms}(c).
The discriminative RBM is a conditional RBM that is trained to
predict the probability of the class labels, $\mathbf{y}$, from the rest of
the inputs, $\mathbf{x}$. Based on the energy function of \eqref{eqn:energy},
it corresponds to setting $\mathbf{u} = \mathbf{x}$ and $\mathbf{a}=0$.
For a set of observed data points $\{(\mathbf{y}_t,\mathbf{x}_t)\}$, the
discriminative RBM optimizes the log
conditional, $\log p(\mathbf{y}_t \,|\, \mathbf{x}_t)$, i.e.\ focusing on predicting
$\mathbf{y}_t$ from $\mathbf{x}_t$ well. A generative variant of this RBM would
instead optimize $\log p(\mathbf{y}_t, \mathbf{x}_t)$, the joint distribution
(in this case, $\mathbf{x}_t$ acts
as an extension of $\mathbf{y}_t$, i.e.\ it is not conditioned on it).
Looking at the parameter gradient of \eqref{eq:cond_grad},
we see that the second expectation requires a sum
over all configurations of $\mathbf{y}$. When $\mathbf{y}$ can take only a
few values, as in ordinary classification
tasks~\cite{Larochelle+Bengio-2008}, this expectation can be computed
efficiently and exactly. However, here $\mathbf{y}$ is a set of $C$ binary
indicators (the presence of a tag) that are not mutually exclusive, so
that the expectation has $2^C$ terms and must be approximated because
it cannot be computed in closed form.
Note that given a value for $\mathbf{y}$, $p(\mathbf{h}|\mathbf{y},\mathbf{x}_t)$
factors and is computed exactly.
\subsection{Approximations to the expectation} \label{sec:approx}
In the case of the discriminative RBM involving $\mathbf{y}$, $\mathbf{h}$, $\mathbf{u} =
\mathbf{x}$, and $\mathbf{a} = 0$, we approximate the $\mathbb{E}_{\mathbf{y},\mathbf{h} \,|\,
\mathbf{x}_t}$ term in \eqref{eq:cond_grad} in three different ways: using
contrastive divergence, mean-field contrastive divergence, and loopy
belief propagation. We also compare a similar, but tractable
computation that maximizes the pseudo-likelihood. The difficulty in
computing this expectation stems directly from the difficulty in
computing $p(\mathbf{y},\mathbf{h} \,|\, \mathbf{x}_t)$, which in turn is caused by the
interdependence of the $\mathbf{y}$ and $\mathbf{h}$ variables.
Contrastive divergence (CD) \cite{hinton02} has proven to be a very
popular algorithm for estimating the log-likelihood gradient in RBMs,
and it can also be used in the case of conditional RBMs.
Typically, it is used to compute
$\mathbb{E}_{\mathbf{y},\mathbf{x},\mathbf{h}}[\cdot]$ as opposed to here, where we compute
$\mathbb{E}_{\mathbf{y},\mathbf{h} \,|\, \mathbf{x}_t}[\cdot]$. To compute the usual
CD-$k$ update, $k$ steps of block Gibbs sampling, {\em starting
from the observed example} $(\mathbf{x}_t,\mathbf{y}_t)$, are used to
approximate the expectation. The block Gibbs chain is obtained
by alternating sampling from $p(\mathbf{h} \,|\,
\mathbf{y}, \mathbf{x})$ and sampling from $p(\mathbf{y},\mathbf{x} \,|\, \mathbf{h})$. In the case of the
conditional CD, we sample from $p(\mathbf{h} \,|\, \mathbf{y}, \mathbf{x}_t)$ and then from
$p(\mathbf{y} \,|\, \mathbf{h})$ (since $\mathbf{h}$ isolates $\mathbf{y}$ from $\mathbf{x}_t$),
keeping $\mathbf{x}_t$ fixed throughout. CD can be noisy
because it uses a small number of samples (usually only one), and it can be biased
because it doesn't necessarily run the Markov chain to convergence (usually only
1 to 10 steps).
The mean-field contrastive divergence approach approximates the $\mathbf{y}$
and $\mathbf{h}$ variables using
their conditional expectations (given each other)
and iteratively updates each one based on the
estimate of the other until convergence (note that $\mathbf{x}_t$ is fixed).
\begin{align}
\mathbf{h}^k &= \mathbb{E}[\mathbf{h} \,|\, \mathbf{y}^{k-1}, \mathbf{x}_t] = \sigm(c +
U \mathbf{y}^{k-1} + W \mathbf{x}_t) \\
\mathbf{y}^{k} &= \mathbb{E}[\mathbf{y} \,|\, \mathbf{h}^{k}, \mathbf{x}_t] = \sigm(b +
W \mathbf{h}^{k}).
\end{align}
In this case, we plug the continuous-valued expectations into these
equations instead of the sampled binary values that should formally be
used. While this method is
straightforward, it cannot capture multimodal distributions in $\mathbf{y}$
and $\mathbf{h}$, which makes it sensitive to initialization. We set the
initial condition $\mathbf{y}^{-1}=\mathbf{y}_t$, i.e.\ we initialize $\mathbf{y}$ at the
training label from which we compute $\mathbf{h}^0$, etc., which is why this is
referred to as mean-field contrastive divergence. We also tried to
use standard mean-field where $\mathbf{y}^{-1}=0$, but found the results to be
much worse.
Loopy belief propagation~\cite{Murphy99loopybelief} (LBP) is another
algorithm for approximating intractable marginals in a graphical
model. It relies on a message passing procedure between the variables
of the graph. While not guaranteed to converge it frequently does in
practice, and gives estimates of the true marginals that are often
more accurate than the iterative mean-field
procedure~\cite{WeissY2001}. In this setting, we used LBP to
estimate the marginals $p(y_j=1|\mathbf{x}_t)$, $p(h_k=1|\mathbf{x}_t)$ and
$p(y_j=1,h_k=1|\mathbf{x}_t)$ for a given $\mathbf{x}_t$ under the discriminative RBM,
and used those marginals to compute the $\mathbb{E}_{\mathbf{y},\mathbf{h} \,|\,
\mathbf{x}_t}$ term in equation \eqref{eq:cond_grad}. The quantity
$p(y_j=1|\mathbf{x}_t)$ can also be estimated at test time to predict the
labels. One method that has been shown to be useful in aiding
convergence is message damped belief propagation \cite{pretti05}. In
this case the updates computed by belief propagation are mixed
with the previous updates for the same variables in order to smooth
them, the damping factor being a parameter of the algorithm.
Another method for tuning the parameters aims to optimize not the
likelihood of the data, but the pseudo-likelihood \cite{besag75}. The
pseudo-likelihood circumvents the intractability of
computing the partition function in \eqref{eq:partition} by
considering only configurations of the visible units that are within a
Hamming distance of 1 from the training observation.
\begin{align}
\log PL(\mathbf{y} \,|\, \mathbf{x}) &= \sum_j \log p(y_j \,|\,
\mathbf{y}_{\backslash j}, \mathbf{x}) \\
= \sum_j &\log p(\mathbf{y} \,|\, \mathbf{x}) - \log \left(
p(\mathbf{y} \,|\, \mathbf{x}) + p(\tilde{\mathbf{y}}_j \,|\, \mathbf{x}) \right) \nonumber
\end{align}
where $\mathbf{y}_{\backslash j}$ is the labels vector $\mathbf{y}$ without the $j$th
variable and $\tilde{\mathbf{y}}_j$ is the labels vector $\mathbf{y}$ with the $j$th bit
flipped (the subscript $t$ is removed here for clarity). The pseudo-likelihood can be optimized using gradient descent.
Because of lack of space, we give pseudocodes for all the
aforementioned algorithms
in the appendix. Additionally, the
python code used for training these models is available on our
website\footnote{\url{http://www.iro.umontreal.ca/~bengioy/code/drbm_tags}}.
Note that while all of these methods can be used for training, not all
of them can be used at test time to estimate $P(y_j=1|\mathbf{x}_t)$.
Specifically, the pseudo-likelihood
requires the knowledge of $\mathbf{y}_{\backslash i}$, which is unavailable at
test time. Similarly, CD must be initialized from the true values of
$\mathbf{y}_t$ and $\mathbf{x}_t$. It is possible to use a Gibbs sampling method
similar to CD starting from an arbitrary initialization of $\mathbf{y}_t$, but
this is costly because the Markov chain may need to be run for many
iterations before it mixes well. We found that mean-field CD could be
successfully initialized with $\mathbf{y}^{-1}=0$ at test time.
\section{Experiments} \label{sec:experiments}
We performed a number of experiments to compare
different hyper-parameter settings, to compare different classifiers, and to
compare different tag smoothing techniques.
These experiments were based on three different datasets: data from
Amazon.com's Mechanical Turk service\footnote{\url{http://mturk.com}},
data from the MajorMiner music labeling
game\footnote{\url{http://majorminer.org}}, and data from Last.fm's
users\footnote{\url{http://last.fm}}. We compare the
discriminative RBM to standard (generative) RBMs, multi-layer
perceptrons, logistic
regression, and support vector machines. All of these algorithms were
evaluated in terms of retrieval performance using the area under the
ROC curve.
\subsection{Datasets}
Three datasets were used in these experiments. All of these datasets
were in the form of (user, item, tag) triples, where the items were
either 10-second clips of tracks or whole tracks. These data were
condensed into (item, tag, count) triples by summing across users.
The first dataset was collected from Amazon.com's Mechanical Turk service and is
described in \cite{mandel10b}. Users were asked to describe 10-second
clips of songs in terms of 5 broad categories including genre,
emotion, instruments, and overall production. The music used in the
experiment consisted of 185 songs selected randomly from the music
blogs indexed by the Hype Machine\footnote{\url{http://hypem.com}}.
From each track, five 10-second clips were extracted from
proportionally equally spaced points, for a total of 925 clips. Each
clip was seen by a total of 3 users, generating approximately 15,500
(user, clip, tag) triples from 210 unique users. We used the most
popular 77 tags for this dataset.
The second dataset was collected from the MajorMiner music labeling
game and is described in \cite{mandel08b}. Players were asked to
describe 10-second clips of songs and were rewarded for agreeing with
other players and for being original. This dataset includes
approximately 80,000 (user, clip, tag) triples with 2600 unique clips,
650 unique users, and 1000 unique tags. We used the most popular
77 tags for this dataset.
The final dataset was collected from Last.fm's website and is
described in \cite{SchifanellaEtAl2010}. The entire dataset consists of
approximately 7 million (user, track, tag) triples from 84,000 unique
users, 1 million unique tracks, and 280,000 unique tags. While only
the textual information was collected from Last.fm, we were able to
match it to 47,000 tracks in our own music collection. While this may
seem like a small fraction of the total number of tracks, the tracks
that were found included 1.5 million of the (user, track, tag)
triples, implying that the tracks we were able to match were tagged
more often than average. Following similar reasoning, many of these
users, tracks, and tags occurred infrequently, with 1
million (user, track, tag) triples in which all three items occurred in
at least 25 triples. Because these tags were
applied at the track level and not at the clip level, we selected
one clip from the center of
each track and assumed that they
should all be described with the track tags. This is the simplest
solution to this problem, although using some form of
multiple-instance learning might find a better solution
\cite{mandel08a}. We used the most popular 100 tags for this dataset.
Converting (item, tag, count) triples to binary matrices for training
and evaluation purposes required some care. In the MajorMiner and
Last.fm data, the counts were high enough that we could require the
verification of an (item, tag) pair by at least two people, meaning
that the count had to be at least 2 to be considered as a positive
example. The Mechanical Turk dataset did not have high enough counts
to allow this, so we had to count every (item, tag) pair. In the
MajorMiner and Last.fm datasets, (item, tag) pairs with only a single
count were not used as negative examples because we assumed that they
had higher potential relevance than (item, tag) pairs that never
occurred, which served as stronger negative examples.
\paragraph{Features}
The timbral and rhythmic features of \cite{mandel08b} were used to
characterize the audio of 10-second song clips. The timbral features
were the mean and rasterized
full covariance of the clip's mel frequency cepstral coefficients.
They capture information about instrumentation and overall production
qualities. The rhythmic features are based on the modulation spectra
in four large frequency bands. In fact, they are closely related to
the autocorrelation in those frequency bands. They capture
information about the rhythm of the various parts of the drum kit (if
present), i.e.\ bass drum, tom tom, snare, hi-hat. They also
discriminate between music that has a strong rhythmic component, e.g.\
dance music, and music that does not, e.g. folk rock. Each dimension
of both sets of features was normalized across the database to have
zero-mean and unit-variance, and then each feature vector was normalized
to be unit norm to reduce the effect of outliers. The timbral
features were 189-dimensional and the rhythmic features were
200-dimensional, making the combined feature vector 389-dimensional.
\subsection{Classifiers}
We compared a number of classifiers including two variants of
restricted Boltzmann machines, and three other standard classifiers.
The RBMs we compared were the discriminative RBM, described in
Section~\ref{sec:drbm}
and a standard generative RBM. Both RBMs
use Gaussian input units~\cite{welling05} in order to deal with the
continuous-valued features for $\mathbf{x}$. The
other classifiers include a multi-layer perceptron, logistic
regression, and support vector machines. For all datasets we select
the hyper-parameters of the model using a 5-fold cross-validation. In order
to increase accuracy of our measure,
for each fold we computed the score as an average across 4 sub-folds. Each
run used a different fold (from the remaining 4 folds) as the validation set
and the other 3 as the training set.
The discriminative RBM uses the gradient updates shown in
\eqref{eq:cond_grad}, while the generative RBM uses a different update
in which the second expectation is $\mathbb{E}_{\mathbf{y},\mathbf{x},\mathbf{h}}$ instead
of $\mathbb{E}_{\mathbf{y},\mathbf{h} \,|\, \mathbf{x}_t}$. The generative RBM attempts to
maximize $\log p(\mathbf{y},\mathbf{x})$, while the discriminative RBM attempts to maximize
$\log p(\mathbf{y} \,|\, \mathbf{x})$. It is also possible to use a mixture of these
two objective functions and maximize $\alpha \log p(\mathbf{y},\mathbf{x}) + \log p(\mathbf{y}
\,|\, \mathbf{x})$, referred to as a hybrid
generative/discriminative RBM~\cite{Larochelle+Bengio-2008}. In our
experiments, however, the hybrid RBM did
not improve on the DRBM, so we will not discuss it further. For each
model and dataset pair we
optimized the hyper-parameters using the cross validation described
above, selecting the hyper-parameters with the best performance on the
validation set averaging across folds and tags. Different
hyper-parameters performed best in each case, which is to be expected given the
differences in the models and in the data. For example, one would expect the
generative RBM to require more hidden units than the discriminative RBM because
it models the joint probability. Also on a
large dataset, one would expect to be able to use more hidden units without
overfitting. The hyper-parameters that performed best on the validation set
can be seen in Table~\ref{tab:params}.
\begin{table}[t]
\caption{Parameter settings found to perform best on validation
sets and used in experiments for discriminative and generative RBMs,
multi-layer perceptrons, and logistic
regression. LR stands for learning rate.} \label{tab:params}
\begin{center}
\begin{tabular}{lcccc}
\toprule
& & \multicolumn{3}{c}{Number of hidden units} \\
Model & LR & MTurk & MajMin & Last.fm \\
\midrule
Disc.\ RBM & 0.01 & 50 & 100 & 200 \\
Gen.\ RBM & 0.01 & 200 & 300 & 300 \\
MLP & 0.001 & 250 & 250 & 250 \\
Log.\ reg.\ & 2.0 & --- & --- & --- \\
\bottomrule
\end{tabular}
\end{center}
\end{table}
The multi-layer perceptron (MLP) is quite similar in structure to the
discriminative RBM in that it has nodes representing the features and
the classes and hidden nodes that capture interactions between them. The
main difference is that in estimating $p(\mathbf{y} \,|\, \mathbf{x})$ there is no modeling of the
interactions between the elements of $\mathbf{y}$ given $\mathbf{x}$. In the
discriminative RBM, at test time the unknown $\mathbf{y}$ and $\mathbf{h}$ interact
with one another through one of the methods described in
Section~\ref{sec:approx} until they reach a mutually agreeable
equilibrium. In the case of the MLP, however, at test time $\mathbf{h}$ is
computed deterministically from $\mathbf{x}$ and $\mathbf{y}$ is computed
deterministically from $\mathbf{h}$. The stochastic hidden units in the
discriminative RBM at test time allow it to better capture
interactions between the variables in $\mathbf{y}$ (i.e. the tags).
An even simpler classifier than the MLP is logistic regression, which
has no hidden layer and predicts each class directly from the input
features. We similarly optimize this using gradient descent, where
the cost function is the cross-entropy between the target
labels and the predictions, like for the MLP.
The final classifier we compared is the support vector machine (SVM).
Specifically we used a linear kernel and a $\nu$-SVM
\cite{scholkopf00} to automatically select
the $C$ parameter. We trained a different SVM for each tag as
an independent two-way decision (e.g. \tg{rock} vs not \tg{rock}).
While the above methods based on stochastic
gradient descent can be trained on all examples, SVMs are more
sensitive to the relative number of positive and negative examples, so
we had to more carefully select the training examples to use for each
tag. To do this, we selected as positive examples those clips to which
users applied a given tag most frequently and as negative examples
those clips to which users applied a given tag least frequently
(generally 0 times). The actual training labels used, however, were
still the standard $\pm1$ targets. We ensured that there were the same number
of positive and negative examples, up to 200 of each.
\paragraph{Metrics}
The performance of all of these algorithms on all of these datasets is
evaluated in terms of retrieval performance using the area under the
ROC curve (AUC) \cite{CortesAndMohri2004}. This metric scores the
ability of an algorithm to rank relevant examples in a collection above
irrelevant examples. A random
ranking will achieve an AUC of approximately 0.5, while a perfect
ranking will achieve an AUC of 1.0.
In certain experiments CRBMs were used to smooth the training data,
but the testing data was always the unsmoothed, user-supplied tags.
We measure the AUC for each tag separately.
We use the average across tags and folds as a overall measure of performance
and consider the standard error across folds for Figure~\ref{fig:params}.
For a more detailed comparison we use a two-sided paired t-test across folds,
per tag, between two models. We count the number
of tags for which each model performs better than the other at a 95\% significance level.
\paragraph{Implementation details / Running time}
In order to find the parameters that worked best for the DRBM, we used
a grid search. To avoid a prohibitive number of combinations, we
settled on a learning rate and number of hidden units before exploring
gradient approximations, Loopy Belief Propagation damping factors, and
numbers of iterations for CD, MF-CD or LBP. We also performed a much
wider parameter search on the smaller datasets, MTurk and MajorMiner,
keeping the same parameters for Last.fm, but varying the number of
hidden units. We found that the DRBM is insensitive to the number of
iteration steps while the computational cost increases considerably.
Training time varies according to many details, but on average, to
train a DRBM on MajorMiner took around 48 CPU-hours.
\subsection{Experiments}
\begin{figure}[t]
\begin{center}
\includegraphics[height=1.9in]{auc_vs_damping}
\end{center}
\caption{ Average area under the ROC curve with
standard errors on the MajorMiner dataset for the discriminative RBM trained
using loopy belief propagation with different damping factors.}
\label{fig:params1}
\end{figure}
\begin{figure*}[t]
\begin{center}
\makebox[\textwidth]{%
\begin{tabular}{cc}
\includegraphics[height=1.9in]{mturk_tr_te_approx} &
\includegraphics[height=1.9in]{majmin_tr_te_approx} \\
\end{tabular}}
\end{center}
\caption{Results on Mechanical Turk and MajorMiner comparing the
performance of different approximations
for the discriminative RBM during training and testing: contrastive
divergence (CD),
pseudo-likelihood (PL), mean field contrastive divergence (MF) and loopy
belief propagation (LBP). The approximations used during training are represented
on the x-axis, while the approximations used during testing are represented through
the gray value of the bar.}
\label{fig:params}
\end{figure*}
\begin{figure*}[t]
\begin{center}
\makebox[\textwidth]{%
\begin{tabular}{cc}
\includegraphics[height=1.6in]{drbm_vs_mlp} &
\includegraphics[height=1.6in]{drbm_vs_log} \\
\includegraphics[height=1.6in]{drbm_vs_rbm} &
\includegraphics[height=1.6in]{drbm_vs_svm} \\
\end{tabular}}
\end{center}
\caption{Comparison of discriminative restricted Boltzmann machine
autotagging retrieval performance to multi-layered perceptron (MLP),
logistic regression (LOG), (generative) restricted Boltzmann machine
(RBM), and support vector machine (SVM). Each bar shows performance
on a different dataset and its height is the number of tags on which
a two-sided paired t-test showed one algorithm to be significantly
better than the other in terms of area under the ROC curve.
Tags that were not significantly different are not included in
this plot.} \label{fig:comparison_results}
\end{figure*}
\begin{figure*}[t]
\begin{center}
\begin{tabular}{ccc}
\includegraphics[height=1.6in]{drbm_with_without_smoothing} &
\includegraphics[height=1.6in]{rbm_with_without_smoothing} &
\includegraphics[height=1.6in]{mlp_with_without_smoothing} \\
\end{tabular}
\begin{tabular}{ccc}
\includegraphics[height=1.6in]{log_with_without_smoothing} &
\includegraphics[height=1.6in]{svm_with_without_smoothing} &
\\
\end{tabular}
\end{center}
\caption{Comparison of autotagging retrieval performance with and
without conditional restricted Boltzmann machine-based smoothing for
discriminative restricted Boltzmann machine (DRBM), multi-layer
perceptron (MLP), (generative) restricted Boltzmann machine (RBM),
logistic regression (LOG) and support vector machine (SVM). Bars
show number of tags on which one set of autotaggers was
significantly better, as in
Figure~\ref{fig:comparison_results}.} \label{fig:smoothing_results}
\end{figure*}
\begin{table}
\newcommand{\hspace{0.8ex}}{\hspace{0.8ex}}
\caption{Average AROC across tags as a percentage for each
algorithm on each dataset with a specified number of tags (Tgs) and
with and without tag smoothing (Sm).} \label{tab:smoothing_results}
\begin{center}
\begin{tabular}{l@{\hspace{.5ex}}c@{\hspace{.5ex}}c@{\hspace{0.8ex}}c@{\hspace{0.8ex}}c@{\hspace{0.8ex}}c@{\hspace{0.8ex}}c@{\hspace{0.8ex}}c}
\toprule
Dataset & Tgs & Sm & DRBM & MLP & RBM & LOG & SVM \\
\midrule
MTurk & 27 & $-$ & 68.8 & 65.4 & 65.4 & 65.7 & 62.3 \\
MTurk & 27 & $+$ & 68.4 & 65.6 & 64.6 & 66.7 & 66.0 \\
MTurk & 77 & $-$ & 65.9 & 65.8 & 62.9 & 63.4 & 59.2 \\
MTurk & 77 & $+$ & 65.9 & 66.1 & 62.4 & 64.6 & 64.0 \\
MajMin & 77 & $-$ & 76.1 & 75.3 & 70.0 & 70.7 & 64.5 \\
MajMin & 77 & $+$ & 74.8 & 74.8 & 68.2 & 73.3 & 71.5 \\
Last.fm & 70 & $-$ & 72.2 & 72.0 & 65.9 & 70.3 & 64.6 \\
Last.fm & 100 & $-$ & 72.4 & 72.4 & 66.1 & 70.2 & 64.5 \\
\bottomrule
\end{tabular}
\end{center}
\end{table}
\paragraph{Experiment 1} The first experiment measured the
effectiveness of different settings of the smoothing hyper-parameter in
loopy belief propagation, meant to aid the
convergence of the algorithm.
Figure~\ref{fig:params1} shows the mean area under the ROC curve
(AUC) on the MajorMiner dataset of discriminative RBMs trained
using loopy belief propagation (LBP) with different damping
factors. We use 10 training iterations. The plots
show that the damping factor does not change the accuracy of the
model appreciably. Very similar results were obtained on the MTurk
dataset (not shown), while for Last.fm
dataset we only use $\beta=0.9$ which performed best on MTurk.
\paragraph{Experiment 2} The second experiment compared discriminative
RBMs trained and tested with different combinations of approximations
to the intractable expectation in \eqref{eq:cond_grad}.
We use different approximations on train and test to fully explore the
space of possibilities.
The left plot in Figure~\ref{fig:params} shows the mean AUC of
these discriminative RBMs on the MTurk dataset, while the right plot shows the
same results for MajorMiner.
The four training approximations, in order of
performance on MTurk, were contrastive divergence (CD), pseudo-likelihood (PL),
loopy belief propagation (LBP) and mean-field contrastive divergence
(MF). On MajorMiner the same order was preserved, except that loopy belief
propagation outperformed pseudo-likelihood.
The testing approximations, in order of performance were
LBP and mean-field (MF). The training approximation had a
larger impact on the final result than the testing approximation. For
Last.fm we only used CD during training and LBP at test time.
We also found that the model is quite robust to the number of training or
testing iterations for CD, MF or LBP.
\paragraph{Experiment 3} The third experiment compares the different
classifiers on the three datasets with and without tag smoothing.
We have also added a slight variation of the MTurk and Last.fm
datasets restricted to a subset
of the most popular tags (27 for MTurk and 70 for Last.fm).
Using a two-sided paired t-test
per tag, we compare all models to a discriminative restricted Boltzmann
machine trained on unsmoothed data.
The same test is done against all of the comparison models:
multi-layer perceptron (MLP), logistic regression (LOG), generative RBM
(RBM), and support vector machines (SVM).
Figure~\ref{fig:comparison_results} shows the number of tags
on which the DRBM outperforms the other algorithm. The DRBM
outperforms all of the other algorithms on many
more tags than it is outperformed. The MLP is evenly matched to it on
the full Last.fm 100 dataset, but on the other four datasets, the DRBM
is significantly better on many more tags than it is worse. The SVM
and logistic regression were previously the best performing
algorithms on these datasets.
Figure~\ref{fig:smoothing_results} shows the same analysis comparing
each classifier trained on the raw, user-supplied tags to the same classifier
trained on the tags smoothed by the proposed tag smoothing conditional
RBM.
Different subsets of the
auxiliary inputs were compared and the smoothing that gave the best
performance on the validation folds was selected. Because of the size
of the Last.fm dataset, only the unsmoothed tags were tested.
A number of interesting trends are visible in
Figure~\ref{fig:smoothing_results}. First, the SVM and logistic
regression
models are helped by the tag smoothing. This makes sense because
they treat each tag as a separate classification task and
cannot by themselves take advantage of the relationships between tags.
The MLP was sometimes helped by tag smoothing, but generally was not.
The fact that the RBMs were not helped by
the tag smoothing suggests that they are able to
capture by themselves the relationships between tags and do not need
the assistance of the tag smoothing.
\section{Conclusion}
This paper has described two applications of conditional restricted
Boltzmann machines to the task of autotagging music. The
discriminative RBM was able to achieve a higher average area under the
ROC curve than the previously best known system for this problem, the
support vector machine, as well as the multi-layer perceptron and
logistic regression. In order to be applied to this problem, the
discriminative RBM was generalized to the multi-label setting and
an in-depth analysis of four different learning algorithms for it were
evaluated. The best results
were obtained for a DRBM using contrastive divergence training and loopy
belief propagation at test time. The performance
of the SVM was improved significantly, although not to the level of
the DRBM, by the purely textual tag smoothing conditional RBM. Both
of these results demonstrate the power of modeling the relationships between
tags in autotagging systems.
\footnotesize
\bibliographystyle{unsrt}
|
\section{Introduction} \label{sec:intro}
In the early universe and in collapsing stars, the density of
neutrinos is so large that they produce significant refractive
effects for each other. In a seminal paper~\cite{Pantaleone:1992eq},
Pantaleone showed that this is not just a correction to the usual
matter effect~\cite{Wolfenstein:1977ue, Mikheev:1986gs, Kuo:1989qe},
but causes qualitatively new phenomena. They originate from flavor
off-diagonal refraction caused by neutrino oscillations which
therefore feed back on themselves. One consequence is self-induced
flavor conversion even for a very small mixing angle, caused by an
instability of the interacting ensemble. When the density decreases
slowly, for example in a supernova as a function of radius, this
instability leads to flavor swaps between neutrino spectra in sharp
energy intervals, causing conspicuous ``spectral splits.'' The body
of literature on collective oscillations has immensely grown, so we
cite only some theoretical key papers~\cite{Sigl:1992fn,
Pantaleone:1998xi, Samuel:1993uw, Kostelecky:1993dm,
Kostelecky:1995dt, Samuel:1996ri, Pastor:2001iu, Wong:2002fa,
Friedland:2003dv, Friedland:2006ke, Duan:2005cp, Duan:2006an,
Duan:2007mv, Raffelt:2007yz, Sawyer:2008zs, Hannestad:2006nj,
Balantekin:2006tg, Raffelt:2007cb, Dasgupta:2009mg, Fogli:2007bk,
Fogli:2008pt, Raffelt:2010za}, whereas we refer to a recent
review~\cite{Duan:2010bg} for the torrent of activities in the area
of supernova neutrinos.
One key feature of collective oscillations is that, given suitable
initial conditions, the entire ensemble evolves in a highly
correlated way (self-maintained coherence). We make this concept
more precise, if only for the simplest case of a homogeneous and
isotropic gas with two mixed flavors and ignoring ordinary matter.
Moreover, we mostly consider fixed density, although we use
motivations related to adiabatic changes of density.
In the absence of neutrino-neutrino interactions, every mode evolves
independently of the others. The ensemble may be described by
two-spinors in flavor space $\psi_E(t)$, occupation-number matrices
$\varrho_{E}(t)$, or the corresponding polarization vectors ${\bf
P}_E(t)$---effectively there are three important numbers for every
mode: the amplitudes of the two flavor components and their relative
phase. We always use polarization vectors and label the modes with
their vacuum oscillation frequency $\omega=\Delta m^2/2E$, where
negative $\omega$ denote antineutrinos. Vacuum oscillations are
described by the polarization vector of a given mode precessing with
frequency $\omega$ around a direction given by a unit vector ${\bf
B}$, the mass direction in flavor space. If neutrinos begin in
weak-interaction eigenstates, all ${\bf P}_\omega$ are initially
tilted relative to ${\bf B}$ by twice the vacuum mixing angle
(Fig.~\ref{fig:precession}). Each ${\bf P}_\omega$ precesses with a
different frequency and so eventually the modes distribute
themselves uniformly on the precession cone centered on ${\bf B}$,
approaching complete kinematical decoherence.
\begin{figure}[b]
\includegraphics[width=0.40\columnwidth]{fig1.eps}
\caption{Precession of a polarization vector ${\bf P}_\omega$ with
frequency $\omega$
around the $z$--direction (the mass direction in flavor space). The
initial orientation is in the weak-interaction direction, tilted by
twice the mixing angle relative to the
$z$--direction.\label{fig:precession}}
\end{figure}
Neutrino-neutrino interactions provide an additional force in terms
of the global ${\bf P}=\int d\omega\,{\bf P}_\omega$, leading to the
equations of motion (EoMs)
\begin{equation}\label{eq:precession}
\dot{\bf P}_\omega=(\omega {\bf B}+\mu{\bf P})\times{\bf P}_\omega\,.
\end{equation}
Here $\mu\sim\sqrt2\,G_{\rm F}n_\nu$ is a typical interaction
energy, its exact value depending on the normalization of the
polarization vectors. In the limit $\mu\to\infty$, and if initially
all ${\bf P}_\omega$ are aligned in the flavor direction, they
``stick together'' and precess around ${\bf B}$ with the common
frequency $\langle\omega\rangle$. These ``synchronized
oscillations'' are the simplest case of self-maintained coherence
\cite{Kostelecky:1995dt, Pastor:2001iu, Wong:2002fa}.
Starting with such a configuration, $\mu$ can slowly decrease as in
the expanding universe or with distance from a supernova. The
ensemble continues to oscillate coherently, but the ${\bf P}_\omega$
slowly separate and fan out in a plane that precesses around ${\bf
B}$. Such a pure precession mode can exist for any strength of $\mu$
\cite{Duan:2007mv, Raffelt:2007cb}. If $\mu$ decreases adiabatically
to zero, the final configuration is that all ${\bf P}_\omega$ with
$\omega$ above some frequency $\omega_{\rm split}$ point in the
positive ${\bf B}$ direction, the others in the opposite direction,
forming a ``spectral split.''
More complicated forms of coherent motion occur in the context of
multiple split formation~\cite{Dasgupta:2009mg}. A generic example
is a gas of neutrinos and antineutrinos of a given flavor with a
thermal distribution. Figure~\ref{fig:spectrum0} shows a Fermi-Dirac
distribution as a function of the re-scaled frequency $\omega=T/E$.
In this way $\omega$ is a dimensionless variable of order unity. The
neutrino flux streaming from a supernova core contains an excess of
$\nu_e$ over $\bar\nu_e$ because the trapped electron-lepton number
needs to be carried away. We mimic this situation schematically by
assuming a small degeneracy parameter $\eta=0.2$, corresponding to
$n_{\bar\nu}\approx0.70\,n_{\nu}$. The region around $\omega=0$
corresponds to the high-$E$ tail and the high-$\omega$ tails
(small~$E$) fall off as~$\omega^{-4}$. The mixing angle is taken to
be very small and so the mass and flavor directions are almost
identical. Notice that ${\bf P}_\omega$ pointing up means one
flavor, pointing down the other. In the flavor-isospin convention
\cite{Duan:2006an, Dasgupta:2009mg}, the interpretation is reversed
for antineutrinos, explaining that the spectrum in
Fig.~\ref{fig:spectrum0} is negative for $\omega<0$.
\begin{figure}[ht]
\includegraphics[width=0.75\columnwidth]{fig2.eps}
\caption{Neutrino spectrum ($z$-component of ${\bf P}_\omega$)
in the variable $\omega=T/E$, assuming that initially only one
flavor is populated with a Fermi-Dirac distribution
(degeneracy parameter
$\eta=0.2$). The integral over positive $\omega$ is normalized to unity.
\label{fig:spectrum0}}
\end{figure}
Without $\nu$-$\nu$ interactions only small-amplitude vacuum
oscillations occur, whereas for non-zero $\mu$ the system is
unstable. The evolution is nicely illustrated by showing the
instantaneous ``swap factor'' $z_\omega(t)$, i.e.\ the
$\omega$-dependent factor by which the original spectrum must be
multiplied to obtain the instantaneous $z$--component of ${\bf
P}_\omega(t)$, i.e.\ $P_\omega^z(t)=z_\omega(t)P_\omega^z(0)$
\cite{Dasgupta:2009mg} . The entire spectrum oscillates coherently.
In the top panel of Fig.~\ref{fig:swapexample1} we show numerical
results ($\mu=10$) for the maximum and minimum $z_\omega$ during an
oscillation period as well as their average.
\begin{figure}
\includegraphics[width=0.79\columnwidth]{fig3.eps}
\vskip-3pt
\caption{Maximum, minimum and average swap factor for
the thermal spectrum of Fig.~\ref{fig:spectrum0} if the interaction
strength begins with $\mu=10$ and then decreases adiabatically to 0.
Snapshots for $\mu=10$, 3, 1, 0.3 and 0 (top to bottom).
The initial spectrum is overlaid using a vertically
compressed scale relative to Fig.~\ref{fig:spectrum0}.\label{fig:swapexample1}}
\end{figure}
In this example and in any unstable spectrum, $z_\omega(t)$ is
initially a Lorentzian of a certain width $\lambda$ that oscillates
like an inverted gravitational pendulum with natural frequency equal
to the same $\lambda$~\cite{Dasgupta:2009mg}. When $\mu$ decreases
adiabatically, $z_\omega$ changes, the oscillation amplitude
decreases, and at $\mu=0$ takes on square-well shape: a spectral
swap with two sharp splits at its edges has formed. Several
snapshots are shown in Fig.~\ref{fig:swapexample1}.
These examples show ``coherence'' in some intuitive sense, but what
precisely does this mean? We here define coherent motion as a form
of evolution where neighboring modes do not separate from each
other. Kinematical decoherence means that different ${\bf
P}_\omega(t)$ within a frequency interval $\delta\omega$ eventually
separate by a large amount, no matter how small $\delta\omega$. If
we coarse-grain the ensemble with a bin width $\delta\omega$, the
coarse-grained $\langle {\bf P}_\omega\rangle$ becomes shorter than
${\bf P}_\omega$ as time goes on~\cite{Raffelt:2010za}. On the other
hand, coherence means that the coarse-grained ensemble remains
identical with the original one, no matter how much time has passed.
Pure precession is a trivial example because the polarization
vectors remain static relative to each other, whereas in the example
of Fig.~\ref{fig:swapexample1}, distant ${\bf P}_\omega(t)$ move
wildly relative to each other, yet neighboring ones do not separate
as time goes~on.
Independent evolution of neighboring modes, no matter how close
their frequencies, implies that every ${\bf P}_\omega(t)$ is
linearly independent. For vacuum oscillations this is obvious
because each ${\bf P}_\omega(t)$ is a harmonic function of frequency
$\omega$. On the other hand, kinematical decoherence can not occur
if every ${\bf P}_\omega(t)$ is a linear combination of a small
number $N$ of independent functions. In this case the system
performs what we call \hbox{$N$-mode} coherent oscillations. The
infinitely dimensional function space required for kinematical
decoherence has reduced to a small number of dimensions.
What is more, the same function space is spanned by the solutions
${\bf J}_i(t)$ of another system ($i=1,\ldots,N$) with vacuum
frequencies $\omega_i$ and the same effective density. So we
consider the EoMs of a new system
\begin{equation}\label{eq:EoMcarrier}
\dot {\bf J}_i=
(\omega_i{\bf B}+\mu{\bf J})\times{\bf J}_i\,.
\end{equation}
The functions ${\bf J}_i(t)$ can be found such that
\begin{equation}\label{eq:selfconsistency}
{\bf P}(t)={\bf J}(t)\,,
\end{equation}
meaning that both systems are dynamically equivalent. The original
modes ${\bf P}_\omega(t)$ and the linearly independent ``carrier
modes'' ${\bf J}_i(t)$ span the same function space: each ${\bf
P}_\omega(t)$ is a certain linear combination of the ${\bf J}_i(t)$.
Notice that usually the carrier modes can not be interpreted as a
coarse-grained representation of the original ensemble, but rather
are more abstract.
The idea that coherence in collective neutrino oscillations
corresponds to linear dependence among different modes and that the
evolution is dynamically equivalent to a small ensemble of carrier
modes provides a simple picture for self-maintained coherence. Our
primary interest is to use this approach for the construction of
explicit classes of solutions, notably for bimodal coherence.
We begin our study in Sec.~\ref{sec:nmode} with the equations of
motion, the idea of $N$-mode coherence, and the concept of carrier
modes. Then we proceed to the explicit cases of single-mode,
two-mode and multi-mode coherence in
Secs.~\ref{sec:singlemode}--\ref{sec:multimode} and conclude in
Sec.~\ref{sec:conclusions}.
\section{N-Mode Coherence} \label{sec:nmode}
\subsection{Equations of motion}
The two-flavor evolution of a homogeneous and iso\-tro\-pic neutrino
gas can be described by flavor polarization vectors ${\bf
P}_\omega(t)$, where $\omega=\Delta m^2/2E$ is the vacuum
oscillation frequency. ${\bf P}_\omega$ pointing up means one
flavor, pointing down the other. In the flavor-isospin convention
\cite{Duan:2006an}, the interpretation is reversed for antineutrinos
($\omega<0$). The flavor evolution under the influence of vacuum
oscillations and neutrino-neutrino interactions is given by
Eq.~(\ref{eq:precession}). This precession equation has been stated
many times~\cite{Sigl:1992fn, Pantaleone:1998xi, Kostelecky:1995dt,
Pastor:2001iu, Wong:2002fa, Raffelt:2007yz, Raffelt:2007cb,
Duan:2010bg}, most recently in great detail in
Ref.~\cite{Raffelt:2010za}, and we use it here without further ado.
A few points still deserve explicit mention. The polarization
vectors derive from an ensemble average and thus are classical
quantities. Our EoMs rely on a mean-field assumption: a given
neutrino is supposed to feel only the average effect of all others,
ignoring fluctuations or quantum correlations. (The quantum to
classical transition was studied in Refs.~\cite{Friedland:2003dv,
Friedland:2006ke}.) The length of ${\bf P}_\omega(t)$ is not fixed
by a quantization condition, but chosen at convenience. The value of
$\mu$ depends on this choice.
The EoMs conserve energy and derive from the classical Hamiltonian
\begin{equation}\label{eq:classHam}
H=\int \omega{\bf B}\cdot{\bf P}_\omega\,d\omega
+\frac{\mu}{2}\,{\bf P}^2\,.
\end{equation}
Here, $H$ and all ${\bf P}_\omega$ are functions of the canonical
coordinates and momenta. The evolution of any function $F$ on phase
space is given by the Poisson bracket $\dot F=[F,H]$. It is crucial
that the ${\bf P}_\omega$ play the role of classical spins, or
rather isospins. They obey the angular-momentum Poisson bracket
$[P_{\omega,x},P_{\omega',y}] =\delta(\omega-\omega')\,P_{\omega,z}$
and cyclic permutations. Notice that $[\,{\cdot}\,,\,{\cdot}\,]$ is
not a commutator---classical variables commute in the sense
$P_{x}P_{y}=P_{y}P_{x}$. The overall ${\bf P}$ is the total angular
momentum (in flavor space). Collective oscillations in the
mean-field approximation are thus equivalent to a set of interacting
classical spins. However, the simplicity of our Hamiltonian is
deceiving because it encapsulates all the complications of
collective flavor oscillations.
\subsection{Rotating frames}
\label{sec:rotatingframes}
Another conserved quantity is ${\bf P}$ projected on ${\bf B}$,
equivalent to angular-momentum conservation in the symmetry
direction. Therefore, we always work in the mass basis and use
coordinates where the $z$--direction coincides with ${\bf B}$
(Fig.~\ref{fig:precession}). Usually all ${\bf P}_\omega$ begin in a
weak-interaction eigenstate and thus in a common direction that is
tilted relative to ${\bf B}$ by twice the mixing angle.
The conservation of $P_z={\bf B}\cdot{\bf P}$ implies that the
internal motion of the ensemble, up to overall precession, does not
depend on $P_z$. Depending on convenience, we can study our system
from the perspective of a frame co-rotating with some
frequency~\cite{Duan:2005cp}. We may choose $\omega_{\rm c}=\mu P_z$
and without loss of generality study the equivalent EoMs
\begin{equation}\label{eq:EoM-transverse}
\dot{\bf P}_\omega=(\omega {\bf B}+\mu{\bf P}_\perp)\times{\bf P}_\omega\,,
\end{equation}
where ${\bf P}_\perp$ is the part of ${\bf P}$ transverse to ${\bf
B}$. We will frequently encounter the degeneracy between a shift of
the neutrino $\omega$ spectrum by some amount $\omega_{\rm c}$ and
$\mu P_z$. On the Hamiltonian level, such shifts amount to
subtracting the conserved quantity $\omega_{\rm c}{\bf B}\cdot{\bf
P}$.
Shifting a spectrum such as Fig.~\ref{fig:spectrum0} by an amount
$\omega_{\rm c}$ modifies the interpretation because
negative-frequency modes (anti-neutrinos) become positive-frequency
modes or vice versa. A mode that used to be occupied by $\bar\nu_e$
is then interpreted as being occupied, for example, by $\nu_\mu$.
However, this modified interpretation does not affect the abstract
dynamics, and we can always shift back at the very end. The
possibility to shift anti-neutrino and neutrino modes seamlessly
into each other is the main advantage of the flavor isospin
convention.
\subsection{Linear dependence of different modes}
Collective neutrino oscillations consist of strong correlations
among all modes and we have advanced that this means that the ${\bf
P}_\omega(t)$ depend linearly on a small number $N$ of independent
functions. To be more precise, we use $N$ to denote the number of
nontrivial functions because in addition the constant ${\bf B}$ will
be seen to appear, so ${\bf P}_\omega(t)$ actually depends on $N+1$
functions.
It is illuminating to diagnose $N$ in numerical examples. The
ensemble is represented by a large discrete set ${\bf P}_i(t)$ with
$i=1,\ldots,n$ and $n\gg 1$. To find the number of linearly
independent functions on a time interval \hbox{$t_1<t<t_2$} we
calculate the Gram matrix \hbox{$G_{ij}=\int_{t_1}^{t_2}dt\,{\bf
P}_{i}(t)\cdot{\bf P}_{j}(t)$}. Its rank reveals the number of
linearly independent functions. In reality, one finds $N+1$ large
eigenvalues of $G_{ij}$ and the rest very much smaller. They can be
made ever smaller by increasing the degree of adiabaticity, i.e.\
slowing down $\mu(t)$ and decreasing the mixing angle, i.e.\ making
the initial polarization vectors more collinear with ${\bf B}$.
Within constraints of numerical accuracy, the chosen time interval
is irrelevant to find $N+1$ if the ${\bf P}_i(t)$ are analytic
functions, but of course influences the numerical realization of the
Gram matrix.
We have always found the expected~$N$. The formation of a single
spectral swap as in Fig.~\ref{fig:swapexample1} corresponds to
bimodal coherence at any stage, i.e.\ the Gram matrix was found to
have rank 3. The evolution from a double-crossed spectrum towards
two swaps~\cite{Dasgupta:2009mg} corresponds to four-mode coherence
(rank~5), and so forth.
\subsection{Carrier modes}
If every ${\bf P}_\omega(t)$ is a linear combination of $N$
independent functions and the constant ${\bf B}$ we may choose any
set ${\bf P}_{\omega_i}(t)$ with $i=1,\ldots,N$ as a basis if they
are not accidentally degenerate. We then have ${\bf P}(t)=\int
d\omega\,{\bf P}_\omega(t)=a_0{\bf B}+\sum_{i=1}^N\,a_i {\bf
P}_{\omega_i}(t)$ as a unique linear combination. We define ${\bf
J}_i(t)\equiv a_i {\bf P}_{\omega_i}(t)$, obeying $\dot{\bf
J}_i=(\omega_i{\bf B}+a_0\mu{\bf B}+\mu{\bf J})\times{\bf J}_i$
where ${\bf J}=\sum_{i=1}^N{\bf J}_i$. In a rotating frame we can
transform away $a_0\mu{\bf B}$, or rather, we may choose another set
${\bf P}_{\omega_i'}(t)$ where $a_0'=0$.
So we can always find a set of linearly independent functions ${\bf
J}_i(t)$ with $i=1,\ldots,N$ that obey Eq.~(\ref{eq:EoMcarrier}),
fulfill the matching condition Eq.~(\ref{eq:selfconsistency}), and
together with ${\bf B}$ span the same function space as the original
ensemble ${\bf P}_\omega(t)$. The internal dynamics of our system is
the same if we study the co-rotating EoMs of
Eq.~(\ref{eq:EoM-transverse}). In this case the matching condition
is ${\bf P}_\perp(t)={\bf J}_\perp(t)$ and the complication about
the $a_0$ coefficient disappears.
We can express each ${\bf P}_\omega(t)$ of the original ensemble as
a linear combination of the chosen ``carrier modes'' ${\bf J}_i(t)$
\begin{equation}\label{eq:linearcombi}
{\bf P}_\omega(t)=b_0{\bf B}+\sum_{i=1}^N b_i\,{\bf J}_i(t)\,.
\end{equation}
This function must obey the original EoM of
Eq.~(\ref{eq:precession}). Inserting Eq.~(\ref{eq:linearcombi}) on
the l.h.s.\ of Eq.~(\ref{eq:precession}) and using
Eq.~(\ref{eq:EoMcarrier}) yields $\sum_{i=1}^N b_i(\omega_i{\bf
B}+\mu{\bf J})\times {\bf J_i}$. On the r.h.s.\ of
Eq.~(\ref{eq:precession}) we use ${\bf P}={\bf J}$ and
Eq.~(\ref{eq:linearcombi}) such that $-b_0\mu{\bf B}\times\mu{\bf
J}+(\omega{\bf B}+\mu{\bf J})\times\sum_{i=1}^N b_i{\bf J}_i$. With
${\bf J}=\sum_{i=1}^N {\bf J}_i$ in the first term and after
collecting everything we find
\begin{equation}
0={\bf B}\times\sum_{i=1}^N\bigl[
b_i\,(\omega-\omega_i)-b_0\mu\bigr]{\bf J}_i(t)\,.
\end{equation}
The functions ${\bf J}_i(t)$ are linearly independent by
construction and not proportional to ${\bf B}$. Therefore, we find
the coefficient $b_i=b_0\mu/(\omega-\omega_i)$ and thus
\begin{equation}\label{eq:linearcombi2}
{\bf P}_\omega(t)=b_0\left[
{\bf B}+\sum_{i=1}^N \frac{\mu}{\omega-\omega_i}\,{\bf J}_i(t)\right]\,.
\end{equation}
The required linear combination is unique up to an overall
proportionality factor. (This linear combination is an example of a
more general transformation briefly discussed in
Appendix~\ref{sec:transformedvectors}.)
The expression, Eq.~(\ref{eq:linearcombi2}), is singular at the
carrier frequencies. However, it is only the direction that matters,
so we introduce the transformed carrier spectrum
\begin{equation}\label{eq:Jhilbert2}
\bar{\bf J}_\omega(t)=\left[{\bf B}+
\sum_{i=1}^N\frac{\mu}{\omega-\omega_i}\,{\bf J}_i(t)\right]
\prod_{i=1}^N\,(\omega-\omega_i)\,,
\end{equation}
where the overbar has nothing to do with antiparticles. Introducing
the spectrum $g_\omega$ with $|{\bf P}_\omega|=|g_\omega|$, the
original ensemble in terms of the carrier modes is
\begin{equation}
{\bf P}_\omega=g_\omega\,\frac{\bar{\bf J}_\omega}{\bar{J}_\omega}\,,
\end{equation}
where $\bar J_\omega=|\bar{\bf J}_\omega|$.
We usually consider initial conditions such that the ${\bf
P}_\omega$ are almost collinear with ${\bf B}$ and we ask for the
probability for a neutrino mode $\omega$ to stay in its original
flavor, perhaps after an adiabatic change of $\mu$. This information
is contained in the $z$--component of $\bar{\bf J}_\omega$ and we
call
\begin{equation}
z_\omega(t)=\frac{{\bf B}\cdot\bar{\bf J}_\omega(t)}{\bar{J}_\omega}
\end{equation}
the time-dependent ``swap factor.''
Our discussion is motivated by the spectral swaps and splits that
form when the effective density encoded in $\mu$ slowly decreases to
zero. Since we are dealing with a Hamiltonian system, we expect the
adiabatic change of the parameter $\mu$ to deform the solution, but
that it remains coherent if it was initially coherent. For $\mu\to0$
we have $\bar{\bf J}_\omega\propto{\bf B}$, implying that in the end
all polarization vectors are collinear with ${\bf B}$. Conceivably
one could define \hbox{$N$-mode} coherence by the very property of
developing $N$ spectral splits when $\mu$ decreases adiabatically to
zero.
Our main interest is to use these ideas to construct explicit
classes of $N$-mode coherent solutions, notably the most general
bimodal solution. For a set of functions ${\bf J}_i(t)$ fulfilling
Eq.~(\ref{eq:EoMcarrier}), the matching condition
Eq.~(\ref{eq:selfconsistency})~is
\begin{equation}\label{eq:selfconsistency1}
{\bf J}(t)=\int d\omega\,g_\omega\,
\frac{\bar{\bf J}_\omega}{\bar{J}_\omega}\,.
\end{equation}
For a given spectrum $g_\omega$ and interaction strength $\mu$ this
condition restricts possible sets of carrier modes to provide an
$N$-mode coherent solution.
\section{Single-Mode Coherence} \label{sec:singlemode}
A first trivial case is what we may call zero-mode coherence, where
all ${\bf P}_\omega$ are exactly aligned with ${\bf B}$ and
therefore stay that way. This configuration also appears as a
limiting case of other forms of oscillation.
The first non-trivial case is single-mode coherence, identical with
pure precession~\cite{Duan:2007mv, Raffelt:2007cb}. The single
carrier mode ${\bf J}(t)$ precesses around ${\bf B}$ with some
co-rotation frequency $\omega_{\rm c}$ and obeys $\dot{\bf
J}=\omega_{\rm c}{\bf B}\times{\bf J}$ where the nonlinear term
$\mu{\bf J}\times{\bf J}=0$ has dropped out. According to
Eq.~(\ref{eq:Jhilbert2}) the transformed spectrum is $\bar{\bf
J}_\omega=(\omega-\omega_{\rm c})\,{\bf B}+\mu{\bf J}$. The matching
condition ${\bf P}={\bf J}$ then implies
\begin{equation}
\bar{\bf J}_\omega=(\omega-\omega_{\rm c})\,{\bf B}+\mu{\bf P}
\end{equation}
and thus $\bar{\bf J}^2=(\omega-\omega_{\rm c}+\mu P_z)^2+(\mu
P_\perp)^2$, where ${\bf P}_\perp$ denotes the part of ${\bf P}$
transverse to~${\bf B}$. The matching condition
Eq.~(\ref{eq:selfconsistency1}) is, in agreement with
Ref.~\cite{Raffelt:2007cb},
\begin{eqnarray}\label{eq:precessionconsisteny}
P_z&=&\int d\omega\,g_\omega\,
\frac{\omega-\omega_{\rm c}+\mu P_z}
{\sqrt{(\omega-\omega_{\rm c}+\mu P_z)^2+(\mu P_\perp)^2}}\,,
\nonumber\\*
P_\perp&=&\int d\omega\,g_\omega\,
\frac{\mu P_\perp}
{\sqrt{(\omega-\omega_{\rm c}+\mu P_z)^2+(\mu P_\perp)^2}}\,.
\end{eqnarray}
We have used the fact that all ${\bf P}_\omega$ lie in the plane
spanned by ${\bf B}$ and ${\bf P}$.
We could have studied our system from the perspective of a rotating
frame as discussed in Sec.~\ref{sec:rotatingframes}. In the above
expressions we recognize that nothing changes if we absorb $\mu P_z$
in the definition of $\omega_{\rm c}$.
For a given spectrum $g_\omega$ we can use these equations in
different ways. If $P_z$ is fixed by an initial condition we may ask
for the value of $P_\perp$ and $\omega_{\rm c}$ corresponding to a
specified value of $\mu$. In particular, for $\mu\to 0$ we can find
the split frequency for a given $P_z$. For explicit solutions in
this sense and further discussions we refer to the
literature~\cite{Duan:2007mv, Raffelt:2007cb}.
We may also use these conditions to construct all possible pure
precession solutions for a given $g_\omega$. To this end we
introduce the parameters $\kappa=\mu P_\perp$ and
$\gamma=\omega_{\rm c}-\mu P_z$, leading to the general pure
precession solution in the form of a two-parameter family depending
on $\gamma$ and $\kappa$,
\begin{eqnarray}\label{eq:pureprecessionsolution}
P_{\omega,z}&=&g_\omega\,
\frac{\omega-\gamma}
{\sqrt{(\omega-\gamma)^2+\kappa^2}}\,,
\nonumber\\*
P_{\omega,\perp}&=&g_\omega\,\frac{\kappa}
{\sqrt{(\omega-\gamma)^2+\kappa^2}}\,.
\end{eqnarray}
In addition one needs to specify the precession phase and thus the
instantaneous orientation of the comoving plane. Integrating over
$d\omega$ yields $P_z$ and $P_\perp$ and the corresponding
$\mu=\kappa/P_\perp$ and $\omega_{\rm c}=\gamma+\mu P_z$.
In terms of the parameters $\kappa$ and $\gamma$, the swap factor
for pure precession is
\begin{equation}\label{eq:swap-precession}
z_\omega=\frac{\omega-\gamma}
{\sqrt{(\omega-\gamma)^2+\kappa^2}}\,.
\end{equation}
Some examples are shown in Fig.~\ref{fig:swap-precession}. The step
occurs at frequency $\gamma$ and the transition region has width
$\kappa$.
\begin{figure}
\includegraphics[width=0.75\columnwidth]{fig4.eps}
\caption{Swap factor for pure precession with
$\kappa=0$, 0.5 and~1 according to
Eq.~(\ref{eq:swap-precession}).\label{fig:swap-precession}}
\end{figure}
For $\kappa\to0$ the swap factor is identical with the sign function
$z_\omega={\rm sgn}(\omega-\gamma)$. In this case the first relation
in Eq.~(\ref{eq:precessionconsisteny}) has the form $P_z=\int
d\omega\,g_\omega\,{\rm sgn}(\omega-\gamma)$ and can be used to find
$\gamma$ for a given $P_z$. The second equation is $ 1=\int
d\omega\,g_{\omega+\gamma}\,\mu/\sqrt{\omega^2+\kappa^2}$ and can be
used to find $\kappa$ explicitly in the limit where $\mu$ is very
small. In this limit, the integral is peaked around $\omega=0$ and
we can approximately pull out $g_\gamma$. The remaining integral
diverges for $\omega\to\pm\infty$, so we cut it off at some
frequencies $\pm\alpha$, leading to $1\sim2\mu g_\gamma
\log(2\alpha/\kappa)$ and implying that for small $\mu$
\begin{equation}\label{eq:kappasteep}
\kappa\propto e^{-1/(2\mu g_{\gamma})}\,.
\end{equation}
In other words, $\kappa$ becomes exponentially small for small $\mu$
or small $g_\gamma$.
\section{Bimodal Coherence} \label{sec:bimodal}
\subsection{Matching conditions}
Turning to bimodal coherence, we consider two carrier modes ${\bf
J}_{1,2}(t)$ with frequencies $\omega_{1,2}$ fulfilling the
precession equation, Eq.~(\ref{eq:EoMcarrier}). According to
Eq.~(\ref{eq:Jhilbert2}) the transformed carrier spectrum is
\begin{eqnarray}
\bar{\bf J}_\omega&=&(\omega-\omega_1)(\omega-\omega_2)\,{\bf B}
\nonumber\\
&&{}
+\mu\,(\omega-\omega_2)\,{\bf J}_1+\mu\,(\omega-\omega_1)\,{\bf J}_2\,.
\end{eqnarray}
Two interacting polarization vectors are dynamically equivalent to a
gyroscopic pendulum (Appendix~\ref{sec:twovecs}). To use this
equivalence we write $\omega_1=\omega_{\rm c}-\beta$ and
$\omega_2=\omega_{\rm c}+\beta$, where $\omega_{\rm c}$ is a
suitable co-rotation frequency such that the two carrier modes have
equal but opposite frequencies $\pm\beta$. To simplify notation we
implement $\omega_{\rm c}$ by shifting the spectrum, i.e.\ instead
of $g_\omega$ we use $g_{\omega+\omega_{\rm c}}$ in the integral for
the matching conditions.
From Appendix~\ref{sec:twovecs} we borrow ${\bf Q}={\bf J}_1-{\bf
J}_2+(\beta/\mu)\,{\bf B}$ for the pendulum's radius vector and find
\begin{equation}
\bar{\bf J}_\omega=\omega^2\,{\bf B}
+\mu\,\omega\,{\bf J}+\mu\,\beta\,{\bf Q}\,,
\end{equation}
where ${\bf J}={\bf J}_1+{\bf J}_2$ is the total angular momentum.
To spell out the matching conditions we need a unit vector in the
$\bar{\bf J}_\omega$ direction and thus need
\begin{eqnarray}
\bar{\bf J}_\omega^2&=&\omega^4+(\mu\beta{\bf Q})^2
+2\mu\omega^2(\beta{\bf B}\cdot{\bf Q}+{\textstyle\frac{1}{2}}\,\mu{\bf J}^2)
\nonumber\\
&&{}+2\mu\omega^3{\bf B}\cdot{\bf J}+2\mu^2\beta\omega\,{\bf Q}\cdot{\bf J}\,.
\end{eqnarray}
From Eq.~(\ref{eq:kappadef}) we take the natural frequency
$\lambda^2=\beta\mu|{\bf Q}|$, so the second term is $\lambda^4$.
The bracket in the third term is the energy $E$ of the pendulum. In
the first term on the second line we have ${\bf B}\cdot{\bf J}=J_z$,
a conserved quantity, and in the final term ${\bf Q}\cdot{\bf
J}=|{\bf Q}|\,S$ with $S$ the spin (conserved angular momentum
projection on the pendulum axis). Ordering by powers of $\omega$ we
find
\begin{equation}
\bar{\bf J}_\omega^2=\omega^4+2\mu\,(\omega^3J_z+\omega^2E +\omega\lambda^2 S)
+\lambda^4\,.
\end{equation}
As expected, this quantity is conserved, i.e.\ it does not depend on
the oscillation phase.
We also need the individual components of $\bar{\bf J}_\omega$. We
select directions in the plane transverse to ${\bf B}$ that move
with the system and we call the direction along ${\bf J}_\perp$ the
comoving $y$--direction, where ${\bf J}_\perp$ is the part of ${\bf
J}$ transverse to ${\bf B}$. The comoving $x$--direction is then
orthogonal and hence collinear with ${\bf B}\times{\bf J}$. We
project out the three components by taking the scalar product of
$\bar{\bf J}_\omega$ with ${\bf B}\times{\bf J}$, ${\bf
J}_\perp={\bf J}-J_z{\bf B}$ and ${\bf B}$ and find
\begin{eqnarray}\label{eq:Jxyz}
\bar{J}_{\omega,x}&=&\frac{\mu\beta\,{\bf Q}\cdot({\bf
B}\times{\bf J})}{J_\perp}\,,
\nonumber\\
\bar{J}_{\omega,y}&=&\frac{\mu\omega\,{\bf J}_\perp^2+
\mu\beta\,{\bf Q}\cdot{\bf J}_\perp}{J_\perp}\,,
\nonumber\\
\bar{J}_{\omega,z}&=&\omega^2+\mu\omega J_z+\mu\beta\,{\bf
Q}\cdot{\bf B}\,.
\end{eqnarray}
On the l.h.s.\ of the matching condition
Eq.~(\ref{eq:selfconsistency1}) we have in these comoving
coordinates ${\bf J}=(0,J_\perp,J_z)$ so that
\begin{equation}\label{eq:bimod}
\begin{pmatrix}0\\ 1\\ 0\end{pmatrix}=
\int d\omega\,
\frac{g_{\omega+\omega_{\rm c}}}{\bar{J}_\omega}\,
\begin{pmatrix}1\\ \mu\omega\\ \omega^2\end{pmatrix}\,.
\end{equation}
In the first line we have divided out $\bar J_{\omega,x}$ because it
does not depend on $\omega$, and using this condition then
simplifies the second line, and both are used to find the third. So
we could find the matching conditions without even spelling out the
scalar products in Eq.~(\ref{eq:Jxyz}).
Every solution is given in terms of the four pendulum parameters
$E$, $S$, $J_z$ and $\lambda$, although we have not spelled out
Eq.~(\ref{eq:Jxyz}) in terms of these conserved quantities. In
addition we have the shift frequency $\omega_{\rm c}$ as a fifth
parameter. However, we will see that $\omega_{\rm c}$ and $\mu J_z$
appear only combined, similar to the pure precession case, so
actually there are only four independent parameters. The dynamics is
given in terms of $\cos\theta$, the angle between ${\bf B}$ and
${\bf Q}$, that follows the pendulum differential equation of
Eq.~(\ref{eq:thetadot3}). In addition we can calculate the
precession motion following the usual pendulum treatment
(Appendix~\ref{sec:gyropendulum}).
Given the pendulum parameters and the EoM for $\cos\theta$ we have
all the information needed to describe the motion. However, these
parameters do not allow us to determine the original carrier modes
in a unique way---there are many solutions. Notice, in particular,
that the frequency $\beta$ and the length of ${\bf Q}$ are
degenerate parameters. Two interacting polarization vectors are
equivalent to a pendulum in flavor space, but the reverse mapping is
not unique. In particular, the carrier frequencies are not unique.
In certain limits, however, a particular choice may be most
convenient.
\subsection{Swap parameters}
While the pendulum parameters are well matched to the underlying
dynamics, they are not very intuitive when studying cases such as
the one shown in Fig.~\ref{fig:swapexample1}. To find a more
suitable parametrization we first spell out the swap factor
$z_\omega(t)=\bar{J}_{\omega,z}/\bar J_\omega$ explicitly,
\begin{equation}\label{eq:pendulumswap1}
z_\omega(t)=\frac{\omega^2+\mu\omega J_z+\lambda^2\cos\theta(t)}
{\sqrt{\omega^4+2\mu\,(\omega^3J_z+\omega^2E +\omega\lambda^2 S)
+\lambda^4}}\,.
\end{equation}
Inspecting the numerical results of Fig.~\ref{fig:swapexample1}, the
analytic swap factor for pure precession of
Eq.~(\ref{eq:swap-precession}) motivates the following ansatz for
the average swap factor
\begin{equation}\label{eq:doubleswap1}
\bar z_\omega=
\frac{(\omega-\gamma_1)}{\sqrt{(\omega-\gamma_1)^2+\kappa_1^2}}
\frac{(\omega-\gamma_2)}{\sqrt{(\omega-\gamma_2)^2+\kappa_2^2}}\,,
\end{equation}
where $\gamma_{1,2}$ give the location of the two steps and
$\kappa_{1,2}$ their respective widths.
Comparing the coefficients of the $\omega$ polynomial in the
denominators of Eqs.~(\ref{eq:pendulumswap1})
and~(\ref{eq:doubleswap1}) provides
\begin{eqnarray}\label{eq:mapparameters}
\lambda^4&=&(\gamma_1^2+\kappa_1^2)(\gamma_2^2+\kappa_2^2)\,,
\nonumber\\
\mu J_z&=&-(\gamma_1+\gamma_2)\,,
\nonumber\\
\mu E&=&\frac{(\gamma_1+\gamma_2)^2+2\gamma_1\gamma_2+\kappa_1^2+\kappa_2^2}{2}\,,
\nonumber\\
\mu S&=&-\frac{\gamma_1(\gamma_2^2+\kappa_2^2)+\gamma_2(\gamma_1^2+\kappa_1^2)}{\lambda^2}\,.
\end{eqnarray}
A further simplification is achieved if we express the time
variation of the swap factor in terms of a new dimensionless
variable
\begin{equation}
u(t)=\frac{\lambda^2\,\cos\theta(t)-\gamma_1\gamma_2}{\kappa_1\kappa_2}\,,
\end{equation}
implying
\begin{equation}\label{eq:bimodz}
z_\omega(t)=
\frac{(\omega-\gamma_1)(\omega-\gamma_2)+\kappa_1\kappa_2\,u(t)}
{\sqrt{[(\omega-\gamma_1)^2+\kappa_1^2]
[(\omega-\gamma_2)^2+\kappa_2^2]}}\,.
\end{equation}
Inserting $u(t)$ in the EoM for $\cos\theta(t)$ of
Eq.~(\ref{eq:thetadot3}) and using the parameter mapping of
Eq.~(\ref{eq:mapparameters}) provides
\begin{equation}\label{eq:ueom}
\dot u^2=(1-u^2)\left[(\gamma_1-\gamma_2)^2+(\kappa_1-\kappa_2)^2
+2\kappa_1\kappa_2(1-u)\right]\,,
\end{equation}
where the range of motion is $-1\leq u\leq+1$. Therefore, the
average is $\bar u=0$ and the average swap factor is indeed given by
the proposed expression Eq.~(\ref{eq:doubleswap1}).
The structure of Eq.~(\ref{eq:ueom}) suggests the interpretation
$u=\cos\vartheta$ with $\vartheta(t)$ some abstract angle variable.
In this way $1-u^2=\sin^2\vartheta$ and one finds
\begin{equation}
\dot\vartheta^2=(\gamma_1-\gamma_2)^2+(\kappa_1-\kappa_2)^2
+2\kappa_1\kappa_2(1-\cos\vartheta)\,.
\end{equation}
This angle always advances in the same direction with a modulated
velocity. The EoM is that of a gravitational pendulum with a speed
so large that it moves through the inverted position.
We show examples for the maximal, minimal and average swap factors,
corresponding to $u=+1$, 0 and $-1$ in Fig.~\ref{fig:swapexample2},
where $\gamma_{1}=-1$, $\gamma_{2}=+1$ and $\kappa_1=0.1$ are held
fixed and we show different cases for $\kappa_2=4$, 2, 1, 0.5 and
0.1 (top to bottom).
\begin{figure}
\includegraphics[width=0.8\columnwidth]{fig5.eps}
\caption{Maximum, minimum and average swap factor according to
Eq.~(\ref{eq:bimodz}) using $\gamma_{1}=-1$, $\gamma_{2}=+1$, $\kappa_1=0.1$,
and $\kappa_2=4$, 2, 1, 0.5 and 0.1 (top to bottom).
\label{fig:swapexample2}}
\end{figure}
We realize that in Eqs.~(\ref{eq:bimodz}) and (\ref{eq:ueom}) the
parameters $\gamma_1$ and $\gamma_2$ always appear either as
$\gamma_1-\gamma_2$ or as $\omega-\gamma_{1,2}$. So we can go to a
frame rotating with some frequency $\omega_{\rm c}$ by shifting
$\omega$ and $\gamma_{1,2}$ by this amount and the solution will be
the same. Since $-(\gamma_1+\gamma_2)/\mu$ has the interpretation of
$J_z$ this means that we can trade $J_z$ for a co-rotation frequency
$\omega_{\rm c}$. Once more this is the same effect discussed in
Sec.~\ref{sec:rotatingframes}. Instead of fixing $J_z=P_z$ it is
enough to match ${\bf P}_\perp={\bf J}_\perp$ and trade the mismatch
of $z$--components for a suitable co-rotation frequency $\omega_{\rm
c}$. Here, in particular, we can adjust $J_z$ such that $\omega_{\rm
c}=0$.
Spelling out the matching conditions in terms of the new parameters,
we find
\begin{equation}\label{eq:bimod2}
\begin{pmatrix}0\\ \mu^{-1}\\ P_z\end{pmatrix}=
\int d\omega\,
\frac{g_{\omega}}{\bar J_\omega}\,
\begin{pmatrix}1\\ \omega\\ (\omega-\gamma_1)(\omega-\gamma_2)\end{pmatrix}\,,
\end{equation}
where
\begin{equation}\label{eq:jbar}
\bar J_\omega=\sqrt{\left[(\omega-\gamma_1)^2+\kappa_1^2\right]
\left[(\omega-\gamma_2)^2+\kappa_2^2\right]}\,.
\end{equation}
For a given spectrum, the first equation establishes a constraint
among the four parameters $\kappa_{1,2}$ and $\gamma_{1,2}$. The
second allows us to calculate the required $\mu$ for this solution,
and the third provides the required $P_z$. We can also use the first
and second condition to simplify the third.
It is also of interest to spell out explicitly all three components
of $\bar{\bf J}_\omega$ of Eq.~(\ref{eq:Jxyz}). For $J_\perp$ we
find
\begin{equation}
(\mu {\bf J}_\perp)^2=
\kappa_1^2+\kappa_2^2-2\kappa_1\kappa_2u\equiv\kappa_u^2
\end{equation}
and
\begin{eqnarray}\label{eq:Jxyz2}
\bar{J}_{\omega,x}&=&\frac{\sqrt{(\gamma_2-\gamma_1)^2+\kappa_u^2}}{\kappa_u}\,
\kappa_1\kappa_2\sqrt{1-u^2}\,,
\nonumber\\
\bar{J}_{\omega,y}&=&\frac{(2\omega-\gamma_1-\gamma_2)\,\kappa_u^2+
(\gamma_2-\gamma_1)(\kappa_2^2-\kappa_1^2)}{2\kappa_u}\,,
\nonumber\\
\bar{J}_{\omega,z}&=&(\omega-\gamma_1)(\omega-\gamma_2)+\kappa_1\kappa_2u\,.
\end{eqnarray}
Divide these components by $\bar J_\omega$ of Eq.~(\ref{eq:jbar}) to
find the unit vector $\bar{\bf J}_\omega/\bar J_\omega$.
For $-\infty<\omega<+\infty$ this vector traces a closed curve on
the unit sphere, where the points $\omega=\pm\infty$ are at the
north pole. This closed curve moves as a function of $\vartheta(t)$.
Notice that for $u=\pm 1$, corresponding to the highest and lowest
swap-factor curves in Fig.~\ref{fig:swapexample1}, the comoving
$x$--component vanishes for all $\omega$, implying that all ${\bf
P}_\omega$ lie in the plane spanned by ${\bf B}$ and ${\bf P}$. The
case $u=-1$ corresponds to the lower swap factor, tracing out all
$z$--values between $-1$ and $+1$, i.e.\ the unit vectors trace out
the great circle spanned by ${\bf B}$ and~${\bf P}$.
\subsection{Pure precession limit: \boldmath$\kappa_1\kappa_2=0$}
Bimodal solutions are described by two carrier modes, which in turn
are equivalent to a gyroscopic pendulum. One possible form of motion
is pure precession: single-mode coherence appears as a limiting form
of two-mode coherence. In terms of our swap parameters, this case is
described by $\kappa_1=0$ or $\kappa_2=0$ where the solution is
static up to an overall precession. The swap factor, no longer
depending on time, is
\begin{equation}\label{eq:doubleswap3}
z_\omega={\rm sgn}(\omega-\gamma_1)\,
\frac{(\omega-\gamma_2)}
{\sqrt{(\omega-\gamma_2)^2+\kappa_2^2}}
\,,
\end{equation}
where we have used $\kappa_1=0$. In other words, we find the
pure-precession swap factor of Eq.~(\ref{eq:swap-precession}),
augmented with a step function centered on a frequency $\gamma_1$
(Fig.~\ref{fig:gyroprecessionswap}). Therefore, our previous result
is equal to the present one for the choice $\gamma_1\to-\infty$.
\begin{figure}
\includegraphics[width=0.75\columnwidth]{fig6.eps}
\caption{Swap factor for pure precession as derived
from the gyroscopic pendulum picture. The chosen parameters are $\gamma_1=-1$,
$\gamma_2=+1$, $\kappa_1=0$ and $\kappa_2=0.5$.\label{fig:gyroprecessionswap}}
\end{figure}
\subsection{Pure nutation limit: \boldmath$\kappa_1=\kappa_2$}
Another limiting case is when the swap factor is symmetric around
$\omega_{\rm c}=\frac{1}{2}(\gamma_1+\gamma_2)$, i.e.\ when
$\kappa_1=\kappa_2\equiv\kappa$. In a frame co-rotating with
$\omega_{\rm c}$ this solution is equivalent to a plane pendulum
where $J_z=S=0$ and was found previously~\cite{Dasgupta:2009mg}.
Shifting $\omega$ and $\gamma_{1,2}$ by $\omega_{\rm c}$ reproduces
the pure nutation solution
\begin{equation}\label{eq:solution}
{\bf P}_{\omega+\omega_{\rm c}}=
\frac{g_{\omega+\omega_{\rm c}}}
{\sqrt{\omega^4+2\omega^2\lambda^2c_{\rm m}+\lambda^4}}
\begin{pmatrix}\lambda^2 s\cr
\omega\lambda\sqrt{2(c_{\rm m}-c)}\cr
\omega^2+\lambda^2 c\cr\end{pmatrix}\,,
\end{equation}
where $s=\sin\theta$, $c=\cos\theta$, and $c_{\rm m}=
\cos\theta_{\rm min}$. In contrast with Ref.~\cite{Dasgupta:2009mg}
we here count the zenith angle $\theta$ from the north pole. In the
co-rotating frame we have $\gamma_{1,2}=\pm\gamma$ and therefore the
natural pendulum frequency is $\lambda^2=\kappa^2+\gamma^2$, whereas
the plane pendulum's highest point is given by $c_{\rm
m}=(\kappa^2-\gamma^2)/(\kappa^2+\gamma^2)$.
If the system has been set up with all polarization vectors
initially almost aligned with ${\bf B}$, the initial swap factor is
unity. The initial swap parameters are
$\gamma_1=\gamma_2=\omega_{\rm c}$ and $\kappa_1=\kappa_2=\kappa$
and are solutions of~\cite{Dasgupta:2009mg}
\begin{eqnarray}\label{eq:stability}
0&=&\int d\omega\,g_\omega\,
\frac{1}
{(\omega-\omega_{\rm c})^2+\kappa^2}\,,
\nonumber\\*
\frac{1}{\mu}&=&\int d\omega\,g_\omega\,
\frac{\omega-\omega_{\rm c}}
{(\omega-\omega_{\rm c})^2+\kappa^2}\,.
\end{eqnarray}
For the example of Fig.~\ref{fig:swapexample1} with the initial
value $\mu=10$ one finds $\omega_{\rm c}=-1.604$ and $\kappa=2.179$.
\subsection{Two pure precessions:
\boldmath$\kappa_{1,2}\ll|\gamma_2-\gamma_1|$}
When we begin with a bimodal system and the density decreases
adiabatically to zero, the end state shows a spectral swap with two
splits at its edges (Fig.~\ref{fig:swapexample1}). When $\mu$ has
become sufficiently small, the two quasi-step functions are well
separated and the overall swap factor looks like the product of two
pure precessions. This observation was the very motivation for
writing the general bimodal solution in terms of the swap parameters
$\gamma_{1,2}$ and $\kappa_{1,2}$ instead of the pendulum parameters
$\lambda$, $E$, $J_z$ and $S$.
In the present limit of $\kappa_{1,2}\ll|\gamma_2-\gamma_1|$, the
two split regions are essentially independent. The carrier modes
${\bf J}_{1,2}(t)$ precess essentially freely with frequencies
$\gamma_{1,2}$. To study this case it is easiest to represent the
$x$--$y$--components of all polarization vectors as complex numbers
of the form $J_x+\I J_y$, so the transverse components of the
carrier modes can be chosen as $\kappa_1 e^{\I\gamma_1 t}$ and
$\kappa_2 e^{\I\gamma_2 t}$ with $\kappa_{1}=\mu|{\bf J}_{1,\perp}|$
and $\kappa_{2}=\mu|{\bf J}_{2,\perp}|$. The abstract angle
characterizing the overall motion is here
$\vartheta=\pm(\gamma_1-\gamma_2)t$ because in the present limit
$\dot\vartheta^2=(\gamma_1-\gamma_2)^2$.
In the lower panels of Fig.~\ref{fig:swapexample1} we see that the
swap factor is asymmetric and steeper in the region where the step
falls into a spectral region where $g_\omega$ is smaller. This is
explained by Eq.~(\ref{eq:kappasteep}) where we have shown that for
a pure precession with very small $\mu$ the width of the step is
exponentially smaller for smaller $g_\omega$ in the step region.
As $\mu$ decreases adiabatically in a case like
Fig.~\ref{fig:swapexample1}, the system begins in a state of pure
nutation and ends in a state of two independent pure precessions.
These limiting forms and all intermediate cases are described by our
four-parameter analytic solution.
\subsection{Numerical examples}
To test if our analytic solution indeed corresponds to numerical
examples such as Fig.~\ref{fig:swapexample1}, we extract the
parameters $\gamma_{1,2}$ and $\kappa_{1,2}$ as $\mu$ slowly
decreases (Fig.~\ref{fig:adiabaticparameters}). Comparing the
analytic swap factor with the numerical one yields perfect
agreement.
\begin{figure}
\includegraphics[width=0.75\columnwidth]{fig7.eps}
\caption{Numerically determined swap parameters $\gamma_{1,2}$
and $\kappa_{1,2}$
for the example of Fig.~\ref{fig:swapexample1} as a function of
$\mu$ that adiabatically decreases from 10 to~0.\label{fig:adiabaticparameters}}
\end{figure}
\begin{figure}
\includegraphics[width=0.75\columnwidth]{fig8.eps}
\caption{Same as Fig.~\ref{fig:adiabaticparameters}, now with $\mu$ being
decreased from $\mu_0=10$ to zero as before, or increased from
$\mu_0=10$ to infinity.\label{fig:adiabaticparameters2}}
\end{figure}
Usually we begin with ${\bf P}_\omega^0$ almost aligned with ${\bf
B}$ and a chosen $\mu_0$. Then $\mu$ decreases from $\mu_0$ to 0,
going through different solutions as in Fig.~\ref{fig:swapexample1}
where $P_z$ remains conserved. If in this example we were to begin
with $\mu_0=3$ instead of 10, the $\mu=3$ solution would be the one
with a Lorentzian pattern. So for the same $g_\omega$, $P_z$ and
$\mu$ there exist different solutions that we can construct
numerically by adiabatic deformations.
We may also begin with a certain $\mu_0$ and then {\it increase} the
density as during supernova collapse. The initial pure nutation is
then deformed to another bimodal solution within our general
four-parameter family. For our usual example of
Fig.~\ref{fig:swapexample1} we show the swap parameters as a
function of $\mu$ in Fig.~\ref{fig:adiabaticparameters2}, assuming
the initial value is $\mu_0=10$. The solutions for $0<\mu<\mu_0$ are
from Fig.~\ref{fig:adiabaticparameters}. We identify the parameters
such that $\gamma_1<\gamma_2$.
\section{Multi-Mode Coherence} \label{sec:multimode}
\subsection{Four-mode coherence} \label{sec:fourmode}
More complicated forms of motion occur if the spectrum has two or
more independent instabilities, typically for multi-crossed
spectra~\cite{Dasgupta:2009mg}. Such spectra arise naturally if we
augment the Fermi-Dirac spectrum of a single species
(Fig.~\ref{fig:spectrum0}) with an initial population of the other
species, but with a larger temperature and smaller flux. This mimics
neutrinos streaming from a supernova core where a larger flux of
$\nu_e$ than $\bar\nu_e$ stream away and the other flavors $\nu_\mu$
and $\nu_\tau$ have smaller fluxes, larger average energies, and no
asymmetry. So if we add to the asymmetric spectrum of
Fig.~\ref{fig:spectrum0} a symmetric component of the other flavor
with a 1.25 times larger $T$ and 0.8 times smaller number density,
the difference spectrum relevant for flavor oscillations is shown in
Fig.~\ref{fig:spectrum1}.
\begin{figure}
\includegraphics[width=0.75\columnwidth]{fig9.eps}
\caption{Difference spectrum between two flavors. One is a
Fermi-Dirac distribution ($\eta=0.2$)
as in Fig.~\ref{fig:spectrum0}, the other
a nondegenerate spectrum with 1.25 larger $T$ and
0.8 times smaller $\nu$ density.
The integral over positive
$\omega$ of the degenerate flavor is normalized to unity.
Overlaid is the final swap factor after
$\mu$ has decreased adiabatically from 3 to~0.\label{fig:spectrum1}}
\end{figure}
This spectrum has two zero crossings with positive slope, allowing
for two independent instabilities, i.e.\ Eq.~(\ref{eq:stability})
can have two solutions~\cite{Dasgupta:2009mg}. If $\mu$ decreases
adiabatically from some value $\mu_0$ to 0, one finds two spectral
swaps and four concomitant splits. As an example we show the final
swap factor for $\mu_0=3$ in Fig.~\ref{fig:spectrum1}.
This case is an example for four-mode coherence. The complicated
motion of all ${\bf P}_\omega$ is equivalent to four carrier modes
as confirmed by numerical tests using the Gram matrix. When $\mu$
has become very small and the swaps are almost complete, we have the
now-familiar pattern of four independent pure precessions in the
four split regions. In the final small-$\mu$ limit, the swap factor
is
\begin{equation}
z_\omega=\prod_{i=1}^4\,
\frac{\omega-\gamma_i}{\sqrt{(\omega-\gamma_i)^2+\kappa_i^2}}\,,
\end{equation}
corresponding to four freely precessing carrier modes.
Equation~(\ref{eq:Jhilbert2}) implies that $\bar{\bf J}_\omega^2$ is
now a polynomial with leading term $\omega^8$, so the above
representation for $z_\omega$ is general, except that in the
numerator complicated time-dependent terms appear when we are not in
the small-$\mu$ limit. One can group the four carrier modes in pairs
and understand the overall motion as two gyroscopic pendulums
interacting by a dipole force. Apart from an overall precession, the
time dependence is then described by two nutation angles and one
relative precession angle.
If we choose some value for $\mu_0$, it is not assured that the
system indeed has two instabilities---it can have only one or none.
Therefore, as $\mu$ adiabatically decreases, the system can at first
show bimodal coherence and at some critical $\mu$-value the second
unstable mode kicks in, taking the system to a state of four-mode
coherence.
\subsection{Stability issues}
It is generic that a system of lower modality can develop higher
modality by an instability. Usually we begin with all polarization
vectors almost aligned with ${\bf B}$. If they were perfectly
aligned, we would have zero-mode coherence for any spectrum
$g_\omega$, but the smallest disturbance allows the unstable modes
to grow exponentially. We implement this disturbance in the form of
a small mixing angle, i.e.\ a small mismatch between the initial
condition and exact zero-mode coherence. The solutions ($\omega_{\rm
c},\kappa$) of Eq.~(\ref{eq:stability}) identify the unstable modes,
each one corresponding to a contribution of 2 to the total modality
$N$ relevant for the given $g_\omega$ and $\mu$.
We can also prepare a state of pure precession. With an arbitrary
spectrum $g_\omega$ and parameters $\gamma$ and $\kappa$, the
initial condition is defined by
Eq.~(\ref{eq:pureprecessionsolution}), allowing us to calculate the
required $\mu$. Henceforth the system evolves in single-mode
coherent motion unless it is unstable. Depending on $g_\omega$ and
with the slightest disturbance it can transit, for example, to
three-mode coherence. Likewise, we can set up the system in some
state of bimodal coherence with suitably chosen parameters
$\kappa_{1,2}$ and $\gamma_{1,2}$, yet it can transit to higher-mode
coherence.
A formal stability criterion is only available for zero-mode
coherence in the form of Eq.~(\ref{eq:stability}) that allows us to
decide, without solving the EoMs, if the system is stuck in its
alignment with ${\bf B}$ or not.
\section{Conclusions} \label{sec:conclusions}
We have studied two-flavor neutrino oscillations in a homogeneous
and isotropic neutrino gas under the influence of neutrino-neutrino
refraction. This system is equivalent to an ensemble of classical
spins, each labeled by its vacuum precession frequency $\omega$
around an external magnetic field ${\bf B}$ and a dipole-dipole
interaction with identical strength $\mu$ between any pair of spins.
We have argued that the conspicuous ``self-maintained coherence''
found in this system can be understood in terms of an equivalent
``carrier system'' of a few discrete modes with the same dynamics.
We have provided an explicit construction of how the original system
depends linearly on the carrier system, assuming certain matching
conditions. We have used this approach to construct the most general
bimodal solution.
Self-maintained coherence is to be contrasted with kinematical
decoherence. We have only studied ``purely coherent'' forms of
motion, but of course, depending on initial conditions, the system
can partly or fully decohere. We have not pursued the fascinating
question of order vs.\ disorder in our
system~\cite{Pantaleone:1998xi,Raffelt:2010za}.
Our results pertain to the simplest possible toy model and as such
do not have any immediate practical impact on issues of supernova
neutrino oscillations. The next step should be to apply these ideas
to more realistic situations, notably ``multi-angle cases,'' where
isotropy is no longer assumed. The main unresolved issues in the
theory of collective neutrino oscillations in the context of
supernova neutrinos remains the role of multi-angle effects. Every
neutrino mode depends on energy and on its direction of motion,
introducing much greater complications. It would be surprising if it
were not possible to develop a more analytical understanding of
collective flavor oscillations than has been achieved by
deconstructing numerical examples. We imagine that our results are
only a first step toward a more complete theory of collective flavor
oscillations. The ultimate goal is to provide a thorough theoretical
underpinning for what one is observing in numerical simulations, and
perhaps eventually in the neutrino signal of the next galactic
supernova.
\section*{Acknowledgements}
I thank B.~Dasgupta, A.~Dighe, A.~Patwardhan and A.~Smirnov for
discussions and critical questions and I.~Tamborra for comments on
the manuscript. This work was partly supported by the Deutsche
Forschungsgemeinschaft under grants TR-27 and EXC-153.
\section*{Note Added}
As this paper went to press, a preprint
appeared~\cite{Pehlivan:2011hp} that studies in detail the
properties of the quantum version of our Hamiltonian,
Eq.~(\ref{eq:classHam}), and in particular its invariants. The
constants of the motion discussed in our
Appendix~\ref{sec:transformedvectors} are identified as the
so-called Gaudin magnet Hamiltonians.
|
\section{Introduction}
The number of cancer cases in the United Kingdom has greatly increased in the last few decades. Approximately 200,000
new cases of cancer are discovered in England per year, causing 120,000 deaths \citep{nhs2000}. Although treatment
has improved recently, there is still much room for improvement \citep{nhs2004}. Several audits conducted in the UK
show that the waiting times for cancer treatment are not yet satisfactory
\citep{rcr1998,spurgeon2000,ash2004,summers2006,drinkwater2008}. Even though cancer care has improved recently within
the United Kingdom, radiotherapy capacity is still an important factor that has not received the adequate amount of
attention \citep{dodwell2006}. New targets set by the \cite{nhs2007}, in a program devised to enhance the effectiveness
of cancer treatment in the UK, make radiotherapy scheduling a very important problem.
This paper deals with a real-world radiotherapy treatment scheduling problem present in the Nottingham University
Hospitals NHS Trust, UK. The main aim is to design, implement and validate a scheduler for the linear accelerators
(hereafter referred to as linacs) used to deliver radiotherapy treatment. The purpose of the scheduler is to assist the
radiographer at the end of each day to create a schedule for the patients who are available to be scheduled.
There are some similarities between radiotherapy treatment scheduling and other appointment scheduling problems. Such
problems have the objective of optimising some quality of service measure, usually related to the waiting time of the
patient before being seen by a doctor, and deal with a stochastic arrival of patients. However, there are some key
differences as well. In other appointment scheduling problems, the objective is, usually, to schedule a single doctor
appointment for a patient, which often has a stochastic duration, and the patient often must be informed of the time of
the appointment immediately at the time of request. Those do not happen in radiotherapy treatment scheduling. The
objective here is to schedule a given number of appointments of deterministic duration with specific time intervals
between them for each patient. In addition, the scheduling of patients is usually done in batches once per day, in
order to find a better schedule.
As far as the authors are aware, few papers in the scientific literature deal with the problem of scheduling
radiotherapy treatments. \cite{petrovic2009} develop a genetic algorithm approach which considers both pre-radiotherapy
and radiotherapy treatment. \cite{kapamara2006} give a review of the radiotherapy patient scheduling (both
pre-treatment and treatment) problem, concluding that this problem is similar to a dynamic stochastic job-shop problem.
Mathematical and simulation models are commonly used to approach the problem or other similar problems.
\cite{conforti2008} define mathematical models for the scheduling of radiotherapy treatment. The objective in their
proposed model is to schedule as many patients as possible in a short period of time (e.g. one week). They consider a
block system, where a workday is split into a fixed number of time blocks/slots. In a successive paper, the same
authors extend the model to take patient availability into account, and run more extensive experiments
\citep{conforti2009} with real world data. \cite{conforti2010} then consider a non-block system, where the session time
may vary from one session to another. They observe that uniform appointment blocks do not represent real workload
properly, since the treatments can take either more or less time than the chosen block time. However, the models do not
consider all the constraints present in real-world radiotherapy scheduling, such as linac eligibility, treatments which
are not held on consecutive days, release dates different from the booking requests and patients who require multiple
sessions per day.
\cite{lev1974} develop a discrete event simulation model of patient flow in a diagnostic radiology department. This
model can be used to evaluate algorithms to be used for scheduling patients in such an environment. \cite{proctor2007}
propose a simulation model for a radiotherapy centre in the UK. The authors analyse two strategies to improve cancer
waiting times: 1) acquiring more equipment, such as a simulator and/or linac and 2) changing the working policy, such
as not requiring that radiographers treat the same patients for all sessions, extend working hours, etc.
The authors' previous research was concerned with constructive approaches to radiotherapy scheduling
\citep{petrovic2006,petrovic2008,petrovic2008a}. These methods vary some scheduling parameters, such as how
frequently to create schedules, and investigate the effect of changing the values of these parameters on the
performance of the algorithm. An algorithm based on the meta-heuristic GRASP was also developed to try to improve the
schedule generated by these constructive approaches.
The main contribution of this paper lies in the introduction of a new Integer Linear Programing (ILP) model for
scheduling radiotherapy treatments. Additional data from the hospital is gathered in order to better understand and
represent the radiotherapy scheduling problem. Experiments are conducted to evaluate the model in a myopic approach
varying the days on which schedules can be created, where no attempt is made to predict the patients who will arrive in
the near future.
This paper is organised as follows: \autoref{sec:problemdefinition} introduces radiotherapy treatment and describes
the procedure in place in the Nottingham University Hospitals NHS Trust, UK. The mathematical formulation of the
radiotherapy treatment scheduling problem is presented. In \autoref{sec:experiments}, experimental results are
presented. \autoref{sec:conclusion} gives conclusions about the experiments and future directions in this research.
\section{The Radiotherapy Treatment Scheduling Problem}
\label{sec:problemdefinition}
The radiotherapy treatment scheduling problem can be defined as the problem of scheduling a given number of
radiotherapy treatment sessions on linear accelerator machines. It is considered as a daily problem in which a number
of patients to be scheduled arrive at a radiotherapy centre. At the end of each day, the radiographer creates a
schedule for patients who arrived that day on a booking system which is partially booked with patients from previous
days.
Patients are grouped into different categories based on their waiting list status, which is classified as emergency,
urgent or routine. The status of a patient is decided by considering the site and the level of advancement of the
tumour. Patients are also grouped according to their treatment intent as radical (with the intent to cure) and
palliative (with the intent to alleviate symptoms and improve a patient's quality of life). This classification is also
used in the most recent audits by the Royal College of Radiologists \citep{summers2006,drinkwater2008}.
Each patient requires one or more types of radiation from the linacs, where the available radiation types are high
energy photon, low energy photon and electron. Since not all linacs can emit all types of radiation, this imposes a
linac eligibility constraint.
Patients can require more than one radiotherapy session, and the session duration can differ from one patient to
another, or even amongst the sessions of the same patient. Commonly, the first session of each patient is longer due to
validation and verification procedures.
Each linac can attend only one patient at a time. The capacity of each linac, given by the number of working hours of
the hospital staff, must not be exceeded on any given day.
Patients must undertake a set of preparatory steps before starting treatment, which are conveniently referred to as the
pre-treatment stage. Each patient can start treatment on or after the date when their pre-treatment finishes. This date
is referred to as the release date.
Patients may require 1, 2, 3 or 5 session days per week, where a session day is simply a day when the patient is
required to go to the radiotherapy centre to receive one or more fractions of the treatment. Patients who can be
treated on weekends can require up to 7 session days per week. Sessions must be scheduled with a strict number of days
between them, as such:
\begin{itemize}
\item patients who have 1 session day per week must have all sessions on the same day of the week for consecutive
weeks,
\item patients who have 2 session days per week must be scheduled either on Mondays and Thursdays consecutively or on
Tuesdays and Fridays consecutively,
\item patients with 3 session days per week must be scheduled on Mondays, Wednesdays and Fridays consecutively,
\item patients with 5 session days per week must have them on consecutive days excluding weekends,
\item and patients with 7 session days per week must have them on consecutive days including weekends.
\end{itemize}
Some patients must have a minimum number of sessions before the first weekend in order to prevent the tumour from
growing back after the first sessions. For example, some palliative patients must have at least 2 sessions before the
first weekend, and therefore, cannot start their treatment on a Friday. Some patients with 5 or less sessions are
required to have them all on the same week on consecutive days, without interruptions.
The majority of patients have only 1 session per day. The exceptions are CHART patients (Continuous Hyper-fractionated
Accelerated Radiotherapy Treatment), who require 3 fractions per day for 12 consecutive days with treatment starting on
a Monday. In addition, the presence of the doctor is required for the first session of some patients. Since each doctor
is available in the radiotherapy centre on only a few days of the week, this imposes one more eligibility constraint.
\autoref{fig:schedule} shows an example of a schedule for one linac where the opening times are set as from 9:00 to
10:00 for simplicity.
\begin{figure}[tb]
\centering
\includegraphics[width=.9\textwidth]{schedule.png}
\caption{An example of a typical schedule where the duration of the first session of each patient is slightly longer
than the others. Patient P1 has 10 sessions, 5 per week; patient P2 has 9 sessions, 5 per week; patient P3 is a
emergency patient with 1 session on Saturday; patient P4 has 1 session; patient P5 has 5 sessions, 3 per week; and
patient P6 has 4 sessions, 2 per week.}
\label{fig:schedule}
\end{figure}
Three target dates are set for each patient. The first is established by the \cite{nhs2004}. It states that each
patient must start their treatment no later than 62 days from the date upon which they are referred to an oncologist by
their general practitioner (GP) and no later than 31 days from the date when the decision to treat with radiotherapy
was made. This is referred to as the \emph{breach date}. The UK Cancer Network evaluates each radiotherapy centre
according to the number of patients that breach this target, thus minimising this number is the primary objective in
this research. The Department of Health targets are illustrated in \autoref{fig:timeline} \citep{nhs2005}.
\begin{figure}[tb]
\centering
\includegraphics[width=.9\textwidth]{timeline.png}
\caption{Time-line and Department of Health targets of patients diagnosed with cancer.}
\label{fig:timeline}
\end{figure}
The other two target dates have been established by the \cite{jcco1993}. They determine the good practice and the
maximum acceptable waiting times from the date the patient is first seen for suspected cancers to the first session of
treatment for each category of patients. \autoref{tab:jccotargets} shows the JCCO waiting time targets which have
been adjusted to the nomenclature used in this work. The JCCO targets are acknowledged by the majority of radiotherapy
centres in the UK \citep{ash2004,summers2006} and considered as a secondary objective in this research. It should
also be noted that the targets used for emergency patients are 24/48 hours for the good practice/maximum acceptable
waiting times, while the targets for urgent and routine patients depend on treatment intent. This better reflects the
nomenclature currently in use in hospitals and has been suggested by \cite{drinkwater2008}.
\begin{table*}[tb]
\centering
\begin{tabular}{l||c|c||c|c||c|c}
\hline
& \multicolumn{2}{c||}{emergency} & \multicolumn{2}{c||}{urgent} & \multicolumn{2}{c}{routine} \\
\cline{2-7}
& palliative & radical & palliative & radical & palliative & radical \\
\hline \hline
Good Practice & 24 hours & 24 hours & 48 hours & 2 weeks & 48 hours & 2 weeks \\
Maximum Acceptable & 48 hours & 48 hours & 2 weeks & 4 weeks & 2 weeks & 4 weeks \\
\hline
\end{tabular}
\caption{Waiting time targets established by the JCCO adjusted to the nomenclature used in this work.}
\label{tab:jccotargets}
\end{table*}
In addition to these three targets, the minimisation of waiting time from the decision to treat to the start of
treatment is considered. The hospital aims to minimise the waiting time while distributing it as evenly as possible
amongst patients. To measure it, the weighted sum of squared waiting times is calculated. This criterion can be
frequently seen in the literature for machine scheduling \citep{bagchi1987}, often used when large deviations of
completion time from the due date are undesirable.
To illustrate how the sum of squared waiting times can be applied to our problem and how it differs from other
frequently used criteria such as the sum of waiting times or the maximum waiting time, let us consider the following
example: on a given day, 3 patients arrive at the radiotherapy centre to be scheduled. Let us suppose that in one
possible schedule, patients 1, 2 and 3 have a waiting time of 1, 3 and 3, respectively, while in a second schedule, the
waiting times are 2, 2 and 3. The hospital would prefer the second schedule since it distributes the waiting time more
evenly among the patients. However, if either the sum of waiting times or the maximum waiting time are used, the value
of the objective function for each schedule will be the same for both schedules (7 and 3 respectively) making them
indistinguishable. If the sum of squared waiting times is used, the value of the objective function will be 19 and 17
for the first and second schedule respectively, enabling the algorithm to correctly choose the second schedule, which
would be preferred by the hospital.
A weight is assigned to each patient, which defines the relative importance of that patient respecting the JCCO
targets. As done in previous work \citep{petrovic2008}, the weights are set to 10 for emergency patients, 3 for urgent
and 1 for routine.
\section{Integer Linear Programming Model}
\label{sec:problemstatement}
The problem described can be formulated as an integer linear programming (ILP) model with the following input data:
\begin{itemize}
\item $M$: number of linacs,
\item $i$: index for linacs ($i = 1, \ldots, M$),
\item $N$: number of patients available to be scheduled,
\item $j$: index for patients available to be scheduled ($j = 1, \ldots, N$),
\item $\set{M}_j$: set of machines which can emit the required radiation types for patient $j$ ($\set{M}_j \subseteq
\{1, \ldots, M\},$ $\set{M}_j \ne \emptyset$),
\item $\set{W}_j$: set containing the days of the week when patient $j$ is allowed to have his/her first session
taking into consideration if the patient can receive treatment on weekends, the number of sessions a patient must
have before the first weekend, and the days the doctor is present at the hospital, if the doctor's presence is
required for the first treatment
($\set{W}_j \subseteq \{$Mon$, \ldots, $Sun$\}, \set{W}_j \ne \emptyset$),
\item $w_j$: relative importance (weight) assigned to patient $j$,
\item $b_j$: date when the booking request of patient $j$ is made,
\item $r_j$: release date of patient $j$,
\item $d^1_j$: breach date by which patient $j$ should start the treatment as established by the \cite{nhs2005},
\item $d^2_j$: maximum acceptable date by which patient $j$ should start the treatment as established by the
\cite{jcco1993},
\item $d^3_j$: good practice date by which patient $j$ should start the treatment as established by the
\cite{jcco1993},
\item $T$: number of days in the scheduling horizon,
\item $k$: index for days in the scheduling horizon ($k = 1, \ldots, T$),
\item $q_k$: day of the week of day $k$ ($q_k \in \{$Mon$, \ldots, $Sun$\}$),
\item $c_{ik}$ - available capacity of linac $i$ on day $k$ given in minutes,
\item $S_j$: number of sessions required for patient $j$,
\item $l$: index for sessions of patient $j$ ($l = 1, \ldots, S_j$),
\item $p_{jl}$: duration of session $l$ of patient $j$ given in minutes,
\item $u_{jkl}$: number of days patient $j$ must wait between sessions $l$ and $l + 1$ if session $l$ is scheduled on
day $k$, or 0 if sessions $l$ and $l + 1$ are on the same day.
\end{itemize}
The model is composed of only one set of decision variables, defined as follows:
\[
x_{ijkl} = \begin{cases}
1 & \mbox{if session } l \mbox{ of patient } j \mbox{ is scheduled on day } k \mbox{ on linac } i,\\
0 & \mbox{otherwise.}
\end{cases}
\]
The first constraints are presented to ensure that sessions are not scheduled on any invalid machine or day.
Constraint (\ref{eq:eligibility1}) imposes that sessions of patient $j$ are not scheduled on machines that do not emit
the types of radiation required for patient $j$. Constraints (\ref{eq:eligibility2})-(\ref{eq:eligibility4}) ensure
that patients are not scheduled on invalid days. Constraint (\ref{eq:eligibility2}) imposes that the day of any session
of patient $j$ cannot be schedule before the release date, constraint (\ref{eq:eligibility3}) guarantees that the first
session of the patient is not on an invalid day of the week for that patient, and constraint (\ref{eq:eligibility4})
ensures that no session other than the first of each patient can take place on the first day of the scheduling horizon.
\begin{eqnarray}
x_{ijkl} = 0 && i = 1, \ldots, M, i \notin \set{M}_j, j = 1, \ldots, N, k = 1, \ldots, T, l = 1, \ldots, S_j
\label{eq:eligibility1}\\
x_{ijkl} = 0 && i = 1, \ldots, M, j = 1, \ldots, N, k = 1, \ldots, r_j - 1, l = 1, \ldots, S_j
\label{eq:eligibility2}\\
x_{ijk1} = 0 && i = 1, \ldots, M, j = 1, \ldots, N, k = 1, \ldots, T, q_k \notin \set{W}_j
\label{eq:eligibility3}\\
x_{ij1l} = 0 && i = 1, \ldots, M, j = 1, \ldots, N, l = 2, \ldots, S_j
\label{eq:eligibility4}
\end{eqnarray}
Each pair of sessions of the same patient must be scheduled $u_{jkl}$ days apart, depending on the day $k$ when session
$l$ is scheduled. To ensure that session $l+1$ is scheduled $u_{jkl}$ days after session $l$, constraint
(\ref{eq:sequence}) is included.
\begin{multline}
x_{ijk'l'} = x_{ijkl} \qquad k' = k + u_{jkl}, l' = l + 1,\\
i = 1, \ldots, M, j = 1, \ldots, N, k = 1, \ldots, T - u_{jkl}, l = 1, \ldots, S_j - 1,
\label{eq:sequence}
\end{multline}
It is necessary to guarantee that all sessions are scheduled, and that each session is scheduled on exactly one day
and one linac. Constraint (\ref{eq:onelinac}) imposes this restriction.
\begin{equation}
\sum_{i=1}^M \sum_{k=1}^T x_{ijkl} = 1 \qquad j = 1, \ldots, N, l = 1, \ldots, S_j
\label{eq:onelinac}
\end{equation}
Finally, the available capacity on linacs must be respected. Constraint (\ref{eq:capacity}) ensures that the total
time used by sessions on day $k$ on linac $i$ does not exceed the linac capacity for that day.
\begin{equation}
\sum_{j=1}^N \sum_{l=1}^{S_j} p_{jl} \times x_{ijkl} \leq c_{ik} \qquad i = 1, \ldots, M, k = 1, \ldots, T
\label{eq:capacity}
\end{equation}
The criteria considered here are described below and are presented in order of their importance. This order has been
decided according to hospital staff preference.
\begin{itemize}
\item Minimisation of the number of patients who miss the breach date:
\begin{equation}
f_1(\mathbf{x}) = \sum_{i=1}^M \sum_{j=1}^N \sum_{k=d^1_j+1}^T x_{ijk1},
\label{eq:objbreach}
\end{equation}
\item Minimisation of the weighted number of patients who miss the JCCO maximum acceptable target:
\begin{equation}
f_2(\mathbf{x}) = \sum_{i=1}^M \sum_{j=1}^N \sum_{k=d^2_j+1}^T w_j \times x_{ijk1},
\label{eq:jccomax}
\end{equation}
\item Minimisation of the weighted number of patients who miss the JCCO good practice target:
\begin{equation}
f_3(\mathbf{x}) = \sum_{i=1}^M \sum_{j=1}^N \sum_{k=d^3_j+1}^T w_j \times x_{ijk1},
\label{eq:jccogood}
\end{equation}
\item Minimisation of the weighted average squared waiting times:
\begin{equation}
f_4(\mathbf{x}) = \sum_{i=1}^M \sum_{j=1}^N \sum_{k=b_j+1}^T (k - b_j)^2 \times w_j \times x_{ijk1}.
\label{eq:waitingtimesqrd}
\end{equation}
It should be noted that, even though the squared waiting time is being calculated, no decision variables are in fact
squared and the model remains linear.
\end{itemize}
In order to handle multiple objectives optimisation, a lexicographical ordering \citep{steuer1986,yu1989} is used. The
set of $Y$ objectives is indexed so that objective $m$ is more important than objective $m + 1$. A lexicographical
ordering preference is defined as follows: solution $\mathbf{x}^1$ is preferred to solution $\mathbf{x}^2$ iff $f_1(\mathbf{x}^1) < f_1(\mathbf{x}^2)$
or there is some $m \in \{2, \ldots, Y\}$ so that $f_m(\mathbf{x}^1) < f_m(\mathbf{x}^2)$ and $f_{m'}(\mathbf{x}^1) = f_{m'}(\mathbf{x}^2)$ for $m' =
1, \ldots, m - 1$.
\section{Experiments and Results}
\label{sec:experiments}
In order to evaluate the model, experiments are run simulating the everyday scheduling of a hospital. Each day, a
number of patients arrive in the radiotherapy department to be scheduled. At the end of the day, the radiographer
creates a schedule for the patients that arrived on that day.
Two sets of data were given to the authors by the Nottingham University Hospitals NHS Trust, UK. The first set contains
the waiting list status, intent, radiation and booking request of patients treated in a period of five years. The
second set contains all the attributes from each patient necessary to build a schedule of patients treated in a period
of one month. Similarly to previous work \citep{petrovic2008,petrovic2008a}, these two sets are combined to create 33
different data sets to be used in the experiments by the following procedure:
\begin{itemize}
\item Select a random period of time of 18 months of length from the first data set.
\item For each patient in that period, use the attributes of a random patient from the second data set with the same
waiting list status, intent and radiation to fill in the missing attributes in the first data set.
\end{itemize}
A warm-up period of 6 months is used, where the patients who arrive in the first 6 months are used only to fill the
booking system.
It should be noted that some combinations of waiting list status, treatment intent and required radiation type are more
frequent than others in the data received, and some are not present. The proportion of each combination of these
attributes can be seen in \autoref{tab:patienttype}, where patients marked as requiring electron radiation may also
require low energy photon and patients marked as requiring high energy photon may also require electron and low energy
photon.
\begin{table}[tb]
\centering
\begin{tabular}{l|c|c||c|c||c|c}
\hline
Required & \multicolumn{2}{c||}{Emergency} & \multicolumn{2}{c||}{Urgent} & \multicolumn{2}{c}{Routine} \\
\cline{2-7}
Radiation Type & Palliative & Radical & Palliative & Radical & Palliative & Radical\\
\hline
\hline
High energy photon & 1.3 & -- & 17.1 & -- & -- & 20.5 \\
Low energy photon & 2.4 & -- & 14.4 & -- & 2.8 & 15.1 \\
Electron & -- & -- & 10.2 & -- & 1.5 & 14.6 \\
\hline
\end{tabular}
\caption{Proportion (\%) of each combination of waiting list status, treatment intent and required radiation type}
\label{tab:patienttype}
\end{table}
A seasonality is identified in the arrival of patients according to the week of the year, as can be seen in
\autoref{fig:patientsperweek}.
\begin{figure}[tb]
\centering
\includegraphics[width=.8\textwidth]{patientsPerWeek.pdf}
\caption{Average number of patients of each waiting list status per week during the year}
\label{fig:patientsperweek}
\end{figure}
During the winter, the number of patients arriving each day is smaller than the year average. It slowly increases in
the next months, coming to a few peaks of patient arrivals in April and May. There is little variation in the next
months, ending with a steep drop in patient arrivals in the last two weeks of December. However, this variation is
slightly different for each waiting list status. The number of emergency patients has a smaller variation in the first
four months of the year, while the drop in the last two weeks of the year is not as steep for emergency and urgent as
it is for routine patients.
\autoref{fig:pretreatment} shows a histogram of the length of the time period between the decision to treat and the
release date for each waiting list status/intent combination from the second data set. During this time, the patient
goes through the pre-treatment phase.
\begin{figure}[tb]
\centering
\includegraphics[width=.8\textwidth]{preTreatmentDurationCom.pdf}
\caption{Histogram of the length of the time period between the decision to treat and release date given in days}
\label{fig:pretreatment}
\end{figure}
There seems to be a strong correlation between the length of this time period and the waiting list status/intent of
patients. Emergency patients have the shortest time period between the decision to treat and the release date of 1 day
in average. In contrast, urgent patients have a release date on average 11 days after the decision to treat has been
made, routine palliative patients an average of 18 days and routine radical patients an average of 33 days. The largest
values in \autoref{fig:pretreatment} can be explained by adjuvant patients, who first have a different treatment, such
as hormone therapy or chemotherapy, and then have adjuvant radiotherapy. For these cases, the breach date is calculated
from the date of the CT scan during pre-treatment instead of from the decision to treat.
It should also be noted that the differences presented in \autoref{fig:pretreatment} make it impossible for some
patients to meet all due dates. It is impossible for 17\% of emergency patients to meet their JCCO good practice of 1
day due to their release date being after this due date, but it is possible for all of them to meet the JCCO maximum
acceptable of 2 days. Around 94\% of non-emergency palliative patients cannot meet the JCCO good practice of 2 days,
and 23\% cannot meet the JCCO maximum acceptable of 14 days. For radical patients, the due dates are even harder to
meet, as 98\% of radical patients cannot meet the JCCO good practice of 14 days and 45\% cannot meet the maximum
acceptable of 28 days. In addition, 12\% of patients, all of which are routine and radical, cannot meet the breach date
of 31 days. This analysis can also give an approximation of the best possible values for the first three criteria.
Other important aspects of the data include the number of sessions of each patient and the number of session
days/sessions per day, shown in \autoref{fig:sessiondays}.
\begin{figure}[tb]
\centering
\subfigure[Histogram of the number of sessions for each patient type]{
\includegraphics[width=.47\textwidth]{patientFractions.pdf}
\label{fig:fractions}}
\subfigure[Number of patients with each number of session days per week/sessions per day of each waiting list status]{
\includegraphics[width=.47\textwidth]{sessionsPerWeek.pdf}
\label{fig:sessionsperday}}
\caption{Frequency of patient types in each data set}
\label{fig:sessiondays}
\end{figure}
There seems to be a strong correlation between the number of sessions and the waiting list status/intent, as can be
seen in \autoref{fig:fractions}. All emergency, and around 63\% of urgent patients, have only one fraction. Routine
patients usually have a very high number of sessions, with an average of 21 sessions for each patient. Also, around
64\% of the patients who have more than 1 session have a number of fractions multiple of 5, showing a preference for
treatments which take a round number of weeks.
\autoref{fig:sessionsperday} shows how many patients of each waiting list status have each possible number of session
days per week/sessions per day. There also seems to be a correlation between these attributes and the waiting list
status of each patient. It is possible to see that the majority of patients (around 68\%) have 5 sessions per week and
1 session per day, where the exceptions are:
\begin{itemize}
\item all emergency and most urgent patients who have 1 session day per week,
\item a few routine patients who have 2 or 3 sessions per week,
\item and a few routine patients classified as CHART, who have 3 sessions per day, 7 days per week for 12 consecutive
days.
\end{itemize}
The oncology ward in the Nottingham University Hospitals NHS Trust, UK, currently has four linacs in total:
\begin{itemize}
\item 1 that emits low energy photon radiation (type $A$),
\item 1 that emits low energy photon and electron radiation (type $B$),
\item 2 that emit all three types of radiation (type $C$).
\end{itemize}
Linacs are available from 8:45 to 18:00 on Monday to Friday for weekday sessions and from 9:00 to 13:00 on Saturdays
and Sundays for weekend sessions.
In order to mimic the current scheduling policies in the hospital, a simplification is made. Patients who require only
low energy photon must be scheduled on linacs of type $A$, and patients who require electron and low energy photon or
only electron radiation must be scheduled on linacs of type $B$. Patients who require high energy photon can only be
scheduled on linacs of type $C$, regardless of requiring additional radiation types. This implies a simplification of
the problem, which is discussed later in this paper.
Two further simplifications are made in this work: random machine down times and patients not showing up for treatments
are not considered. However, the model can still be used as presented in both situations by adjusting each available
capacity $c_{ik}$ and by implementing a recovery protocol to reschedule sessions missed by patients. This protocol
could be based on integer linear programming or on other techniques.
In this work, all experiments are run using ILOG CPLEX 12.1, an optimisation software package, on a PC with an AMD
Opteron 2.2GHz CPU and 2GB of RAM under the Scientific Linux operating system.
\subsection{Experiments with Schedule Creation Day (SCD)}
Although using the model on its own on a daily basis can cause the earliest appointments to be always used first,
changes can be made to the scheduling policy to counteract this effect. Possible changes have been presented in
\cite{petrovic2008,petrovic2008a} involving the preferred date for patients to start, days when schedules are created
and machine reservation. Two of those approaches are also analysed in this paper by using the proposed model with a
standard ILP solver in place of the constructive algorithm previously presented. The first approach introduces a
parameter called \emph{schedule creation day} (SCD) to specify days of the week for each patient when a schedule can be
created for patients of a given waiting list status. If a patient has a decision to treat made on a date when the
creation of a schedule is not allowed, the schedule is created on the first following allowed day.
The intention is to investigate whether the accumulation of patients to be scheduled will lead to better schedules.
Obviously, the search space becomes larger and it may lead to solutions of higher quality. The SCD values considered
are 5 (every weekday), 3 (on Mondays, Wednesdays and Fridays), 2 (on Tuesdays and Fridays) and 1 (only on Fridays) for
urgent and routine patients and fixed as 7 (every day) for emergency patients.
The model presented is dependent on the value of the scheduling horizon $T$, which must be supplied as input and must
be large enough to accommodate all patients. To calculate this value, the constructive algorithm presented in
\cite{petrovic2008a} is used. The proposed algorithm sorts the patients to be scheduled according to their waiting list
status, breach date, JCCO maximum acceptable target and number of sessions. Following that order, each patient is
scheduled on the earliest day possible. The value of $T$ is taken from the date of the last session in this solution
and increased by 14 days in order to augment the search space for CPLEX. Since $T$ is taken from a feasible solution,
it is impossible for the instance to have no feasible solutions. The solution found by the constructive approach is
also used as starting point by CPLEX in order to increase its execution speed.
The schedule of patients who must be scheduled on linacs of one type has no influence on the schedule of patients who
must be scheduled on linacs of other types. Therefore, it is safe to split the problem into three sub-problems in order
to speed up the process of finding a schedule each day. Each sub-problem is composed of linacs of one type and all
patients who must be scheduled on linacs of that type. Each sub-problem is solved individually and the schedules of
patients are combined to form a complete schedule. With this simplification, linac eligibility constraints are no
longer necessary and can be removed. However, they are kept in order to avoid loss of generality in the model. The time
limit of 10 minutes is equally divided amongst the three sub-problems.
Each configuration is run on the 33 instances described in \autoref{sec:experiments}. When running a number of
experiments with randomly generated data, it is possible that one configuration achieves a better average result than
the other configurations simply by chance. In order to determine whether or not the means of the criteria values of two
sets of experiments are significantly different, the Mann-Whitney U (MWW) test is used on experiment values re-sampled
from a bootstrap approximation \citep{leger1992}. The MWW test \citep{mann1947} is able to determine if there is
significant statistical evidence that one configuration achieves better results than another configuration.
Bootstrapping is a computationally intensive technique based on data re-sampling. By using bootstrapping, it is
possible to perform valid statistical tests without making unrealistic or unverifiable assumptions about the criteria
values, like their distribution or variance \citep{efron1979,yuan2007}.
The MWW tests are run for each pair of configurations with an overall confidence of 90\% and the bootstrap is set to
1000 replications. Results for each criterion are shown in \autoref{tab:scd}, where ``Breach'' is the percentage of
patients who miss their breach date, ``JMax'' and ``JGood'' are the weighted percentage of patients who miss their JCCO
maximum acceptable and good practice targets, respectively, and ``Waiting'' is the average weighted squared waiting
time per patient.
The values in bold are the ones where there was no significant statistical evidence of any of the values found by the
other configurations being better for that criterion. This form of evaluation is also used in the subsequent sections.
\begin{table*}[tb]
\centering
\begin{tabular}{cc|cccc}
\hline
\multicolumn{2}{c|}{SCD} & Breach (\%) & JMax (\%) & JGood (\%) & Waiting \\
\cline{1-2}
Urgent & Routine & & & & \\
\hline \hline
5 & 5 & 34.98 & 47.67 & 86.59 & 1,670 \\
5 & 3 & 34.79 & 47.49 & 86.57 & 1,649 \\
5 & 2 & 34.89 & 47.56 & \textbf{86.51} & 1,670 \\
5 & 1 & 34.13 & 47.22 & \textbf{86.47} & \textbf{1,623} \\
3 & 5 & 35.08 & 47.58 & 86.59 & 1,682 \\
3 & 3 & 34.74 & 47.39 & \textbf{86.57} & 1,650 \\
3 & 2 & 35.01 & 47.48 & \textbf{86.53} & 1,678 \\
3 & 1 & 34.04 & \textbf{47.13} & \textbf{86.47} & \textbf{1,611} \\
2 & 5 & 34.92 & 47.53 & \textbf{86.54} & 1,665 \\
2 & 3 & 34.85 & 47.42 & \textbf{86.52} & 1,663 \\
2 & 2 & 34.68 & 47.40 & \textbf{86.49} & 1,657 \\
2 & 1 & \textbf{33.95} & \textbf{47.06} & \textbf{86.41} & \textbf{1,624} \\
1 & 5 & 35.37 & 47.34 & 89.05 & 1,680 \\
1 & 3 & 35.22 & 47.22 & 89.03 & 1,678 \\
1 & 2 & 34.92 & \textbf{47.24} & 89.03 & 1,668 \\
1 & 1 & \textbf{33.77} & \textbf{46.89} & 88.99 & \textbf{1,624} \\
\hline
\end{tabular}
\caption{Results obtained varying the frequency of creating schedules for patients of each waiting list status.}
\label{tab:scd}
\end{table*}
In general, good results for the breach date are found when routine patients are scheduled once a week, with the best
of these being when urgent patients are scheduled either once or twice per week. As the breach date is the least
restrictive target (it is the largest target date), it is possible to achieve better results by slightly delaying the
creation of schedules for patients, so that schedules are created only once a week to increase the search space.
For the JCCO maximum acceptable target, the best results are found when creating a schedule for urgent patients between
one and three times per week and for routine patients once per week, or for urgent patients once per week and for
routine patients twice a week. For a large portion of urgent patients, it is impossible to meet this target due to a
large time interval between the decision to treat and the release date. This is likely why the frequency of creating
schedules for urgent patients does not have such a large influence on the quality of this criterion.
When considering the JCCO good practice target, the best results were obtained either creating schedules for urgent
patients five times per week and for routine patients twice or less, or for urgent patients three times per week and
for routine three times or less, or for urgent patients twice a week. It is possible to see a pattern with the value of
the SCD parameter and the restrictiveness of each target date. The more restrictive the target date is, the more
frequently the creation of schedules achieve better results. This makes sense since the more restrictive the target
date is, the more likely it is that the patient's release date will have already passed when the schedule is created
for that patient if the schedule for them is created with a low frequency.
For the squared waiting time, the best results are found when routine patients are scheduled once a week, similarly to
the breach date. The authors believe that the configurations that achieve the best results for the breach date also
achieve the best results for the squared waiting time because both criteria give a much greater penalty to patients who
have large waiting times than patients with shorter waiting times.
When using a lexicographical approach, it is possible that the first criteria constrain the search space in a way that
there is no room for improvement for the other criterion. However, this is rarely the case in the problem investigated
in this work, as the criteria presented are increasingly more restrictive. Apart from situations where the criteria
value is the same for all feasible solutions (e.g. a day when all patients have their release date after their JCCO
good practice target), objectives are improved in around 56\% of the times.
The values for the SCD parameter of 2/1 for urgent/routine patients achieve the best results for all criteria and are,
therefore, used in the remaining experiments.
\subsection{Experiments with Maximum Number of Days in Advance (MNDA)}
Two important dates in radiotherapy treatment scheduling are the decision to treat date and the release date. Since
these two dates can be very far apart, one possibility of achieving a better schedule is to not create a schedule for
the patients immediately when they arrive in the radiotherapy centre, but to wait for the release date to become
closer, i.e. towards the end of their pre-treatment phase. This might give better chances of good quality schedules for
patients that will arrive in the near future, while still obtaining good performance for current patients.
The maximum number of days in advance (\emph{MNDA}) parameter is introduced to limit the creation of schedules based on
the patient's release date. The values used in the experiments for the MNDA parameter of urgent and routine are:
\begin{itemize}
\item $\infty$ (infinity) - the schedule is created as soon as the patient arrives,
\item 21, 14, 7 - the schedule is created when the release date is within 21, 14 or 7 days, respectively,
\item and 0 - the schedule is created on the release date or afterwards.
\end{itemize}
For emergency patients, the MNDA value is fixed at $\infty$. Results are presented in \autoref{tab:mnda}.
\begin{table*}[tb]
\centering
\begin{tabular}{cc|cccc}
\hline
\multicolumn{2}{c|}{MNDA} & Breach (\%) & JMax (\%) & JGood (\%) & Waiting \\
\cline{1-2}
Urgent & Routine & & & & \\
\hline \hline
$\infty$ & $\infty$ & 33.95 & 47.06 & 86.41 & 1,624 \\
$\infty$ & 21 & 29.84 & 45.55 & 86.11 & 1,416 \\
$\infty$ & 14 & 27.72 & 44.04 & 85.95 & 1,293 \\
$\infty$ & 7 & \textbf{26.81} & 40.92 & 85.75 & 1,155 \\
$\infty$ & 0 & 31.46 & 40.25 & \textbf{85.60} & \textbf{1,122} \\
21 & $\infty$ & 33.96 & 47.04 & 86.41 & 1,624 \\
21 & 21 & 29.85 & 45.53 & 86.11 & 1,416 \\
21 & 14 & 27.73 & 44.02 & 85.94 & 1,293 \\
21 & 7 & \textbf{26.82} & 40.90 & 85.75 & 1,155 \\
21 & 0 & 31.47 & 40.23 & \textbf{85.60} & \textbf{1,122} \\
14 & $\infty$ & 34.00 & 46.97 & 86.40 & 1,628 \\
14 & 21 & 30.00 & 45.42 & 86.10 & 1,419 \\
14 & 14 & 27.90 & 43.87 & 85.92 & 1,307 \\
14 & 7 & \textbf{26.84} & 40.76 & 85.73 & 1,155 \\
14 & 0 & 31.51 & \textbf{40.01} & \textbf{85.60} & \textbf{1,136} \\
7 & $\infty$ & 34.38 & 46.85 & 86.38 & 1,684 \\
7 & 21 & 30.82 & 45.38 & 86.08 & 1,500 \\
7 & 14 & 28.53 & 44.14 & 85.90 & 1,346 \\
7 & 7 & 27.30 & 41.16 & \textbf{85.65} & 1,191 \\
7 & 0 & 31.64 & 40.48 & \textbf{85.57} & 1,169 \\
0 & $\infty$ & 35.44 & 47.66 & 88.73 & 1,772 \\
0 & 21 & 32.18 & 46.16 & 88.45 & 1,587 \\
0 & 14 & 29.85 & 45.49 & 88.36 & 1,421 \\
0 & 7 & 28.05 & 43.69 & 88.21 & 1,267 \\
0 & 0 & 32.30 & 43.61 & 88.31 & 1,225 \\
\hline
\end{tabular}
\caption{Results obtained varying the antecedence with which schedules are created.}
\label{tab:mnda}
\end{table*}
For the breach date, the best results are obtained when creating schedules for urgent patients when their release date
is within 14 or more days and for routine patients when their release data is within 7 days. It is possible to see a
large difference in the number of late patients between creating schedules when the release date is within 7 days and
on the release date or afterwards (MNDA values 7 and 0, respectively). This can be explained by the fact that when
creating schedules once per week on the release date or after it (MNDA value 0), there might be a patient whose
schedule is created only after the breach date, while if the schedule had been created when the release date was within
7 or more days, it would have been created before the breach date, thus not violating it.
For the JCCO maximum acceptable target date, which is slightly more restrictive than the breach date, the best results
are obtained by scheduling the urgent patients when their release date is within two weeks and routine patients either
on their release date or after. This way, it is possible to give a higher priority for creating schedules for the
urgent patients, without compromising the schedules of emergency patients.
The JCCO good practice is the most restrictive target date. The best results for this criterion are found when creating
schedules for urgent patients when their release date is within 7 or more days and routine patients on their release
date or afterwards, or when creating schedules for both types of patients when their release date is within 7 days.
The values of the squared waiting time criterion vary similarly to the values of the JCCO good practice objective
function. Good results are achieved when creating schedules for urgent patients within 14 or more days of their release
date and for routine patients on their release date or after. These values lead to schedules where only few patients
have very large waiting times.
Using the parameters values of 14/0 for urgent/routine patients achieves the best results for 3 of the 4 examined
criteria. However, the authors recommend using the values $\infty$/7 instead, as they achieve the best value for the
breach criterion, and values very close to the best values obtained for the remaining criteria. In comparison, using
the values 14/0 achieves the best values for the JCCO maximum acceptable, good practice and squared waiting time, but
the values achieved for the breach criteria are considerably worse than the best found.
\section{Conclusions and Future Work}
\label{sec:conclusion}
This research investigates the problem of scheduling the treatments of radiotherapy patients in the Nottingham
University Hospitals NHS Trust, UK. A new integer linear programming model with four optimisation criteria is
formulated and data for the problem is generated based on real-world data from the hospital.
Throughout this paper, different policies for scheduling treatments for radiotherapy patients were experimented with.
It was demonstrated to be possible using an efficient optimisation tool, such as CPLEX, for finding a schedule of good
quality. It might be worth waiting for an accumulation of patients in order to increase the search space and find a
schedule of good quality for a larger number of patients. This can be done, for instance, by using this tool once per
week for creating schedules for patients.
The experiments also suggest that it is better to not create schedules for patients immediately when they arrive, but
to wait for their release date to become closer. It is understood that patients prefer to know their treatment schedule
as soon as possible, but in view of the possible improvements in the quality of their schedule, the authors recommend
waiting for their release date to become closer before creating their schedule. The best results in the experiments
were achieved by creating schedules for routine patients only when their release date was within 7 days.
Future work includes the investigation of look-ahead techniques to try to anticipate how many patients of each category
might arrive in the succeeding days. Criteria to evaluate the robustness of a solution with regard to future patients
can also be investigated as well as the implementation of rescheduling policies.
So far, we have considered allocation of treatment days to patients. More constraints will be included to realistically
capture the real-world radiotherapy scheduling problem, such as considering the random machine down times, patient
preference for being treated in morning or afternoon sessions, requirements for transportation to and from the
hospital, etc.
The implementation of a graphical user interface for every day use in the hospital is currently in development, and it
will be placed for use in the hospital at the end of the project.
\section*{Acknowledgements}
The authors would like to thank the Nottingham University Hospitals NHS Trust, UK, and the Engineering and Physics
Sciences Research Council (EPSRC), UK for supporting this research (Ref No. EP/C549511/1).
|
\section{Introduction}
When Brownian particles are confined along a line in a quasi one dimensional channel so narrow that they cannot cross each other, anomalous diffusion appears and strongly subdiffusive behaviour can be observed. This phenomenon called Single File Diffusion was first noticed in 1955 by Hodgkin and Keynes~\cite{Hodgkin55} who were studying water transport through molecular-sized channels in biological membranes. Since then, SFD also appeared in the diffusion of molecules in porous materials like zeolites~\cite{Gupta95,Hahn96,Chou99}, of charges along polymer chains~\cite{Wang09}, of ions in electrostatic traps~\cite{Seidelin06}, of vortices in band superconductors~\cite{Besseling99,Kokubo04} and of colloids in nanosized structures~\cite{Wei00,Cui02,Lin02,Lin05,Koppl06,Henseler08} or optically generated channels~\cite{LutzJPC04,LutzPRL04}. Even though SFD can be encountered in a lot of various physical systems, most of the theoretical studies devoted to it are generally restricted to the simplest case~: an infinite overdamped system with hard core interactions.
In this paper, we present simulations results concerning finite systems of long range interacting particles. In particular, we focus on the dependency of the diffusion properties with the number of particles $N$ and the damping coefficient $\gamma$. We exhibit new behaviors, specifically associated to systems of small number of particles and to small damping. In order to interpret those results, we present an original analysis based on the decomposition of the particles motion in the normal modes of the chain.
In the thermodynamic limit (infinite systems with finite density $\rho$), for overdamped dynamics with hard core interactions, several analytical models~\cite{Harris65,Levitt73,vanBeijeren83,Arratia83,DeGennes71} predict that at long times, the mean square displacement of a particle of mass $m$ grows as $F_{H} \sqrt{t}$ with the mobility $F_{H}$ given by
\begin{equation}
F_{H}=\frac{2}{\rho} \sqrt{\frac{D_0}{\pi}} = \frac{2}{\rho} \sqrt{\frac{k_B T}{\pi m \gamma}},
\label{eq:mobhard}
\end{equation}
with $D_0 = k_B T/( m \gamma)$ the single particle free diffusion coefficient at temperature $T$, and $k_B$ Boltzmann's constant. If the interactions are long-ranged, only two analytical approaches have been undertaken so far~\cite{Kollmann03,Sjogren07}, for overdamped systems in the thermodynamic limit. There it is proven that the m.s.d. grows as $F_S \sqrt{t}$ at long times, with a mobility $F_S$ that depends on the interaction potential, through the isothermal compressibility $\kappa_T$ \cite{Coste10} or the spring constant $K \equiv \rho/\kappa_T$,
\begin{align}
\label{eq:mobform}
F_S = {2\over \rho} S(0,0)\sqrt{{ D_{eff} \over \pi}} = 2 k_B T \sqrt{\frac{ \kappa_T}{\pi m \gamma \rho}} = 2 k_B T \frac{ 1}{\sqrt{\pi m \gamma K}},
\end{align}
where $S(0,0) \equiv S(q \to 0,t = 0)$ is the long wavelength static structure factor of the particles. The diffusivity $D_{eff}$ is the effective diffusivity of a Brownian particle, taking into account its interactions with the other Brownian particles \cite{Nagele96}, and differs from the single particle diffusivity $D_0$. In its last version, the expression~\eqref{eq:mobform} can be interpreted by considering that we can derive $F_S$ from $F_H$ if we replace the interparticle distance $1/\rho$ by the mean square displacement $k_B T/K$ of a particle in the potential well due to its neighbors. In appendix~\ref{sec:asympt}, we recover the formula \eqref{eq:mobform} without the assumption of overdamped Langevin dynamics \refpar[langevinbase].
In numerical simulations and experiments, the systems are obviously finite. Periodic boundary conditions are used in simulations, and annular geometries in experiments. As a consequence of finite size effects, the asymptotic behavior at long time is always $\langle\Delta x^2\rangle = D_N t$. All particles in the system are then totally correlated and diffuse as a single effective particle of mass $N \times m$ \cite{vanBeijeren83}. We will recover it from our analytical model and provide measurements of the diffusion coefficient $D_N$ in good agreement with this interpretation. The SFD regime may nevertheless be observed in finite systems if the damping and the particles number are high enough, in a manner that will be precised by an analytical approach of finite systems dynamics.
However, most theoretical and experimental studies have been performed for overdamped systems only. This is the initial assumption in the existing models for long range interacting particles \cite{Kollmann03,Sjogren07}. The relevant experiments were generally done with solutions of collo\"ids~\cite{Wei00,Cui02,Lin02,LutzJPC04,LutzPRL04,Lin05,Koppl06,Henseler08} for which overdamping is a safe assumption. The simulations \cite{Herrera07,Herrera08} are shown in \cite{Coste10} to be in good agreement with the theoretical prediction of Kollmann, but they also assume overdamping in the choice of the simulation algorithm. In order to explore \emph{underdamped} systems, we have previously studied the diffusion in a circular channel of millimetric steel balls electrically charged \cite{Coste10}. In this experiment, identical metallic beads are held in a plane horizontal condenser made of a silicon wafer and a glass plate covered with an optically transparent metallic layer. A constant voltage is applied to the electrodes, inducing a charge distribution of the beads. The condenser is fixed on loudspeakers excited with a white noise voltage, and we have checked that this mechanical shaking behaves as an effective thermal bath \cite{Coupier06,Coupier07,Coste10}. In this system the measurement of the damping constant $\gamma$ proved that the balls diffusion is underdamped \cite{Coupier06}. We have observed that the {\rm m.s.d.}~ of the particles exhibit the SFD scaling predicted for overdamped systems, with a prefactor that is only slightly higher than the theoretical prediction~\eqref{eq:mobform}. Unfortunately, we were not able to tune the damping constant experimentally. Thus, in order to investigate the specific role of damping on the diffusion of finite systems we have developped numerical simulations that allows easy changes of the damping constant.
The paper is organised as follows~: Section~\ref{sec:simulation} is devoted to the description of the algorithm used in our numerical simulation. In section~\ref{sec:results}, we present our numerical results and exhibit new behaviors specific to systems of small numbers of underdamped particles. We characterize the various regimes for the {\rm m.s.d.}~ scaling with time, and define the crossover times between those regimes. In section~\ref{sec:discussion}, we give a physical interpretation of those regimes in the framework of our analytical model. We recover the various scaling laws for the {\rm m.s.d.}~, and provide estimates of the various crossover times, showing their dependency on the damping and particle numbers. We summarize our results in section~\ref{sec:conclusion}. Two appendices are devoted to complementary calculations.
\section{Description of the simulation}
\label{sec:simulation}
\subsection{A line of particles with long-ranged interactions}
\label{subsec:model}
We consider point particles of mass $m$ in a horizontal plane $(xOy)$, submitted to a thermal bath at temperature $T$. The particles are confined by a quadratic potential in $y$ in such a way that they cannot cross each other, as if they were diffusing in a narrow channel (see Fig.~\ref{schem}). This lateral confinement is chosen to mimic as well as possible experimental situations. We have checked that its strength do not influence the system behavior provided the beads stay ordered.
\begin{figure}[htb]
\centering
\includegraphics[width=7cm]{schema2.eps}
\caption{\label{schem}Scheme of the system}
\end{figure}
We describe the dynamics with the Langevin equation. Let $\mathbf{r}_i = (x_i,y_i)$ be the position of the particle $i$. We do not take into account the gravity, thus describing horizontal systems. The particle is submitted to a confinement force $-\beta y_i \mathbf{e}_y$ of stiffness $\beta$ and to the interaction potential $U(\mathbf{r}_i)$, so that Langevin equation reads
\begin{align}
\label{langevinbase}
\ddot{\mathbf{r}}_i + \gamma \dot{\mathbf{r}}_i +\frac{\boldsymbol{\nabla} U(\mathbf{r}_i)}{m} + \frac{\beta}{m}y_i\mathbf{e}_y = \frac{\boldsymbol{\mu}(t)}{m}
\end{align}
with $\gamma$ the damping constant and $\boldsymbol{\mu}$ a random force. In our simulations, the random force has the statistical properties of a white gaussian noise. Therefore, its components on both axes must satisfy:
\begin{align}
\label{eq:wgn1}
\langle\mu_x(t) \rangle = 0,\quad \langle\mu_y(t) \rangle = 0,\quad \langle\mu_x(t) \mu_y(t')\rangle = 0,\\
\label{eq:wgn2}
\langle \mu_x(t)\mu_x(t') \rangle = \langle \mu_y(t)\mu_y(t') \rangle = 2 k_B T m \gamma \delta (t-t') \end{align}
where $k_B$ is Boltzmann's constant and $\langle\cdot\rangle$ means statistical averaging.
It is suitable to put those equations in dimensionless form, defining the following dimensionless variables: $t=\widetilde{t}/\gamma$ and $x=\widetilde{x}\sqrt{{k_B T}/({m \gamma^2})}$. It gives us
\begin{align}
\ddot{\widetilde{\mathbf{r}_i}} + \dot{\widetilde{\mathbf{r}_i}} + \widetilde{\boldsymbol{\nabla}} \widetilde{U}(\widetilde{\mathbf{r}_i}) + \frac{\beta}{m\gamma^2}\widetilde{y_i}\mathbf{e}_y = \widetilde{\boldsymbol{\mu}}(\widetilde{t})
\label{eq:langevinadim}
\end{align}
with the dimensionless quantities:
\begin{align}
\widetilde{U}(\widetilde{\mathbf{r}_i})=\frac{U(\widetilde{\mathbf{r}_i})}{k_B T} , \qquad
\widetilde{\boldsymbol{\mu}}(\widetilde{t})=\frac{\boldsymbol{\mu}(\widetilde{t})}{\sqrt{k_B T m \gamma^2}}
\end{align}
and the only nonzero correlation (\ref{eq:wgn2}) now reads
\begin{align}
\label{eq:wgn4nodim}
\langle \widetilde{\mu}_x(\widetilde{t})\widetilde{\mu}_x(\widetilde{t'}) \rangle = \langle \widetilde{\mu}_y(\widetilde{t})\widetilde{\mu}_y(\widetilde{t'}) \rangle = 2 \delta (\widetilde{t}-\widetilde{t'}).
\end{align}
For the sake of simplicity, we drop the "tildes" $\widetilde{\quad}$ in the rest of this section.
In order to allow a direct comparison between simulations and experiments, we take the same interaction potential as in our experimental set-up \cite{Coupier06,Coupier07,Coste10}. It reads
\begin{align}
U(\mathbf{r}_i)= {U_0} \sum_{j\ne i}K_0\left({\left\vert\mathbf{r}_i-\mathbf{r}_j\right\vert \over \lambda}\right),
\label{eq:gammabeads}
\end{align}
where $K_0$ is the modified Bessel function of second order and index $0$, $\lambda$ and $U_0$ two constants. In principle, the sum extends to all particles, but in practice the summation is limited to the first five neighbours of each particle, which ensures a relative precision better than $10^{-7}$ and reduces the calculation time.
To decrease the computation time further, we replace the Bessel functions in the expression of the force ${\displaystyle \mathbf{F}({\mathbf{r}_i})=-{\boldsymbol{\nabla}} {U}({\mathbf{r}_i}) = \sum_{j \neq i} F_{ij}(\left\vert\mathbf{r}_i-\mathbf{r}_j\right\vert){\mathbf{r}_i-\mathbf{r}_j\over \left\vert\mathbf{r}_i-\mathbf{r}_j\right\vert}}$ by asymptotic expressions,
\begin{equation}
\left\{
\begin{array}{l l l}
{\displaystyle{F}_{ij}(x)} & ={\displaystyle\frac{U_0}{{\lambda }} \left[-\frac{1}{x}+bx+cx\ln(x)\right]} & \mbox{for }x<1\\
\\
{\displaystyle{F}_{ij}(x)} & = {\displaystyle\frac{U_0}{{\lambda }} \left[\sqrt{\frac{\pi}{2 x}} e^{-x}\left(1+\frac{a}{x}\right) \right]} & \mbox{for }x>1
\end{array}
\right.
\label{eq:forceapprox}
\end{equation}
where $a$, $b$, and $c$ are constants. They are chosen in such a way that the force and its derivative are continuous for $x=1$, and that the force is equal to its actual value at this point. Those two approximations fit very well the actual force (see Fig.~\ref{pot}).
\begin{figure}
\centering
\includegraphics[width=6cm]{potentiel.eps}
\caption{\label{pot}Force approximation. The thick black line represents the actual force derived from eqn.~\eqref{eq:gammabeads}, the green and orange dotted lines are respectively the logarithmic and exponential approximations in eqn.~\eqref{eq:forceapprox}.}
\end{figure}
\subsection{Algorithm}
\label{sec:algorithm}
The simulation is based on the Gillespie algorithm~\cite{Gillespie96PRE,Gillespie96AJP} that allows a consistent time discretization of the Langevin equation (\ref{eq:langevinadim}). We introduce a time step value ${\Delta t}$, which for consistency has to be much smaller than any other characteristic time-scale of the system. In dimensionless units $\Delta t = 10^{-3}$. Then the velocities $\dot{{x_i}}({t}+{\Delta t})$ and $\dot{{y_i}}({t}+{\Delta t})$ are calculated from updating formula derived from (\ref{eq:langevinadim}),
\begin{equation}
\left\{
\begin{array}{l}
\dot{{x_i}}({t} +{\Delta t}) = \dot{{x_i}}({t}) - \bigl[\dot{{x_i}}({t}) + \boldsymbol{\nabla} {U}(r_i(t))\cdot\mathbf{e}_x\bigr]\times{\Delta t} + \sqrt{2 {\Delta t}} \times {\mu_x}({t})\\
\\
{\displaystyle\dot{{y_i}}({t} + {\Delta t}) = \dot{{y_i}}({t}) - \left[\dot{{y_i}}({t}) + \frac{\beta}{m \gamma^2} {y_i}({t})+ \boldsymbol{\nabla} {U}(r_i(t))\cdot\mathbf{e}_y\right]\times{\Delta t} + \sqrt{2 {\Delta t}} \times {\mu_y}({t})}
\end{array}
\right.
\label{eq:langevinnum}
\end{equation}
where $r_i(t) = \sqrt{x_i(t)^2 + y_i(t)^2}$. The positions ${x_i}({t}+ {\Delta t})$ and ${y_i}({t}+ {\Delta t})$ of all the particles are then calculated from
\begin{equation}
{x_i}({t}+{\Delta t})={x_i}({t})+ \dot{{x}_i}({t}){\Delta t}, \qquad {y_i}({t}+{\Delta t})={y_i}({t})+ \dot{{y}_i}({t}){\Delta t}.
\label{eq:langevinposnum}
\end{equation}
The components of the random noise ${\mu_y}$ and ${\mu_x}$ are sampled in such a way that they have the properties given by equations~(\ref{eq:wgn1}) and~(\ref{eq:wgn4nodim}), hence that they are unit normal random numbers.
We simulate systems of $N$ particles, with periodic boundary conditions. We get from \refpar[eq:langevinposnum] $N$ equivalent trajectories, because all beads play the same role. The system is simulated during a dimensionless time of $10^3$, which means $10^6$ time-steps. The quantity of interest is the mean square displacement (m.s.d.) along the $x$ direction,
\begin{equation}
\langle\Delta x^2(t)\rangle = \langle \left[x(t + t_0) - x(t_0) - \langle x(t + t_0) - x(t_0)\rangle\right]^2\rangle.
\label{eq:defmsd}
\end{equation}
where $t_0$ is an arbitrary initial time. The ensemble averaging is done on every beads, since they all play an equivalent role. Moreover, the phenomenon is assumed to be stationary, so that $\Delta x^2(t)$ do not depend on $t_0$. For a given time $t$, it makes thus sense to average on the initial time $t_0$. Let $n$ be the overall number of time-steps in one simulation, and $n_t = t/\Delta t$. Then the averaging on the initial time $t_0$ reads
\begin{equation}
\langle\langle\Delta x^2(t)\rangle_e\rangle_0 = \sum\limits_{i = 0}^{n - n_t}{\left\{x[(n_t+i)\Delta t] - x(i \Delta t)\right\}^2 \over n - n_t + 1} -\left( \sum\limits_{i = 0}^{n - n_t}{x[(n_t+i)\Delta t] - x(i \Delta t) \over n - n_t + 1}\right)^2,
\label{eq:calcavg}
\end{equation}
where the index $i$ is such that $t_0 = i \Delta t$, $\langle \cdot \rangle_e$ means ensemble averaging and $\langle \cdot \rangle_0$ means averaging on the initial time $t_0$. This way of averaging greatly improves the statistics when $n_t$ is smaller than $n$. We will use it henceforward, denoting it with the simplified notation $\langle \cdot \rangle$ except in appendix~\ref{sec:average} where it is specifically discussed.
\subsection{Orders of magnitude of the various parameters.}
In our simulations, we work at densities $\rho = 33$, $100$ and $533$ particles per meter. The temperature and interaction strength are such that $\Gamma$ ranges as in experiments \cite{Wei00,LutzPRL04,LutzJPC04,Coste10} and numerical simulations \cite{Herrera07,Nelissen07,Herrera08}. The interest of the simulations is to get acces to parameter values that are difficult or impossible to obtain experimentally. We vary the particles number $N$ between 32 and 1024. This last value is comparable to some simulations \cite{Herrera07,Herrera08} but much greater than in experiments \cite{Wei00,LutzPRL04,LutzJPC04,Coste10}. We vary the damping constant $\gamma$ between 0.1~s$^{-1}$ and 60~s$^{-1}$, extending the experimental range toward small values of $\gamma$. This is to be compared to the cut-off frequency of the chain (see section~\ref{sec:theory}). With our damping constant range, we get access to both the overdamped and underdamped dynamics of the particles, and are thus able to exhibit the subtle behaviors linked to underdamping.
\bigskip
We simulate the same system that was experimentally studied in \cite{Coste10}. In the experiments, the beads number $N$ vary between 12 and 37, and the density $\rho$ is 477, 566 or 654 particles per meter. The mean interparticle distance is thus such that $1.53<1/\rho < 2.10$~mm, to be compared to the range $\lambda = 0.48$~mm of the potential. The dimensionless potential energy is such that $6 < \Gamma < 55$. The damping constant $\gamma$ ranges between 10~s$^{-1}$ and 30~s$^{-1}$ (see \cite{Coupier06}, Fig.~6). For the experimental values of density and potential energy, the cut-off frequency ranges between $21$~s$^{-1}$ and 37~s$^{-1}$. Experimentally, we are thus in the underdamped regime, as was already noticed in \cite{Coste10}.
\section{SFD of finite systems~: the different regimes}
\label{sec:results}
In this section, we present our results about the evolution of the mean square displacement (m.s.d.) $\langle\Delta x^2(t)\rangle$ as a function of the time $t$, and focus on the effects of the particles number $N$ (at fixed density) and on the damping constant $\gamma$. Two typical examples are provided by Fig.~\ref{regimes}. The evolution of the m.s.d. may be described by power laws $\langle\Delta x^2(t)\rangle \propto t^\alpha$, with an exponent $\alpha$ that depends on the observation time. The interpretation that will be detailed in part IV allow us to regroup them into three different regimes~:
\begin{itemize}
\item during the first regime I, $0 \leq t \leq \tau_{\rm ball}$, the m.s.d. grows according to $H_1 t^2$;
\item during regime II, $\tau_{\rm ball} \leq t \leq \tau_{\rm coll}$, $\langle\Delta x^2\rangle$ may be proportional to $D t$ only, $F_S \sqrt{t}$ only or to $D t$ then $F_S \sqrt{t}$, depending on the parameters of the simulation. The coefficient $D$ is \emph{not necessarily} the free diffusion constant $D_0$. When both scalings $D t$ and $F_S \sqrt{t}$ are observed, we define the crossover time $\tau_{\rm sub}$ between them.
\item A final regime III takes place for $\tau_{\rm coll} \leq t$, $\langle\Delta x^2\rangle = D_N t$ at long times with $D_N \neq D$ and $D_N \neq D_0$. This final asymptotic behavior is sometimes preceded by the scaling $H_Nt^2$ with $H_N \neq H_1$, the crossover time being noted $\tau_{\rm lin}$ .
\end{itemize}
\begin{figure}[!ht]
\centering
\includegraphics[width=7cm]{linagainI.eps}
\includegraphics[width=7cm]{linagainII.eps}
\caption{\label{regimes}Evolution of the m.s.d. (in mm$^2$) according to the time (in s.) for a chain of 32 particles with density 533~m$^{-1}$, temperature $T=10^{12}$ K and interaction potential $\Gamma \approx 7$. The solid line scales as $t^2$, the dashed line scales as $t$, the dotted line scales as $t^{1/2}$.\\
(a) Damping constant $\gamma = 0.1$~s$^{-1}$. In this low damping case, regime II is characterized by a $t$ scaling, regime III by a $t^2$ then a $t$ scaling.\\
(b) Damping constant $\gamma = 60$~s$^{-1}$. In this strong damping case, regime II is characterized by a $\sqrt{t}$ scaling, regime III by a $t$ scaling.}
\end{figure}
\begin{figure}[!ht]
\centering
\includegraphics[width=6cm]{vargamma.eps}
\includegraphics[width=6cm]{varN.eps}
\includegraphics[width=6cm]{varV.eps}
\includegraphics[width=6cm]{varT.eps}
\caption{\label{variances} Plot of $\langle\Delta x(t)^2 \rangle$ (in mm$^2$) according to the time (in s.) for a density $\rho = 533$~m$^{-1}$. Unless otherwise specified, the parameters are $N = 32$, $T=10^{12}$~K, $\Gamma \approx 6.8$ and $\gamma = 10$~s$^{-1}$. Specific values are as follows~: (a) $\gamma=0.1\mbox{, }1\mbox{, }10\mbox{ and }60\mbox{ s}^{-1}$ (blue, red, green and orange respectively). (b) $\gamma=1 \mbox{ s}^{-1}$ and $N=4\mbox{, }16\mbox{, }64\mbox{ and }128$ (blue, orange, red and green respectively). (c) $\Gamma \approx 4.4\mbox{, }6.8\mbox{, }9.8\mbox{ and }13.4$ (green, red, orange and blue respectively). (d) $T=10^{10}\mbox{, }10^{11}\mbox{, }10^{12}\mbox{ and }10^{13}$~K, (green, red, orange and blue respectively). The black thick line is~\eqref{eq:solcorreltpscourt}, the dashed line is $F_S \sqrt{t}$ with the mobility $F_S$ given by \eqref{eq:mobform}
and the dotted line is either $D_N t$ with $D_N$ given by \eqref{cdiffN}, in (a), (c) and (d) or $D t$ with $D$ given by \eqref{eq:diffeffect} in (b). There are no free parameters in the calculations.}
\end{figure}
\subsection{The small time regime (regime I)}
\label{sec:regimeIdata}
Regime I is defined by an evolution $\langle\Delta x^2\rangle = H_1 t^2$. It is observed in all data displayed in Fig.\ref{variances}, and the prefactor $H_1$ is independent on the damping constant [see Fig.\ref{variances}--a)] , on the system size [see Fig.\ref{variances}--b)] and on the interaction potential $\Gamma$ [see Fig.\ref{variances}--c)]. In this time range ($0 \leq t \leq \tau_{\rm ball}$), each particle behaves independently from the others and ensures a ballistic flight at its thermal velocity $\sqrt{k_B T/m}$, so that the constant $H_1$ should thus be equal to $k_B T/m$. The duration of this first \emph{ballistic regime} is called $\tau_{\rm ball}$. From our data summarized in Fig.\ref{variances}, we measure the constant $H_1$ and show in Fig.~\ref{prefactors}--a) that it is indeed in perfect agreement with its predicted value.
This behavior is obviously not observed in the simulations of the overdamped Langevin equation \cite{Herrera07,Herrera08,Henseler08,Centres10}, but as already been seen in simulations of the full dynamics \cite{Taloni08}. In \cite{Nelissen07}, they simulate the full Langevin equation but they do not display data for a sufficiently small time to observe the $t^2$ scaling.
\subsection{The intermediate time regime (regime II)}
\label{sec:regimeIIdata}
If we consider now the second regime, two different behaviors with distinct power laws can be observed~: Fig.~\ref{variances}--a) shows that for the highest values of $\gamma$, $\langle\Delta x^2\rangle$ only grows as $\sqrt{t}$. When $\gamma$ is decreased, a linear evolution in $D t$ appears for $\tau_{\rm ball} < t <\tau_{\rm sub}$. For the lowest values $\gamma = 1$~s$^{-1}$ and $\gamma = 0.1$~s$^{-1}$, the $\sqrt{t}$ scaling completely disappears. This is clearly a \emph{finite size effect}, as seen in Fig.~\ref{variances}--b). The data displayed in this picture are recorded at a low value $\gamma = 1$~s$^{-1}$, and the $\sqrt{t}$ scaling is not recovered until large numbers of particles, typically $N> 128$. Data from simulations with 256, 512 and 1024 particles (at constant density) superimpose exactly on the data for 128 particles.
We could be tempted to explain the linear evolution in $D t$ by arguing that we observe the diffusion of a free particle, that needs a finite time to feel the effect of confinement. If this should be the case, the diffusion coefficient $D$ should be the diffusion constant for a free particle, which is:
\begin{align}
\label{cdiff}
D_0=\frac{k_B T}{m \gamma}
\end{align}
In Fig.~\ref{prefactors}--b), we compare our numerical values of $D$ to $D_0$. It is obvious that $D$ is very different from $D_0$ except at the lowest values of the density $\rho$ (that is, low interactions). We shall see in section~\ref{sec:regimeIIdiscussion} that for high interactions the coefficient $D$ actually results from a collective behavior of the particles. In our model [see eqn.~\eqref{eq:appvariancelowgamma}], when $\rho$ is high (high interactions), the coefficient $D$ doesn't depend upon $\gamma$ and is given by
\begin{align}
\label{eq:diffeffect}
D = \frac{k_B T}{2 \pi} \sqrt{\frac{\kappa_T}{m \rho}} = \frac{k_B T }{2 \pi\sqrt{m K}}.
\end{align}
In Fig.~\ref{prefactors}--c), we see that at high density the coefficient $D$ is actually a function of $k_B T/\sqrt{m K}$, but with a numerical coefficient that is rather equal to $1/2$. The dependency of $D$ on either the spring constant $K = U''(1/\rho)$ or the compressibility $\kappa_T = \rho/K$ indicates that collective phenomenons are responsible of this behavior, and that the "free particle" hypothesis does not account for the linear behavior observed in strongly interacting systems. The modified Bessel function $K_0$ which gives the behavior of the potential $U(1/\rho)$ [see~\eqref{eq:gammabeads}] is a very quickly increasing function of the density, which explains why this behavior is typical of high density systems.
When the subdiffusive regime $\langle\Delta x^2\rangle=F\sqrt{t}$ is observed, as in Fig.~\ref{variances}--c) and Fig.~\ref{variances}--d), we may measure the mobility $F$. As seen in Fig.~\ref{prefactors}--d), our numerical data are in excellent agreement with the expression $F_S$ given in formula \refpar[eq:mobform], even if our system is underdamped. We discuss in section~\ref{sec:regimeIIdiscussion} the relevance of this equality for \emph{finite} systems.
The numerical data displayed in Fig.~\ref{variances}--c) and Fig.~\ref{variances}--d) are calculated for parameters values that are very close (in particular, the system sizes are equal) to the relevant parameters of the experiments reported in \cite{Coste10}. One can check (see Fig.~5 of \cite{Coste10}) that the value of the {\rm m.s.d.}~ is the same. The duration of the experiments is unsufficient to see the final $D_N t$ scaling described in the next section.
\subsection{The large time regime (regime III)}
\label{sec:regimeIIIdata}
It is quite intuitive that, for very long times and finite systems, all particles become fully correlated and behaves as a single effective particle of mass $N \times m$. It is a property of the translationally invariant mode (see section~\ref{sec:theory}). For small values of $\gamma$ and $N$, the m.s.d. grows according to $H_N t^2$, with a prefactor $H_N$ that is different from the constant $H_1$ introduced in section~\ref{sec:regimeIdata} above. $H_N$ should thus be given by the resolution of the Langevin equation for a free particle of mass $N \times m$~:
\begin{align}
\label{cballN}
H_N=\frac{k_B T}{N m},
\end{align}
which is in very good agreement with our simulations as shown by Fig.~\ref{prefactors}--e).
At higher values of $\gamma$, and for larger systems, we only observe a linear evolution of the {\rm m.s.d.}~ $\langle\Delta x^2\rangle = D_N t$ as shown by Fig.~\ref{variances}--c) and \ref{variances}--d). Similarly, $D_N$ should be given by the resolution of the Langevin equation for a free particle of mass $N \times m$:
\begin{align}
\label{cdiffN}
D_N=\frac{k_B T}{N m \gamma}.
\end{align}
This expression is in very good agreement with our results Fig~\ref{prefactors}--f).
The interpretation of section~\ref{sec:regimeIIIdiscussion} will show that this regime is dominated by the \emph{collective} behavior of the particles, so that we call $\tau_{\rm coll}$ the time at which these collective behavior takes place. We recover the values of $D_N$ and $H_N$ in section~\ref{sec:regimeIIIdiscussion} from our analytical solution \eqref{eq:solcorrel}, and give an estimate showing that this quadratic scaling is favored by small damping constants and small particles number, as is the case in Fig.~\ref{variances}--a) and \ref{variances}--b).
\begin{figure}[!ht]
\centering
\includegraphics[width=6cm]{Htot.eps}
\includegraphics[width=6cm]{coeffdiff0tot_a.eps}
\includegraphics[width=6cm]{coeffdiff0tot_b.eps}
\includegraphics[width=6cm]{mobilitestot.eps}
\includegraphics[width=6cm]{Hntot.eps}
\includegraphics[width=6cm]{coeffdifftot.eps}
\caption{\label{prefactors}a) Coefficient $H_1$ as a function of $k_B T/m$.\\
b) Coefficient of diffusion $D$ according to $k_BT/(m \gamma)$ [see~\eqref{cdiff}]. The green squares represent $D$ for systems of low densities ($\rho \approx 100 \mbox{ and } 33 \mbox{ m}^{-1}$) and the purple circles for systems with higher densities ($\rho \approx 533 \mbox{ m}^{-1}$).\\
c) Coefficient of diffusion $D$ according to $k_B T\sqrt{\kappa_T}/(2 \sqrt{m \rho})$ [the numerical coefficient is slightly different from that of~\eqref{eq:diffeffect}]. The green squares represent $D$ for systems of low densities ($\rho \approx 100 \mbox{ and } 33 \mbox{ m}^{-1}$) and the purple circles for systems with higher densities ($\rho \approx 533 \mbox{ m}^{-1}$)\\
d) Mobility $F_S$ according to $2 k_B T\sqrt{\kappa_T}/\sqrt{\pi m \gamma \rho}$ [see~\eqref{eq:mobform}]. The blue correspond to overdamped systems and the red triangles to underdamped systems.\\
e) Coefficient $H_N$ according to $k_BT/(N m)$ [see~\eqref{cballN}].\\
f) Coefficient of diffusion $D_N$ according to $k_B T/(N \gamma)$ [see~\eqref{cdiffN}].\\
All axes in mm$^2\cdot$s$^{-1}$ and logarithmic scales. All dotted lines are of slope 1.}
\end{figure}
\subsection{Crossover times}
\label{sec:crossovertimes}
Now that we know the evolution of $\langle\Delta x(t)^2\rangle$ in the different regimes , we can estimate the different crossover times. In this section, we will proceed heuristically, defining the crossover times by requiring continuity of the {\rm m.s.d.}~ for successive scalings.
Following this method, the ballistic time $\tau_{\rm ball}$ will be the time for which the curves of equation $H_1 t^2$ and $2Dt$ will intersect. Depending on the expression of the diffusion coefficient $D$, we obtain:
\begin{align}
\frac{k_B T}{m} \tau_{\rm ball}^2 \sim 2\frac{k_B T}{m \gamma}\tau_{\rm ball} \quad \Longrightarrow \quad
\tau_{\rm ball} \sim \frac{2}{\gamma}
\label{t1weak}
\end{align}
for weakly interacting systems. For strongly interacting systems, one has to considere the effective diffusion coefficient $D$ of \eqref{eq:diffeffect}, which gives
\begin{align}
\frac{k_B T}{m} \tau_{\rm ball}^2 \sim \frac{k_B T}{\pi} \sqrt{\frac{\kappa_T}{m\rho}}\tau_{\rm ball} \quad \Longrightarrow \quad
\tau_{\rm ball} \sim \frac{1}{\pi}\sqrt{\frac{\kappa_T m}{\rho}}.
\label{t1strong}
\end{align}
We performed measurements of $\tau_{\rm ball}$ for different values of parameters and reported them on Fig.~\ref{tlin}--a) and Fig.~\ref{tlin}--b). One can clearly see that two different mecanisms are at stake: the green circles which represent systems of high densities ($\rho \approx 533 \mbox{ m}^{-1}$), which are associated to strongly interacting particles, can be easily distinguished from the red squares which represent systems with lower densities ($\rho \approx 100 \mbox{ and } 33 \mbox{ m}^{-1}$), thus weakly interacting particles. Formula~\eqref{t1weak} seems in good agreement with the transition times of weakly interacting systems whereas formula~\eqref{t1strong} fits the values of $\tau_{\rm ball}$ for strongly interacting ones.
Let us now consider $\tau_{\rm coll}$. For overdamped and large systems, it will be the time for which the curves of equation $F_S\sqrt{t}$ and $D_{N}t$ intersect, thus giving
\begin{align}
F_S \sqrt{\tau_{\rm coll}} &= 2D_N\tau_{\rm coll} \quad \Longrightarrow \quad \tau_{\rm coll} = \frac{F^2}{4D_N^2} = \frac{S(0,0)^2}{D_0^2} \times D_{eff} \times \frac{N^2}{\pi \rho^2}= \frac{m \gamma N^2 \kappa_T}{\pi \rho}.
\label{t2form}
\end{align}
Note that starting from equation~\eqref{eq:mobform}, using the fact that $D_{eff}=D_0/S(0,0)$, introducing the length $L = N/\rho$ of the chain, we may recast this expression to obtain
\begin{equation}
\tau_{\rm coll} = \frac{L^2}{\pi D_{eff}}
\end{equation}
which is interesting as it tells us that $\tau_{\rm coll}$ can be seen as the time necessary for a given particle to diffuse over the length of the system $L$, with the effective diffusion coefficient that takes into account its interactions with the other particles.
In the case of small damping and small systems, the SFD behavior is not observed in the intermediate regime [see Fig.~\ref{variances}--a) and Fig.~\ref{variances}--b)], being replaced by a $D t$ scaling with $D$ given by \eqref{eq:diffeffect}, and the collective regime begins by a $t^2$ evolution [see~\ref{sec:regimeIIIdata} and \eqref{cballN}]. The time $\tau_{\rm coll}$ may thus be estimated by
\begin{align}
2 D \tau_{\rm coll} = H_N\tau_{\rm coll}^2\quad \Longrightarrow \quad \frac{k_B T}{\pi \sqrt{m K}}\tau_{\rm coll} = \frac{k_B T}{N m}\tau_{\rm coll}^2 \quad \Longrightarrow \quad \tau_{\rm coll} = \frac{N}{\pi}\sqrt{\frac{m}{K}}.
\label{t2form2}
\end{align}
One can see on figure~\ref{tlin}--c) that $\tau_{\rm coll}$ is in very good agreement with equations~\eqref{t2form} for large values of $\gamma N^2$. For small values of $\gamma N^2$, the data rather follow~\eqref{t2form2}, according to the analysis provided in section~\ref{sec:regimeIIIdiscussion} and particularly~\eqref{eq:tpslongbis}.
\begin{figure}[!ht]
\centering
\includegraphics[width=6cm]{t1_a.eps}
\includegraphics[width=6cm]{t1_b.eps}
\includegraphics[width=6cm]{t2_a.eps}
\includegraphics[width=6cm]{t2_b.eps}
\caption{\label{tlin}a) Measures of the transition time $\tau_{\rm ball}$ according to $2/\gamma$ [Log scale; see~\eqref{t1weak}]. The green circles represent $\tau_{\rm ball}$ for systems of high densities ($\rho \approx 533 \mbox{ m}^{-1}$) and the red squares for systems with lower densities ($\rho \approx 100 \mbox{ and } 33 \mbox{ m}^{-1}$)\\
b) Measures of the transition time $\tau_{\rm ball}$ according to $\sqrt{\kappa_T m/\rho}$ [Log scale; see~\eqref{t1strong}]. The green circles represent $\tau_{\rm ball}$ for systems of high densities ($\rho \approx 533 \mbox{ m}^{-1}$) and the red squares for systems with lower densities ($\rho \approx 100 \mbox{ and } 33 \mbox{ m}^{-1}$)\\
c) Measures of the transition time $\tau_{\rm coll}$ according to $m \gamma N^2 \kappa_T/\pi \rho$ [Log scale; see~\eqref{t2form}]. The blue circles represent $\tau_{\rm coll}$ for systems with $(N/2\pi)\sqrt{\gamma^2 m \kappa_T/\rho}>1$ and the purple diamonds for systems with $(N/2\pi)\sqrt{\gamma^2 m \kappa_T/\rho}<1$ (see~\eqref{eq:criterium}).\\
d) Measures of the transition time $\tau_{\rm coll}$ according to $N\sqrt{\kappa_T m/\rho}$ [Log scale; the numerical coefficient is slightly different from that of~\eqref{t2form2}]. The blue circles represent $\tau_{\rm coll}$ for systems with $(N/2\pi)\sqrt{\gamma^2 m \kappa_T/\rho}>1$ and the purple diamonds for systems with $(N/2\pi)\sqrt{\gamma^2 m \kappa_T/\rho}<1$ (see~\eqref{eq:criterium}).\\
All axes in s. All dashed lines are of slope 1.}
\end{figure}
\section{Theoretical analysis}
\label{sec:discussion}
\subsection{A chain of springs and point masses in a thermal bath}
\label{sec:theory}
In order to analyse the results presented in the last section, we have studied the Langevin dynamics of a chain of $N$ beads of mass $m$, aligned along the $x$ axis, interacting with a pair potential $U(x)$, with nearest neighbors interactions. Those two simplifying assumptions, a strictly 1D system with nearest neighbors interactions only, are as we will see in excellent agreement with the actual dynamics.
Small oscillations around the equilibrium position are described by linear springs of force constant $K = U''(1/\rho)$ where $\rho$ is the particle density at equilibrium \cite{Brillouin53}. Let $x(l,t)$ be the position of particle $l$ at time $t$. The equation of motion reads
\begin{equation}
{\mathrm{d}^2 \over \mathrm{d} t^2}x(l,t) = -\gamma{\mathrm{d} \over \mathrm{d} t}x(l,t) + {K \over m}\left[x(l+1,t) - 2x(l,t) + x(l-1,t)\right] + \frac{{\mu}(l,t)}{m},
\label{eq:chain}
\end{equation}
with the same notations as in \S~\ref{subsec:model}. Let's consider a chain with periodic boundary conditions. We may introduce the discrete Fourier transform
\begin{equation}
X(q,t) = \sum\limits_{l = 1}^N e^{i q l}x(l,t), \qquad x(l,t) = \frac{1}{N}\sum\limits_{k = 1}^N e^{-i q_k l}X(q_k,t),
\label{eq:Fourier}
\end{equation}
with $q_k = -\pi + {2\pi k / N}$ for $k = 1, \ldots, N$. From now on, we simplify the notations, dropping the dependency of the modes $q_k$ on the natural number $k$ and replacing summations on $k$ by summations on $q$. The variance of the displacement $x$ may be calculated from the Fourier modes $X(q,t)$, as $\langle \Delta x^2 (t)\rangle = \sum_q \langle \Delta X^2(q,t) \rangle/N^2$ with
\begin{equation}
\langle \Delta X^2(q,t) \rangle \equiv \left\langle \bigl[X(q,t) - \langle X(q,t)\rangle\bigr]\bigl[X(-q,t) - \langle X(-q,t)\rangle\bigr]\right\rangle.
\label{eq:defrmsFourier}
\end{equation}
Let us first considere the mode $q = 0$, that will be noted simply $X(t)$. It follows the equation
\begin{equation}
{\mathrm{d}^2 \over \mathrm{d} t^2}X(t) +\gamma{\mathrm{d} \over \mathrm{d} t}X(t) = \frac{{\mu}(q = 0,t)}{m}.
\label{eq:modezero}
\end{equation}
Physically, this is the equation for a free particle of mass $m$ in a thermal bath at temperature $T$, with damping constant $\gamma$. The solution is composed of two parts, $X_d(t)$ which corresponds to the deterministic motion of the particle, and the fluctuating part $X_\mu(t)$ which depends linearly on the random forcing $\mu(q = 0,t)$. It reads
\begin{equation}
X(t) - X^0 = \frac{\dot X^0}{\gamma}\left[1 - e^{-\gamma t}\right] + \frac{1}{m}\int\limits_0^t \mathrm{d} t' \int\limits_0^{t'} \mathrm{d} t'' e^{-\gamma(t' - t'')}\mu(q = 0,t''),
\label{eq:modezerosolu}
\end{equation}
where $X^0 \equiv X(t = 0)$ and $\dot X^0 \equiv \dot X (t = 0)$ are the initial conditions. The corresponding mean square displacement contribution measured in the simulations is the double average defined by \refpar[eq:calcavg]. It is calculated in the appendix~\ref{sec:average} and reads
\begin{equation}
\langle \Delta X^2 \rangle = 2\frac{N k_B T}{m \gamma}\left[t - \frac{1}{\gamma}\left(1 - e^{-\gamma t}\right)\right].
\label{eq:doubleavgzero}
\end{equation}
\bigskip
We now considere the modes $q \neq 0$. Using the periodic boundary conditions $x(l,t) = x(l+N,t)$, we see that each mode $X(q,t)$ of wave number $q\neq 0$ follows the equation
\begin{equation}
{\mathrm{d}^2 \over \mathrm{d} t^2}X(q,t) +\gamma{\mathrm{d} \over \mathrm{d} t}X(q,t) + {2K \over m}(1 - \cos q)X(q,t) = \frac{{\mu}(q,t)}{m}.
\label{eq:mode}
\end{equation}
The roots of the characteristic polynomial associated to this equation are
\begin{equation}
\omega_\pm(q) \equiv -{\gamma \over 2} \pm \sqrt{{\gamma^2 \over 4} - \omega_q^2},\qquad \omega_q^2 \equiv 2{K \over m}(1 - \cos q).
\label{eq:defomega}
\end{equation}
Physically, the modes $X(q,t)$ behaves as a particle of mass $m$ in an harmonic potential well with pulsation $\omega_q$, forced by the random force $\mu(q,t)$. The solution of \refpar[eq:mode] is readily obtained as
\begin{equation}
X(q,t) = \frac{\dot X(q,0) + \omega_-(q)X(q,0) }{ \omega_+(q) - \omega_-(q)}e^{\omega_+(q)t} + \frac{\omega_+(q)X(q,0) - \dot X(q,0) }{ \omega_+(q) - \omega_-(q)}e^{\omega_-(q)t} + X_\mu(q,t).
\label{eq:solcorrelnonnul}
\end{equation}
The translationally invariant mode $q = 0$ scales as $t^2$ at small time $t \leq 1/\gamma$, then scales as $t$ when $t \gg 1/\gamma$. The modes with nonzero wave number scales as $k_B T t^2/(N m)$ at small time, and saturate toward the constant value $2 k_B T/(N m \omega_q^2)$ at very large time.
\begin{figure}[htb]
\centering
\includegraphics[width=6.5cm]{reldispschema.eps}
\caption{\label{fig:reldisp} Dispersion relation $\omega(q)$ as a function of $q$, for $q \geq 0$ (the curve is obviously symmetric for $q \leq 0$). The continuous line is valid for the infinite chain, the dots represent the modes for $N = 32$. The modes such that their frequency is less than $\gamma/2$ are overdamped, the other ones are underdamped.}
\end{figure}
At intermediate time, the behavior of the modes $q \neq 0$ is determined by the relative values of $\omega_q$ and $\gamma/2$, as illustrated in Fig.\ref{fig:reldisp}.\\
-- The modes such that $\omega_q < \gamma/2$ are overdamped. They reach their saturation value at a time $t_{\rm sat} \sim 1/\vert \omega_+(q)\vert$. The shortest saturation time is associated to $q = \pi$, and reads $t_{\rm sat}^{\rm min} \sim \gamma/\omega_{\pi}^2$ if $\omega_\pi <\gamma/2$ or $t_{\rm sat}^{\rm min} \sim 2/\gamma$ otherwise.\\
-- The modes such that $\omega_q > \gamma/2$ are underdamped. They oscillate at a frequency $\omega(q) \equiv \sqrt{\omega_q^2 - \gamma^2/4}$. Below a time which is roughly $1/\omega(\pi)$, all underdamped modes scale as $t^2$.\\
Finally, all modes $q \neq 0$ saturate toward $2 k_B T/(N m \omega_q^2)$ at large time.
We do not distinguish the overdamped ($\gamma > 2\omega_q$) and the underdamped ($\gamma < 2\omega_q$) modes, because the final result \refpar[eq:flucqnonnul] is in both cases a real function with the same formal expression. The averaging process is explained in appendix~\ref{sec:average}. One gets
\begin{equation}
\langle \Delta X^2(q,t) \rangle = {2 N k_B T \over m \omega_q^2}\left[1 + {\omega_-(q)e^{\omega_+(q)t} \over \omega_+(q) - \omega_-(q)} - {\omega_+(q)e^{\omega_-(q)t} \over \omega_+(q) - \omega_-(q)}\right].
\label{eq:flucqnonnul}
\end{equation}
We remark that the limit $q \to 0$ is not singular, and by taking it properly in this expression one recovers \refpar[eq:doubleavgzero]. We think that it is nevertheless physically convenient to distinguish between the translationally invariant mode $q = 0$ and the others, because they lead to different asymptotic behaviors.
Using \refpar[eq:flucqnonnul] together with \refpar[eq:doubleavgzero], we obtain the mean square displacement as
\begin{equation}
\langle \Delta x^2 (t)\rangle = {2 k_B T\over N m}\left\{\frac{t}{\gamma} - \frac{1}{\gamma^2}\left(1 - e^{-\gamma t}\right) + \sum\limits_{q\neq 0} {1 \over \omega_q^2}\left[1 + {\omega_-e^{\omega_+t} \over \omega_+ - \omega_-} - {\omega_+e^{\omega_-t} \over \omega_+ - \omega_-}\right]\right\},
\label{eq:solcorrel}
\end{equation}
which will be the basis for the following discussion.
In the limit of very large damping, that is in the absence of the inertial term, a solution of \refpar[eq:chain] have been provided by L. Sj\"ogren \cite{Sjogren07}. We have thus extended his calculations to underdamped systems~\footnote{More precisely ${\protect\langle} \Delta x^2 (t){\protect\rangle}$ is equal to the correlation $C(0,t)$ defined by Sj{\protect\"o}gren \cite{Sjogren07}, and to $2 W(t)$ introduced by Kollmann \cite{Kollmann03}. This is easily seen from the definition~\eqref{eq:defmsd}, because with the double averaging on the initial conditions and on the random noise [see appendix~\ref{sec:average} and eqns.~\eqref{eq:modezerosolu} and \eqref{eq:solcorrelnonnul}] we get ${\protect\langle}{\protect\langle} x(t + t_0) - x(t_0){\protect\rangle}_e{\protect\rangle}_0 = 0$, and then stationarity ensures ${\protect\langle}{\protect\langle}[x(t + t_0) - x(t_0)]^2{\protect\rangle}{\protect\rangle} = {\protect\langle}{\protect\langle} [x(t ) - x(0)]^2{\protect\rangle}{\protect\rangle}$ which is precisely the definition of Sj{\protect\"o}gren and Kollmann.}, and calculated the r.m.s. displacement in a different way to take into account our peculiar way of averaging \refpar[eq:calcavg]. We also extend this discussion, in the rest of this section, to the case of finite systems.
\subsection{The ballistic regime (regime I)}
\label{sec:regimeIdiscussion}
Since the small time behavior of each mode in \refpar[eq:solcorrel] is $(k_B T / N m)t^2$, and there are $N$ equivalent contributions to the sum we get
\begin{equation}
\langle \Delta x^2 (t)\rangle \stackrel{t \to 0}{\sim} { k_B T\over m}t^2.
\label{eq:solcorreltpscourt}
\end{equation}
This result is independent of $N$, and thus valid in the thermodynamic limit too. Because of the inertial term in the Langevin equation \eqref{eq:chain}, at very small time each particle behaves independently from the others and undergoes ballistic flight at her thermal velocity $\sqrt{k_B T/m}$.
In order to discuss the duration $\tau_{\rm ball}$ of this regime, let us assume a finite, but large (in a sense to be precised later) particle number $N$. The time evolution of $\langle \Delta x^2 (t)\rangle$ is determined by the mode $q = 0$ and a summation on all modes $q \neq 0$. All modes in the summation \eqref{eq:solcorrel}, hence the sum itself, behaves as $t^2$ on a timescale such that
\begin{equation}
t \leq \tau_{\rm ball} \equiv \hbox{min}\left(\frac{2}{\gamma}, \frac{1}{\sqrt{\omega_\pi^2 - \gamma^2/4}},\frac{\gamma}{\omega_\pi^2}\right).
\label{eq:tundeux}
\end{equation}
For weakly interacting or equivalently low density systems $\tau_{\rm ball} \approx 2/\gamma$, which was already heuristically derived in \eqref{t1weak}, and observed in Fig.~\ref{tlin}--a). For strongly interacting or equivalently high density systems, we get $\tau_{\rm ball} \approx 1/\omega_\pi \propto \sqrt{m/K} \propto \sqrt{m \kappa_T/\rho} $, in perfect agreement with our observations Fig.~\ref{tlin}--b) and the heuristic derivations \eqref{eq:diffeffect} and \eqref{t1strong}.
\subsection{The collective regime (regime III)}
\label{sec:regimeIIIdiscussion}
This collective regime (regime III of Fig.~\ref{regimes}, see section~\ref{sec:regimeIIIdata}) is a property of finite systems only. This asymptotic behavior may be easily deduced from the sum \refpar[eq:solcorrel], which is dominated by the contribution of the mode $q = 0$ that scales as $[2 k_B T /(N m \gamma)] \times t$. This corresponds to the free diffusion of a particle of mass $N m$ in a thermal bath at temperature $T$. At very large time, the particles are completely correlated and behaves as a single particle of effective mass the sum of all masses. The same result has been obtained in \cite{vanBeijeren83}.
It is not difficult to estimate the time $\tau_{\rm coll}$. It is the time at which the contribution of the mode $q = 0$ dominates the sum of the contributions of all $N - 1$ other modes. Let us use the simplifying Debye approximation $\omega_q^2 = (K/m)q^2$. We thus get
\begin{equation}
\frac{2 k_B T}{N m \gamma}\tau_{\rm coll} \sim \sum\limits_{q \neq 0}\frac{2 k_B T}{N m \omega_q^2} \sim \frac{2 N k_B T}{4 \pi^2 K}\sum\limits_{i = 1}^{(N-1)/2}\frac{1}{i^2} \qquad \Longrightarrow \qquad \tau_{\rm coll} \sim \frac{N^2 m \gamma}{12 K},
\label{eq:tpslong}
\end{equation}
where we have used the fact that the sum is the generalized harmonic number $H_{(N-1)/2,2}$ which is equal to $\pi^2/6$ up to corrections of order $1/N$. The fact that this time scales as $N^2$ explains why this long time regime is seldom seen in simulations (to our knowledge, the only exception is \cite{Nelissen07}) or experiments. This is clearly illustrated by our Fig.~\ref{variances}--b), where we show that increasing $N$ shifts the long time regime toward longer times. This expression of $\tau_{\rm coll}$ is equal to our previous heuristic estimate \refpar[t2form]. It means that the reasoning at the basis of the derivation of eqn.~\eqref{eq:tpslong} includes in the right way the physical origin of the long time collective behavior of the finite chain.
As was already quoted in section~\ref{sec:regimeIIIdata}, at very small damping constant it is possible to observe at large time an evolution $\langle \Delta x^2 (t)\rangle = H_N t^2$. This is possible if the modes $q \neq 0$ are saturated while the mode $q = 0$ still evolves as $[ k_B T/(Nm)]t^2$. This requires $t < 1/\gamma$ (otherwise the mode $q = 0$ scales as $t$), and that the contribution of the $q\neq 0$ modes to the sum in \refpar[eq:solcorrel] be less than that of the mode $q = 0$. Roughly speaking, we get
\begin{equation}
\left.\left\langle \Delta x^2 \left(t = \frac{1}{\gamma}\right)\right\rangle\right\vert_{q = 0} \sim \frac{k_B T}{N m}\left(\frac{1}{\gamma}\right)^2 > \sum\limits_{q \neq 0}\frac{2 k_B T}{N m \omega_q^2} \sim \frac{N m k_B T }{12 K} \quad\Longrightarrow\quad \gamma^2 < \frac{12 K}{N^2 m},
\label{eq:tpslongbis}
\end{equation}
where we have used the Debye approximation to estimate the contribution of the $q\neq 0$ modes. This means that this regime is to be observed at small damping $\gamma$ and small particles number $N$, which is precisely the case in Fig.~\ref{variances}--a) and Fig.~\ref{variances}--b). In this case, the estimate~\eqref{eq:tpslong} should be replaced by
\begin{equation}
\frac{k_B T}{N m} \tau_{\rm coll}^2 \sim \sum\limits_{q \neq 0}\frac{2 k_B T}{N m \omega_q^2} \sim \frac{2 N k_B T}{4 \pi^2 K}\sum\limits_{i = 1}^{(N-1)/2}\frac{1}{i^2} \qquad \Longrightarrow \qquad \tau_{\rm coll} \sim N\sqrt{\frac{ m }{6 K}}.
\label{eq:tpslongbis}
\end{equation}
This expression of $\tau_{\rm coll}$ is equal to our previous heuristic estimate \refpar[t2form2], allowing us to interpret the $t^2$ scaling at long time, in systems of few particles with small damping, as a collective behavior linked to the translationally invariant mode. This regime takes place when the greatest saturation time of the modes $q \neq 0$ is smaller than the time above which the mode $q = 0$ evolves as $t$ rather than $t^2$. The behavior $\langle \Delta x^2 (t)\rangle = H_N t^2$ may thus be observable when
\begin{equation}
\frac{1}{\omega_{2\pi/N}} < \frac{1}{\gamma} \qquad \Longrightarrow \qquad \frac{N}{2\pi}\sqrt{\frac{\gamma^2 m \kappa_T}{\rho}} < 1.
\label{eq:criterium}
\end{equation}
The relevance of this estimate is proved by Fig.\ref{tlin}--c) and Fig.\ref{tlin}--d). It also show that the time $\tau_{\rm lin}$ introduced in section~\ref{sec:discussion} is equal to $1/\gamma$.
\begin{figure}[htb]
\centering
\includegraphics[width=7cm]{fig5amath.eps}
\caption{\label{fig:fig5amath} Plot of $\langle\Delta x(t)^2 >$ (in mm$^2$) according to the time (in s.), in logarithmic scale, calculated from the analytical solution~\eqref{eq:solcorrel} for a density $\rho = 533$~m$^{-1}$, $N = 32$, $T=10^{12}$~K, $\Gamma \approx 6.8$ and $\gamma = 0.1$s$^{-1}$. The solid line is of slope 2, the dashed line of slope 1. This plot is to be compared to the relevant plot in Fig.~\ref{variances}--a). We have indicated the times $\tau_{\rm ball}$ and $\tau_{\rm coll}$ respectively given by \eqref{eq:tundeux} and \eqref{eq:tpslongbis}. We see that both equation provides lower bounds for $\tau_{\rm coll}$.}
\end{figure}
\subsection{The correlated regime (regime II)}
\label{sec:regimeIIdiscussion}
Let us now discuss the intermediate regime~\footnote{In simulations of overdamped Langevin equation, for which the ballistic regime cannot be seen, this intermediate regime is the first to be observed \cite{Herrera07,Herrera08,Henseler08,Centres10}.} which, between the individual ballistic regime (regime I) and the collective behavior (regime III), exhibit the correlated behavior of the particles. While at asymptotically large time, all modes with finite (non zero) wave numbers have reached a constant value, in the intermediate regime, the physical behavior at a given time $t$ of the chain results from a subtle balance between the modes that are already saturated and those that still evolve. To simplify somewhat the discussion, we will considere the limit $\gamma \ll 2 \omega_{2\pi/N}$ when all modes are oscillating (underdamped) and the limit $\gamma \gg 2 \omega_\pi$, when all modes are overdamped.
Let us first assume a very low damping. The modes oscillates until they reach a stationary value. The time evolution of $\langle \Delta x^2 (t)\rangle$ is due to the progressive disappearance of the contributions of the first oscillation of those modes $q \neq 0$ that have reached their maximum value. In Fig.~\ref{fig:modes}--a) this is graphically illustrated with several underdamped modes ($\gamma/2 = 0.05$~s$^{-1} \ll \omega_\pi = 25$~s$^{-1}$), together with the complete sum~\eqref{eq:solcorrel}. As a first approximation,
\begin{equation}
\frac{1}{N^2}\left\langle \Delta X^2(q,t) \right\rangle \sim {2 k_B T \over N m \omega_q^2}\left[1 - e^{-\gamma t/2} \cos\omega(q) t\right] \sim { k_B T \over N m}t^2.
\label{eq:appmodelowgamma}
\end{equation}
At a given time $t$, the sum is dominated by the contributions of the modes that have not reached their first maximum, that is those modes such that $t < 1/\omega_q$. Let $n(t)$ be the number of such modes. In the Debye approximation $\omega_q = q\sqrt{K/m}$, the maximum wave number of those modes is $1/t\sqrt{K/m}$ so that an estimate of $n(t)$ is given by $n(t) \sim 2(N/2\pi)(1/t\sqrt{K/m})$ (the factor 2 takes into accounts the modes $\pm \vert q\vert$). The variance may thus be estimated as
\begin{equation}
\langle \Delta x^2(t) \rangle \sim { k_B T \over N m}t^2 \times n(t) \sim { k_B T \over \pi \sqrt{ m K}}t.
\label{eq:appvariancelowgamma}
\end{equation}
This is a normal diffusion with a diffusivity $k_B T /( 2\pi \sqrt{ m K})$ that depends on the stiffness of the interaction $K$, showing that it is a collective effect. This is in agreement with our observations, see Fig.~\ref{prefactors}--c).
In the opposite limit of a very strong damping, all modes evolve monotonously toward their saturation value. As a first approximation,
\begin{equation}
\frac{1}{N^2}\left\langle \Delta X^2(q,t) \right\rangle \sim {2 k_B T \over N m \omega_q^2}\left\{1 + \frac{\omega_-(q)[1 - \omega_+(q) t]}{\omega_+(q) - \omega_-(q)} \right\} \sim { 2 k_B T \over N m \gamma}t,
\label{eq:appmodehighgamma}
\end{equation}
where we have used $\omega_+(q) - \omega_-(q) \approx \gamma$, $\omega_+(q) \approx \omega_q^2/\gamma$ and $\omega_-(q) \approx -\gamma$. A mode is saturated at a time $t > \gamma/\omega_q^2$. At a given time $t$, in the Debye approximation, the modes that increase with time are such that $q < \sqrt{m \gamma/(K t)}$. These reasoning is graphically illustrated in Fig.~\ref{fig:modes}--a) where we show several overdamped modes ($\gamma/2 = 30$~s$^{-1} > \omega_\pi = 25$~s$^{-1}$), together with the complete sum~\eqref{eq:solcorrel}. Their number is thus $n(t) \sim 2(N/2\pi)\sqrt{m \gamma/(K t)}$. The contributions of all such modes thus gives
\begin{equation}
\langle \Delta x^2(t) \rangle \sim { 2 k_B T \over N m \gamma}t \times n(t) \sim { 2 k_B T \over N m \gamma}t \times \frac{2 N}{2\pi}\sqrt{\frac{m \gamma}{K t}} = \frac{2 k_B T }{\pi \sqrt{m K \gamma}} t^{1/2} = \frac{2 k_B T }{\pi} \sqrt{\frac{\kappa_T}{m \rho \gamma}} t^{1/2},
\label{eq:appvariancehighgamma}
\end{equation}
where we have introduced the isothermal compressibility $\kappa_T$ in the last expression to ease the comparison with the exact expression of $F_S$~\eqref{eq:mobform}. Taking into account the crudeness of our approximations, this estimate is extremely satisfactory because we recover the SFD behavior $\langle \Delta x^2(t) \rangle \propto t^{1/2}$, with a prefactor that is almost the exact one.
\begin{figure}[htb]
\centering
\includegraphics[width=7cm]{modesunderdamped.eps}
\includegraphics[width=7cm]{modesoverdamped.eps}
\caption{\label{fig:modes} (Color online) Thick solid line~: plot of $\langle\Delta x(t)^2 >$ (in mm$^2$) according to the time (in s.), in logarithmic scale, for a density $\rho = 533$~m$^{-1}$, $N = 32$, $T=10^{12}$~K, $\Gamma \approx 6.8$, calculated from \eqref{eq:solcorrel} for the relevant damping constant. (a) Underdamped regime, $\gamma = 1$~s$^{-1}$. The blue (gray) curves are, from top to bottom, the modes $q = \pi/16, q = \pi/4, q = 7\pi/16$. The dashed lines are, from left to right, of slopes 2 and 1. (b) Overdamped regime, $\gamma = 60$~s$^{-1}$. The blue (gray) curves are, from top to bottom, the modes $q = \pi/16, q = \pi/4, q = 7\pi/16, q = 5\pi/8$. The dashed lines are, from left to right, of slopes 2, $1/2$ and 1.}
\end{figure}
In the general case, the modes with wave number $q$ such that $\gamma < 2\omega_q$ contribute to a $t$ scaling of the {\rm m.s.d.}~, whereas the modes such that $\gamma > 2\omega_q$ contribute to the SFD behavior, that is a $t^{1/2}$ scaling of the {\rm m.s.d.}~ The typical time $\tau_{\rm sub}$ at which the \emph{subdiffusive} SFD behavior takes place is thus the inverse of this cut-off frequency, $\tau_{\rm sub} = 2/\gamma$. At a given particle number $N$, the minimum nonzero frequency is $\omega_{2\pi/N}$. If the damping constant is so small that all modes are underdamped ($\gamma < 2\omega_{2\pi/N}$) , the {\rm m.s.d.}~ scales as $t$ in the collective regime II. Increasing the damping, at fixed $N$, amounts to increase the number of overdamped modes, and favors the subdiffusive $t^{1/2}$ scaling for the {\rm m.s.d.}~ This is exemplified by Fig.~\ref{variances}--a). Increasing the particle number $N$, at fixed $\gamma$, amounts to decrease the frequency $\omega_{2\pi/N}$, hence to increase the number of overdamped modes, and favors the subdiffusive $t^{1/2}$ scaling for the {\rm m.s.d.}~ This is exemplified by Fig.~\ref{variances}--b).
The result \eqref{eq:appvariancehighgamma} is only approximate, so that the numerical prefactor cannot be trusted, but it explains under which conditions the SFD regime may be seen in finite systems with periodic boundary conditions [we remind the reader that this latter is the key assumption leading to \eqref{eq:mode}]. In appendix~\ref{sec:asympt}, we prove that the expression \eqref{eq:mobform} for the mobility of long ranged interacting systems is valid in the underdamped case $\gamma < 2\omega_\pi$ too, in the thermodynamic limit.
\section{Conclusion}
\label{sec:conclusion}
In this paper, we study the SFD of a chain of particles with long ranged interactions, without the simplifying assumption of overdamped dynamics. We have focused our discussion on finite size effects and the influence of low damping. We use numerical simulations of the Langevin equation with the Gillespie algorithm \cite{Gillespie96AJP,Gillespie96PRE} and modelize the system as a chain of linear springs (spring constant $K$) and point masses ($m$) in a thermal bath at temperature $T$.
In our simulations data, we have identified several regimes for the time evolution of the mean square displacement ({\rm m.s.d.}~ ) $\langle \Delta x(t)^2\rangle$. At small times ($0 \leq t \leq \tau_{\rm ball}$), it evolves as $\langle \Delta x(t)^2\rangle = (k_B T/m) t^2$. This is a ballistic flight that traces back to the inertial effects, and is observed whatever the damping $\gamma$ or the particle number $N$. We recover this behavior in the thermodynamic limit ($N \to \infty$ at finite density) from our model. The prefactor of the $t^2$ scaling measured in our simulations is in excellent agreement with the theory.
For finite systems with periodic boundary conditions, an intermediate regime ($ \tau_{\rm ball} \leq t \leq \tau_{\rm coll}$) takes place. Depending on the respective values of the damping constant and the number of particles, we may observe either a diffusive behavior, a SFD behavior or successively both. We provide a physical explanation of those observations when we express the motion of the chain in terms of normal modes of oscillations. The {\rm m.s.d.}~ $\langle \Delta x(t)^2\rangle$ of the chain results from the superposition of all those modes. The mode associated to the null wave number is always overdamped, and similar to the motion of a free particle in a thermal bath, since no restoring force is exerted on it. The modes of finite (non zero) wave numbers have the same dynamics as an oscillator in an harmonic well. At long time, the {\rm m.s.d.}~ of all modes with non zero wave numbers saturate toward a constant value. The overdamped modes, which do not oscillate until they saturate, contribute to the SFD scaling $\langle \Delta x(t)^2\rangle \propto t^{1/2}$. The underdamped modes oscillate before their saturation, and contribute to a linear scaling $\langle \Delta x(t)^2\rangle \propto t$.
In the thermodynamic limit, for systems of infinite number of particles, we exhibit analytically the SFD behavior $\langle \Delta x(t)^2\rangle = F_S t^{1/2}$ at asymptotically large time. We recover the mobility $F_S$ that was previously calculated for long ranged interactions and overdamped dynamics \cite{Kollmann03,Sjogren07}, thus extending the previous calculation to systems with arbitrary damping.
At asymptotically long time ($ t \gg \tau_{\rm coll}$), for a finite number of particles with periodic boundary conditions, the system behaves as an effective particle of mass $N \times m$. The physical origin of this behavior is the motion of the collective mode of null wave number, which is linked to the translational invariance of the system. For $t \geq \tau_{\rm lin} \geq \tau_{\rm coll}$, the system undergoes a linear diffusion with a diffusion coefficient $D_N = 2 k_B T/(N m \gamma)$. We show that this regime takes place whatever the value of $\gamma$. The duration of our simulations allows us to see this regime, and the measured diffusivity is in excellent agreement with its predicted value. We provide estimates of the time $\tau_{\rm coll} \sim N^2 m \gamma/K$ at large damping and $\tau_{\rm lin}\sim 1/\gamma$ at small damping. Those estimates are in good agreement with our simulations data. At small particle number and small damping, a new regime takes place at times $\tau_{\rm lin} \geq t \geq \tau_{\rm coll}$. It corresponds to the ballistic flight of the effective particle of mass $N \times m$, with $\langle \Delta x(t)^2\rangle = (k_B T/N m)t^2$. In this case, the time $\tau_{\rm coll} \sim N\sqrt{m/K}$.
\acknowledgements
We thank J. Mokhtar and F. van Wijland for helpful discussions.
|
\section{Introduction}
The coalgebraic approach to modal logic has been pursued
successfully over the last years. The basic ideas (see eg
\cite{moss:cl,pattinson:cml-j,schroeder:fossacs05,kurz:sigact06}),
are the following.
\begin{itemize}
\item A $T$-\emph{coalgebra}, consisting of a carrier $X$ and a
`next-step' map $\xi:X\rightarrow TX$, represents a transition system. For
example, with $\mathcal{ P} X$ the set of finite subsets of $X$ and
$\mathit{Act}$ a set of actions, $X\rightarrow\mathcal{ P}(\mathit{Act}\times X)$ is a
labelled transition system.
\item Any particular choice of $T$ yields a canonical notion of
$T$-\emph{bisimilarity}. For example, for $X\rightarrow\P(\mathit{Act}\times X)$
we obtain the Milner-Park notion of bisimilarity~\cite{aczel:nwfs}
whereas for $X\rightarrow\mathcal{D}(\mathit{Act}\times X)$, with $\mathcal{D} X$ denoting the
set of probability distributions on $X$, we obtain the
notion of bisimilarity described in \cite{vink-rutt:prob-bisim}.
\item Moreover, for any choice of $T$, we can find a logic for
$T$-coalgebras which is expressive (ie distinguishes non-bisimilar
states) and comes with a complete calculus. These logics are modal
logics in the sense that formulas are invariant under
$T$-bisimilarity.
\end{itemize}
The work on coalgebraic logic so far is focused on $T$-bisimilarity.
\medskip In parallel, Jacobs and collaborators
\cite{Jacobs04TraceSemantics,HasuoJacobsSokolova06GenericTraceTheory,Hasuo08PhD} showed that coalgebras not only
provide a framework for bisimilarity, but also for trace semantics:
\begin{itemize}
\item A $(B,T)$-coalgebra $X\rightarrow BTX$ is now given wrt a `transition
type' $T$ and a `branching type' $B$. For example, with $BX=\mathcal{ P} X$
and $TX=\{*\}+\mathit{Act}\times X$, a $X\rightarrow\P(\{*\}+\mathit{Act}\times X)$ is a
non-deterministic automaton.
\item Different choices of $B$ yield different notions of trace
semantics. With $B=\P$, the trace semantics of
$X\rightarrow\mathcal{ P}(\{*\}+\mathit{Act}\times X)$ identifies states that accept the same
language. With $B=\mathcal{D}$, the trace semantics of
$X\rightarrow\mathcal{D}(\{*\}+\mathit{Act}\times X)$ identifies states that accept the same
(finite) traces with the same probabilities.
\end{itemize}
The work of Jacobs et al is build on several assumptions, which limit the
generality of the definition of trace semantics. For instance, it is not
possible to define the trace semantics of finitely branching transition
systems.
\medskip\noindent\textbf{Results } In this paper, we reconsider the
definition of trace semantics in the category of algebras for the
branching type $B$. This allows us to includes the often occuring finite
non-determinism and finitely graded branching.
Moreover we propose a generic definition of coalgebraic logics characterising
states up to trace equivalence. Our definition of trace logics is build upon a
dual adjunction on the category of algebras for the branching type, and matches
the definition of coalgebraic modal logics for $T$-bisimulation.
\medskip\noindent\textbf{Structure of the paper } After reviewing
material known from the literature, Section~\ref{sec:EM} introduces
trace semantcis in the category of Eilenberg-Moore algebras of the
monad $B$ describing the branching type. Section~\ref{MT-logic}
describes trace logics using the adjunction induced by the closed
structure a the commutative monad $B$. Section~\ref{sec:pred-lift}
explains how to define logics via predicate lifting, a notion known
set-coalgebras, which is adapted to our
setting. Section~\ref{sec:gen-logic} introduces the notion of a
generic trace logic and uses it to prove a particular instance to be
sound, complete, and expressive.
\medskip\noindent\textbf{Acknowledgements} We would like to thank
Ichiro Hasuo and Bart Jacobs.
\section{Two Examples}\label{sec:examples}
\label{S:Examples}
Consider $\gamma:X\rightarrow\P_{\omega}(\{*\}+\mathit{Act}\times X)$. $(X,\gamma)$ is a
finitely non-deterministic automaton. Indeed, with $1$ as $\{*\}$ and
$+$ as (disjoint) union, we read $(a,x')\in\gamma(x)$ as $x$ can input
$a$ and go to $x'$ and we read $*\in\gamma(x)$ as $x$ is an accepting
state.
Now consider a logic
\begin{equation}\label{equ:exle1-syntax}
\phi ::= 0\mid \surd \mid \phi\vee\phi\mid \diam{a}\phi
\end{equation}
with compositional semantics
\begin{align}
x\not\Vdash 0 & \\
x\Vdash \surd & \Leftrightarrow \ *\in\gamma(x)\\
x\Vdash \phi\vee\psi \ & \Leftrightarrow \ x\Vdash\phi\ \textrm{or} \
x\Vdash\psi \\
x\Vdash\diam{a}\phi \ & \Leftrightarrow \ (a,x')\in\gamma(x)\ \textrm{and}
\ x'\Vdash\phi \label{equ:exle1-semantics-diam}
\end{align}
and as axiomatisation the usual laws for falsum (0) and disjunction
($\vee$) plus the axioms
\begin{equation}\label{equ:exle1-axioms}
\diam{a}0 = 0 \quad\quad \diam{a}(\phi\vee\psi) =
\diam{a}\phi\vee\diam{a}\psi
\end{equation}
Note that this implies the typical axiom we would expect for trace
logics
\begin{equation}
\label{Eq:Bifurcation}
\diam{a}(\diam{b}\phi\vee\diam{c}\psi) =
\diam{a}\diam{b}\phi\vee\diam{a}\diam{c}\psi
\end{equation}
Our development will not only provide a generic proof for the fact
that this logic is sound, complete and expressive, but also provide
conceptual explanations for why we can have falsum and disjunction,
but not negation and conjunction.
\medskip To see that the interaction of the modal operators $\diam{a}$
with the propositional operators (0,$\vee$) is subtle, consider as a
second example $\gamma:X\rightarrow\mathcal{D}(\{*\}+\mathit{Act}\times X)$ where $\mathcal{ D} Y$ is the
set of finitely supported discrete probability distributions on
$Y$. $\gamma(x,*)\in[0,1]$ is the probability of terminating successfully
and $\gamma(x,a,x')\in[0,1]$ is the probability of continuing with $a$
and transiting to $x'$. Two states $x,x'$ are trace equivalent if
(inventing an adhoc notation similar to the logic above)
\begin{equation}\label{equ:exle:prob}
x\Vdash p\cdot\diam{a_0}\ldots\diam{a_n}\surd \ \Leftrightarrow \
x'\Vdash p\cdot\diam{a_0}\ldots\diam{a_n}\surd
\end{equation}
which we read as stating that the probability of $x$ (and $x'$) to
terminate successfully after the sequence $a_0\ldots a_n$ is $p$.
The notation in (\ref{equ:exle:prob}) indicates that there must be a
definition of logic, semantics, axiomatisation paralleling the example
of non-determinstic automata and we will show how to obtain in a
systematic fashion from the functors involved.
\section{Preliminaries}
\label{sec:preliminaries}
\subsection{Monads, Algebras and Coalgebras}
\begin{definition
A \emph{coalgebra} for an endofunctor $T$ on a category $\mathcal{C}$ is a
morphism $\gamma:X\rightarrow TX$ for an object $X$ of $\mathcal{C}$, that we call
$\gamma$'s domain. A $T$-coalgebra morphism between coalgebras
$\gamma:X\rightarrow TX$ and $\delta:Y\rightarrow TY$ is a morphism $f:X\rightarrow Y$ such
that $Tf\circ \gamma=\delta\circ f$ commutes. Dually, a $T$-algebra
is an arrow $\alpha:TX\rightarrow X$.
\end{definition}
\begin{definition
A \emph{monad} on $\mathit{Set}$ is an endofunctor $B:\mathit{Set}\rightarrow\mathit{Set}$ with
natural transformations $\eta:\mathit{Id}\Rightarrow B$ and $\mu:BB\Rightarrow B$ such that
$\mu\circ\eta_T=\mathit{id}_T=\mu\circ T\eta$ and $\mu\circ\mu_T=\mu\circ
T\mu$. If $B$ preserves filtered colimits, the monads is called
\emph{finitary}.
\end{definition}
\begin{example}[finitary monads]
\label{Ex:Monads}
\begin{enumerate}
\item The finite powerset $\P_{\omega}$, equipped with the singleton map
$\{(-)\}$ and set-union.
\item The bag functor $\mathcal{B}$ takes a set $X$ to the set
$(\mathbb{N}^X)_{\omega}$ of its finite multisets, and functions $f:X\rightarrow
Y$ to multiset-functions $\mathcal{B} f:\mathcal{B} X\rightarrow\mathcal{B} Y$ taking multisets
$m\in(\mathbb{N}^X)_{\omega}$ to $\lambda y.\sum_{x\in
f^{-1}(y)}m(x)$.
\item A (sub-)distribution of a set $X$ is a function $d:X\rightarrow[0,1]$
such that $\sum_{x\in X}d(x)=1$ ($\sum_{x\in X}d(x)\leq 1$). The
(sub-)distribution functor $\D_{=1}$ ($\D_{\leq 1}$) takes a set $X$ to the
set of its (sub-)distributions, and functions $f:X\rightarrow Y$ to
$\lambda m.\lambda y.\sum_{x\in f^{-1}(y)}m(x)$. For the sake of a
brevity we write both, $\D_{=1}$ and $\D_{\leq 1}$, as $\mathcal{D}$ when it is
clear from context, which functor we mean.
For each $X$ we can define functions
\[
\begin{array}{lll}
\mu_X(d'\in\mathcal{D}^2 X)(x):=\sum_{d\in\mathcal{D} X}d'(d)\cdot d(x) & ~~ &
\eta_X(x):=\lambda y.\left\{\begin{array}{ll}1&\mbox{ if }y=x\\0&\mbox{ otherwise}\end{array}\right.
\end{array}
\]
$\mu$ and $\eta$ are transformations natural in $X$ and form with
$B$ a monad.
\item All of the above are examples of functors which take a set $X$ into
the set $\srngsom{X}$ of evaluations of $X$ into a semiring $\mathcal{S}$
with finite support, and functions $f:X\rightarrow Y$ into functions
$\srngsom{X}\rightarrow\srngsom{Y}$ such that $m\in\srngsom{X}\mapsto\lambda
y.\sum_{x\in f^{-1}(x)}m(x)$. For $\P_{\omega}$ the semiring is the boolean
algebra $\struct{\{\top,\bot\},\wedge,\vee,\top,\bot}$, and for $\mathcal{B}$
the semiring are the natural numbers $\struct{\mathbb{N},+,*,0,1}$.
\item If we take for $\mathcal{S}$ the real numbers with addition and
multiplication, then the category of algebras for the semiring monad
is (isomorphic to) the category of vector-spaces. See
Semadeni~\cite{semadeni:monads} for more on this perspective. More
generally, if the semiring does not happen to be a field, the
category of algebras for the monad is known as the category of
modules for the semiring.
\item Another example of a semiring monad uses the min-semiring
$\struct{\mathbb{N}\cup\{\infty\},\min,+,\infty,0}$ of natural numbers
augmented with a top element, $\infty$, with an idempotent additive
operation, $\min$, and a commutative multiplicative operation, $+$,
such that $\infty$ is neutral wrt $\min$ and $0$ wrt $+$, and $0$
absorbs wrt $\min$.
\item Another example of semiring monads can be found in the weighted
automata of Rutten~\cite{Rutten03Weighted}, where the stream
behaviour is an instance of the finite trace semantics presented in
this paper.
\end{enumerate}
\end{example}
An (Eilenberg-Moore-) algebra for a monad $B$ is an algebra for the
functor $B$ satisfying additionally $\alpha\circ\mu_X=\alpha\circ
B\alpha$ and $\alpha\circ\eta_X=id_X$.
The algebras for a monad $B$ form a category, the Eilenberg-Moore
category $B\mbox{-}\mathit{Alg}$. $U:B\mbox{-}\mathit{Alg}\rightarrow\mathcal{C}$ maps an algebra to its carrier. $U$
has a left adjoint $F$ and we write $\eta:\mathit{Id}\rightarrow UF$ and
$\varepsilon:FU\rightarrow\mathit{Id}$ for the unit and counit of the adjunction. Recall that
$UF=B$ and $F\varepsilon_{UX}=\mu_X$.
Each monad admits and initial and a final $B$-algebra, respectively
$\struct{B\emptyset,\mu_{\emptyset}B^2\emptyset\rightarrow B\emptyset}$ and
$\struct{\{*\},(\lambda _.*):B\{*\}\rightarrow\{*\}}$. Synonymously, we denote
by $1$ a singleton set, when the domain ($\mathit{Set}$ or $B\mbox{-}\mathit{Alg}$) is clear
from context.
For our definition of generic trace logics, it may be useful when
$B\mbox{-}\mathit{Alg}$ is \emph{closed} in the sense that homsets in $B\mbox{-}\mathit{Alg}$ have
$B$-algebra structure themselves. Kock~\cite{Kock70MonadsOnSMC}
showed that this is true for commutative monads.
\begin{definition}[Strength Laws]
A strength law for a monad $B$ is a transformation
$\mathit{st}_{X,Y}:=BX\times Y\rightarrow B(X\times Y)$ natural in $X$ and $Y$ and
commutes with the monad's unit and multiplication law such that
$\mathit{st}_{X,Y}\circ(\eta_X\times\mathit{id}_Y)=\eta_{X\times Y}$ and
$\mu_{X\times Y}\circ B\mathit{st}_{X,Y}\circ\mathit{st}_{BX,Y}=\mathit{st}_{X\times
Y}\circ(\mu_X\times\mathit{id}_Y)$.
A double strength law is a natural transformation given as the
diagonal $\mathit{dst}_{X,Y}:BX\times BY\rightarrow B(X\times Y)$ of $\mu_{X\times
Y}\circ B\mathit{st}_{Y,X}\circ\mathit{st}_{X,BY}=\mu_{X\times Y}\circ
B\mathit{st}_{X,Y}\circ\mathit{st}_{Y,BX}$, given it exists consistently.
A monad is commutative if it has a double strength law.
\end{definition}
\noindent The proof of the following can be found in \cite{Kock70MonadsOnSMC}.
\begin{proposition}
The Eilenberg-Moore category of a commutative monad is closed.
\end{proposition}
\subsection{The Kleisli Construction and Functor Liftings}
\begin{definition}[Kleisli-Categories]
\label{D:KleisliCats}
The Kleisli-category $\Kl B $ of a monad $B$ on $\mathcal{C}$ has as objects
the objects of $\mathcal{C}$ and arrows $f:X\rightarrow Y$ are the arrows $f:X\rightarrow BY$
in $\mathcal{C}$. The identity is given by $\eta:X\rightarrow BX$ and composition of
$f:X\rightarrow Y$ and $g:Y\rightarrow Z$ in $\Kl B $ is given by $g\circ
f:=\mu_Z\circ Bg\circ f$.
The adjunction $F'\dashv U':\mathcal{C}\rightarrow\Kl B $ is defined such that for
all sets $X$, $F'X:=X$, all functions $f:X\rightarrow Y$ in $\mathit{Set}$,
$F'f:=\eta_Y\circ f$, and for all objects $X$ in $\Kl B $, $U'X:=BX$
and for all morphisms $f:X\rightarrow Y$, $U'f:=\mu_Y\circ Bf$.
\end{definition}
\begin{example}
\begin{enumerate}
\item The Kleisli-category for the powerset monad $\P$ is $\mathit{Rel}$,
the category of sets as objects and relations as morphisms.
\item The Kleisli-category for the semiring monad $\srngsom{(-)}$ is the category of free (left) modules for the semiring $\mathcal{S}$.
\end{enumerate}
\end{example}
A coalgebra $\gamma:X\rightarrow BTX$ in $\mathit{Set}$ is a morphisms $X\rightarrow TX$ in
$\Kl B $. In order to exhibit $\gamma$ as a coalgebra in $\Kl B $ and
to have coalgebra morphisms, one defines the lifting of
$\mathit{Set}$-functors $T$ to $\Kl B $. The lifted functor $\overline{\T}$ makes
$FT=\overline{\T} F$ commute. The existence of the functor lifting is
equivalent to the existence of a distributive law.
\begin{definition}[Distributive Laws]
A distributive law for a monad $B$ and a functor $T$ is a natural
transformation $\pi:TB\Rightarrow BT$ such that $\pi\circ T\eta=\eta_T$ and
$\pi\circ T\mu=\mu_T\circ B\pi\circ\pi_B$ commute.
\end{definition}
\begin{example}
\label{Ex:DistrLaws}
Let $T(-):=\{*\}+\mathit{Act}\times (-)$ be a $\mathit{Set}$-functor for a fixed set
$\mathit{Act}$. With each of the monads in Example~\ref{Ex:Monads} $T$ has a
distributive law.
\begin{enumerate}
\item $\pi:T\P\rightarrow \P T$: $\pi_X(*):=\{*\}$, $\pi_X(a,Y\subseteq X):=\{(a,x)\mid x\in Y\}$.
\item $\pi:T\mathcal{B}\rightarrow\mathcal{B} T$: $\pi_X(*):=\eta_{\{*\}+\mathit{Act}\times X}(*)$, and
$\pi_X(a,m)(a,x):=\{(a,x)\mapsto m(x),(b,x)\mapsto 0,*\mapsto 0\mid a\in\mathit{Act},b\in\mathit{Act},b\neq a,x\in X\}$
\item $\pi:T\mathcal{D}\rightarrow\mathcal{D} T$: $\pi_X(*):=\eta_{\{*\}+\mathit{Act}\times X}(*)$, and
$\pi_X(a,d):=\{(a,x)\mapsto d(x),(b,x)\mapsto 0,*\mapsto 0\mid a\in\mathit{Act},b\in\mathit{Act},b\neq a,x\in X\}$ where $\mathcal{D}\in\{\D_{\leq 1},\D_{=1}\}$
\end{enumerate}
\end{example}
\begin{definition}[Functor Lifting by Distributive Law]
Given a distributive law $\pi:TB\rightarrow BT$ we can define $\overline{\T}$ on
objects $\overline{\T} X:=T X$ and on morphisms $\overline{\T}(f:X\rightarrow
Y):=\pi_Y\circT f$
\end{definition}
There is a full and faithful functor $K:\Kl B \toB\mbox{-}\mathit{Alg}$ mapping $X$ to
the free algebra over $X$, see~\cite{MacLane98Categories}. In other
words, we can think of $\Kl B $ as the full subcategory of $B\mbox{-}\mathit{Alg}$
consisiting of the free algebras.
\section{Coalgebraic Logic for Trace Semantics}
\label{S:TraceLog}
In this section we show how to set up trace logics in a coalgebraic
framework. But first we review some basic of coalgebraic logic (more
can be found in \cite{kurz:sigact06}) and the fundamentals of generic
trace semantics \cite{Jacobs04TraceSemantics}.
\subsection{A Brief Review of Logics for $T$-Bisimilarity }
\label{S:CoalgLogic}
Suppose we are looking for a logic for $T$-coalgebras built upon
classical propositional logic. Such a logic would be based on Boolean
algebras which precisely capture the axioms of propositional logic.
Then, in the same way as $T$ is a functor $\mathit{Set}\rightarrow\mathit{Set}$ on the models
(coalgebras) side, the logic will contain modalities given in terms of
a functor $L:\mathit{BA}\rightarrow\mathit{BA}$ on the category $\mathit{BA}$ of Boolean algebra.
The situation is depicted in
\begin{equation}\label{E:Duality}
\xymatrix{
\mathit{Set}\ar@/^.5pc/[rr]^{Q}\POS!R(-.7),\ar@(ul,dl)_{T} & \bot
& \op\mathit{BA}\ar@/^.5pc/[ll]^{S}\POS!R(.7),\ar@(ur,dr)^{L}
}
\end{equation}
$Q$ contravariantly takes sets $X$ to their powersets $2^X$ and
$S$ maps a Boolean algebra to the set of
maximal consistent theories (ultrafilters). For example, if $T=\mathcal{ P}$
we may define $L$ by saying that $LA$ is the Boolean algebra generated
by $\Diamond\phi, \phi\in A$, modulo the axioms
\begin{equation}\label{equ:K-axioms}
\Diamond 0=0 \quad\quad
\Diamond(\phi\vee\psi)=\Diamond\phi \vee \Diamond\psi
\end{equation}
Note how this definition of $L$ captures the usual modal logic for
(unlabelled) transition systems. The semantics of the logic is given
by a map
\begin{equation}\label{equ:delta}
\delta_X:LQX\rightarrow QTX
\end{equation}
In the example we define $\delta_X(\Diamond\phi)=\{\psi\in \mathcal{ P} TX
\mid \phi\cap\psi\not=\emptyset\}$ in order to capture that
$\Diamond\phi$ holds if the set `of successors' $\psi$ satisfies
$\phi\cap\psi\not=\emptyset$. Finally, $(L,\delta)$ gives rise to a
logic in the usual sense as follows. The set of formulas of the logic
is the carrier of the initial $L$-algebra. The semantics of a formula
wrt to a coalgebra $X\rightarrow TX$ is given by the unique homomorphism from
the initial $L$-algebra $LI\rightarrow I$ as in:
\begin{equation}\label{equ:semantics}
\xymatrix{
LI\ar[d]_{L(\sem{\cdot})}\ar[rr] && I\ar[d]^{\sem{\cdot}}\\
LQX\ar[r]^{\delta_X} & QTX \ar[r]^{Q\gamma} & QX
}
\end{equation}
\begin{theorem}
Any $(L,\delta)$ with $\delta$ as in (\ref{equ:delta}) gives rise to
a logic for $T$-coalgebras. The semantics $\sem{\cdot}$ as in
(\ref{equ:semantics}) is invariant under $T$-bisimilarity. The logic
is expressive for (finite) coalgebras, if $\delta_X$ is onto for
(finite) $X$ and the equational logic given by the axioms defining
$L$ is complete if $\delta_X$ is injective for all $X$.
\end{theorem}
\noindent Suppose we are given $T$, how can we find a logic
$(L,\delta)$? Two answers:
\begin{remark}\label{rem:generic-logics}
\begin{enumerate}
\item Moss~\cite{moss:cl} takes $LA$ to be the free $\mathit{BA}$ generated by
$TU\!A$ where $U\!A$ is the underlying set of $A$. A complete
calculus has been given in \cite{kkv:aiml08}.
\item The standard modal logic for $T=\mathcal{ P}$ above arises from
$LA=QTSA$ on finite $A$ and extending continuously to all of $\mathit{BA}$
\cite{kurz-rosi:calco07}. It is always complete.
\end{enumerate}
\end{remark}
Both logics are expressive. A detailed comparison has been given in
\cite{KurzLeal09Equational}.
\subsection{A Brief Review of Finite Trace Semantics}
\label{S:FinTraceSem}
\textbf{The basic construction } Consider a coalgebra $X\rightarrow BTX$, the
running example being $B=\mathcal{ P}$ and $TX=\{*\}+\mathit{Act}\times X$ as
discussed in Section~\ref{sec:examples}. The set of traces will be
the carrier of the initial $T$-algebra given by the colimit (or union) of the
sequence
\begin{equation}\label{equ:traces}
\xymatrix{
\emptyset\ar@{^(->}[rr]^{\emptyset} &&
T\emptyset\ar@{^(->}[rr]^{T\emptyset} &&
T^2\emptyset\ar@{^(->}[r] &
\cdots & T^\omega\emptyset}
\end{equation}
In the example $T^n\emptyset=\{a_1\ldots a_n\mid a_i\in\mathit{Act}\}$ and
$T^\omega\emptyset=\mathit{Act}^*$, ie the set of finite words over $\mathit{Act}$.
The set of traces of length $n$ will be given by a map
\begin{equation}\label{equ:trace-map}
\mathit{tr}_n:X\rightarrow BT^n\emptyset
\end{equation} In the example, $\mathit{tr}_n(x)$ is the set of traces of
length $n$ that lead from $x$ to an accepting state. To compute it, we
need the following ingredients.
\begin{assumption}\label{ass:B}\
\begin{itemize}
\item a map $\mu_X:BBX\rightarrow BX$ (for this we assume that $B$ is a monad)
\item a map $\pi_X:TBX\rightarrow BTX$ (for this we assume that $\pi$ is a distributive
law)
\item an algebra morphism $e:A\rightarrow F\emptyset$ from any $B$-algebra $A$
into $F\emptyset$.\footnote{This means that we assume from hereon
$B\emptyset\not=\emptyset$. Also note that in all our examples $B$
is a commutative monad, hence $B\emptyset\not=\emptyset$ implies
$B\emptyset=1$, so that $F\emptyset$ is the final algebra.}
\end{itemize}
\end{assumption}
\noindent The maps $\mathit{tr}_n$ then arise from taking $n$ steps of
$\gamma$, eg in the case $n=2$, as
\begin{equation*}
\xymatrix{
X \ar[r]^{\gamma\ \ } & BTX \ar[r]^{BT\gamma\ \ } & BTBTX\ar[r]^{BTBTe} &
BTBTB\emptyset\ar[r]^{p} & BBBTT\ar[r]^{m}\emptyset & BT^2\emptyset}
\end{equation*}
($p$ stands for 3 applications of $\pi$ and $m$ for 2 applications of
$\mu$.)
\begin{definition}\label{def:tr-sem}
Two states $x,y\in X$ of a coalgebra $X\rightarrow BTX$ are trace equivalent
if $\mathit{tr}_n(x)=\mathit{tr}_n(y)$ for all $n<\omega$.
\end{definition}
For the purposed of the current paper, we consider this the essence of
the trace semantics of
\cite{HasuoJacobsSokolova06GenericTraceTheory}. But
\cite{HasuoJacobsSokolova06GenericTraceTheory} do much more and, in
particular, they show that under additional assumptions the trace
semantics can be given by a final coalgebra in the Kleisli category.
\medskip\noindent\textbf{Trace semantics in the Kleisli category }
\cite{HasuoJacobsSokolova06GenericTraceTheory} show not only that the
ingredients of a monad $B$ and a distributive law $TB\rightarrow BT$ give rise
to trace semantics, they also show that it can be elegantly formulated
in the so-called Kleisli category of the monad $B$ (see
Section~\ref{sec:preliminaries}). The objects in the Kleisli category
are the same as in $\mathit{Set}$, but arrows $X\rightarrow Y$ in $\Kl B $ are maps
$X\rightarrow BY$ in $\mathit{Set}$. In case of the powerset functor $B=\mathcal{ P}$,
$\Kl B $ is the category of sets with relations as arrows.
The distributive law $TB\rightarrow BT$ gives rise to a lifting of
$T:\mathit{Set}\rightarrow\mathit{Set}$ to $\overline{\T}:\Kl B \rightarrow\Kl B $. The definition of $\mathit{tr}_n$
can then be defined inductively as
\begin{equation}\label{equ:def-tr}
\mathit{tr}_{n+1}=\overline{\T}(\mathit{tr}_n)\circ\gamma
\end{equation}
where we assume a morphism $\mathit{tr}_0:X\rightarrow 0$ in the base case. The
following diagram illustrates the above definition
\begin{equation}
\label{equ:traces-in-Kl}
\xymatrix{
&
X\ar[r]^{\gamma}\ar@/_.5pc/[ld]_{\mathit{tr}_0}\ar@/^1pc/[rd]_{\mathit{tr}_n}\ar@/^1pc/[rrd]_{\mathit{tr}_{n+1}}&
\overline{\T} X\ar@/^1pc/[rd]^{\overline{\T}\mathit{tr}_n} &&\\
\emptyset &
\cdots &
\overline{\T}^n\emptyset &
\overline{\T}^{n+1}\emptyset &
\cdots\\
}
\end{equation}
\hide{
in $\Kl B $
\begin{equation}\label{equ:traces-in-Kl}
\xymatrix{
\emptyset\ar[r]^{\emptyset} &
\overline{\T}\emptyset\ar[r]^{T\emptyset} &
\overline{\T}\olT\emptyset &
\cdots & \overline{\T}\olT^n\emptyset & \cdots \\
X \ar[r]^{\gamma} \ar[u]^{\mathit{tr}_0} & \overline{\T} X \ar[r]^{\overline{\T} \mathit{tr}_0}
\ar[u]^{\overline{\T} \mathit{tr}_0} & \overline{\T}\olT X \ar[u]^{\overline{\T}\olT\mathit{tr}_0}&
\cdots & \overline{\T}^n X
\ar[u]^{} & \cdots
}
\end{equation}
}
Furthermore, under conditions for which we refer to
\cite{HasuoJacobsSokolova06GenericTraceTheory}, the final
$\overline{\T}$-coalgebra $Z$ exists.\footnote{Moreover,
\cite{HasuoJacobsSokolova06GenericTraceTheory} prove the beautiful
result that show that the final $\overline{\T}$-coalgebra is given by the
initial $T$-algebra with the carrier $T^\omega\emptyset$ as in
(\ref{equ:traces}).} Therefore, with the notation of
Definition~\ref{def:tr-sem}, there is a map $\mathit{tr}:X\rightarrow BZ$ with the
property $$\mathit{tr}(x)=\mathit{tr}(y) \Leftrightarrow \mathit{tr}_n(x)=\mathit{tr}_n(y)$$ for all
$n<\omega$. Thus, the trace semantics via the final coalgebra (if it
exists) in the Kleisli-category is equivalent to the one of
Definition~\ref{def:tr-sem}. The advantage of the trace semantics via
the final coalgebra in the Kleisli-category is that it gives a
coinductive account of trace semantics. The disadvantage is that it
excludes some natural examples such as \emph{finite} powersets or
multisets. The next section shows that these examples can be treated
via final coalgebras if we move from the Kleisli-category to the
category of algebras for the monad.
\subsection{Trace Semantics in the Eilenberg-Moore
Category}\label{sec:EM}
In this section we propose to move the trace semantics from the
Kleisli-category $\Kl B $ to the category $B\mbox{-}\mathit{Alg}$ of
Eilenberg-Moore-algebras. There are at least two reasons why this of
interest. The first is that the duality we will exploit for the logic
takes place in $B\mbox{-}\mathit{Alg}$. The second is that, in general, the limit of
Diagram~(\ref{equ:traces-in-Kl}) is not a free $B$-algebra and hence
not in $\Kl B $, but it always exists in $B\mbox{-}\mathit{Alg}$.
Let $K$ denote the functor which embeds $\Kl B$ into $B\mbox{-}\mathit{Alg}$. Our
first task is to extend $\overline{\T}:\Kl B\rightarrow\Kl B$ to $\widetilde{\T}:B\mbox{-}\mathit{Alg}\toB\mbox{-}\mathit{Alg}$
so that $\widetilde{\T} K\cong K\ol T$ (hence $\widetilde{\T} F\cong F T$).
\begin{equation}
\xymatrix{
\Klom B\ar@/_1pc/@{^(->}[rr]_{J}\ar@(ul,ur)^{\overline{\T}'} &
\Kl B\ar@/^.5pc/@{^(->}[r]^{K}\ar@(ul,ur)^{\overline{\T}} &
B\mbox{-}\mathit{Alg}\ar@(ul,ur)^{\widetilde{\T}}
}
\end{equation}
On the full subcategory of free algebras we can define $\widetilde{\T} FX=K\overline{\T}
X=FTX$. To extend this to arbitrary algebras $A$ recall first that any
$A\inB\mbox{-}\mathit{Alg}$ is a coequaliser of $FU\varepsilon_A,\varepsilon_{FUA}:FUFUA\rightarrow FUA$. We
then define $\widetilde{\T} A$ as the coequaliser of $\widetilde{\T} FU\varepsilon_A$ and $\widetilde{\T}
\varepsilon_{FUA}$. It can be shown that $\widetilde{\T}$ is the left Kan-extension of
$K\overline{\T}$ along $K$.
\begin{example}\label{exle:wtT}
Let $B=\P_{\omega}$ and $T=\{\surd\}+\mathit{Act}\times\mathit{Id}$. Then $\widetilde{\T} A\cong
F1+\mathit{Act}\cdot A$. Indeed, by definition, we have $\widetilde{\T} FX= FTX \cong
F1+\mathit{Act}\cdot FX$. Now the claim follows from the fact that the
functor $F1+\mathit{Act}\cdot\mathit{Id}$, being a coproduct, preserves
coequalisers.
\end{example}
It is convenient for us to make the following assumptions.
\begin{assumption}\label{ass:fin}
$\mathcal{B}:\mathit{Set}\rightarrow\mathit{Set}$ is a finitary commutative monad with
$B\emptyset\not=\emptyset$ and $T:\mathit{Set}\rightarrow\mathit{Set}$ is a finitary functor
with a distributive law $TB\rightarrow BT$.
\end{assumption}
\begin{remark}\label{rmk:wtT-fb}
If $\mathcal{B}$ and $T$ are finitary, then $\widetilde{\T}$ is determined by finitely
generated free algebras, or, in other words, $\widetilde{\T}$ preserves sifted
(hence filtered) colimits \cite{arv} and falls within the framework
considered in \cite{kurz-rosi:strcompl,vele-kurz:fb}. For a functor
$H:\mathcal{ A}\rightarrow\mathcal{ A}$ on a finitary algebraic category $\mathcal{ A}$ to be
strongly finitary means that $H$ is determined by its action on
finitely generated free algebras. More formally, $H$ is a left
Kan-extension of $HK$ along $K$ where $K$ is the inclusion
$\mathcal{ A}_0\rightarrow\mathcal{ A}$ of the full subcategory $\mathcal{ A}_0$ of finitely
generated free algebras. A pleasant consequence is that all concrete
calculations of some $HA$ can be restricted to the case $A=Fn$,
where $F$ is the left adjoint of the forgetful functor
$\mathcal{ A}\rightarrow\mathit{Set}$ and $n$ is finite. This will be exploited in the
following for $\mathcal{ A}=B\mbox{-}\mathit{Alg}$. Other consequences of our assumption
then are:
\begin{itemize}
\item $F\emptyset$ is the initial and final object of $B\mbox{-}\mathit{Alg}$.
\item The final $\widetilde{\T}$ sequence converges after $\omega$ steps.
\end{itemize}
\end{remark}
In a second step, we can now map a coalgebra $\gamma:X\rightarrow BTX$ (ie
$\gamma:X\rightarrow\overline{\T} X$) to ${\widetilde\gamma}:FX\rightarrow \widetilde{\T} FX$ (ie ${\widetilde\gamma}:KX\rightarrow
K\overline{\T} X$). Thus ${\widetilde\gamma}$ is a coalgebra for a functor
$\widetilde{\T}:B\mbox{-}\mathit{Alg}\toB\mbox{-}\mathit{Alg}$. Moreover we observe that we can factor
$\mathit{tr}_n:X\rightarrow BT^n\emptyset$ from Diagram~(\ref{equ:traces-in-Kl}) as
$$ \mathit{tr}_n : X\rightarrow BX\cong UFX
\stackrel{U\widetilde\tr_n}{\longrightarrow}\widetilde{\T}^nF\emptyset\cong BT^n\emptyset$$ where
we define $\widetilde\tr_0$ via $e$ as in Assumption~\ref{ass:B} and
$\widetilde\tr_{n+1}=\widetilde{\T}\widetilde\tr_n\circ{\widetilde\gamma}$. Let us summarise this in a
definition and a proposition.
\begin{definition}
Recall Assumption~\ref{ass:fin}. For any coalgebra $\alpha:A\rightarrow\widetilde{\T}
A$ we define the trace semantics as follows. First, $\widetilde\tr:A\rightarrow
F\emptyset$ is given by finality; then, inductively
$\widetilde\tr_{n+1}=\widetilde{\T}\widetilde\tr_n\circ{\widetilde\gamma}$. This defines a cone on the
final $\widetilde{\T}$-sequence so we can define the trace semantics
$\widetilde\tr:A\rightarrow Z$, where $Z\rightarrow\widetilde{\T} Z$ is the final
$\widetilde{\T}$-coalgebra. For a coalgebra $\gamma:X\rightarrow BTX$ we define
$\mathit{tr}:X\rightarrow UZ$ as $U\widetilde\tr\circ\eta_X$, where $\widetilde\tr$ is the trace
semantics of ${\widetilde\gamma}:FX\rightarrow\widetilde{\T} FX$.
\end{definition}
To emphasise that this definition agrees with the one of the previous
subsection we state
\begin{proposition}
Consider $\gamma:X\rightarrow BTX$ and ${\widetilde\gamma}:FX\rightarrow FTX= \widetilde{\T}
FX$. Then $U\widetilde\tr_n\circ \eta_X=\mathit{tr}_n$.
\end{proposition}
\noindent Thus, $Z$ and $\widetilde\tr$ and $\mathit{tr}$ are just a convenient way to
talk about the maps $\mathit{tr}_n$ for all $n\in\mathbb{N}$
simultaneously. In particular, we have now again a coinductive account
of trace semantics. This technique will give, for example, a short and
conceptual proof of Theorem~\ref{thm:main}. Under
Assumption~\ref{ass:fin}, and if the final $\overline{\T}$-coalgebra of
\cite{HasuoJacobsSokolova06GenericTraceTheory} exists, then both the
trace semantics in $\Kl B $ and the trace semantics in $B\mbox{-}\mathit{Alg}$ are
equivalent as both boil down to Definition~\ref{def:tr-sem}. (Of
course, this is due to the fact that the definition of $\widetilde{\T}$ extends
to all algebras the lifting $\overline{\T}$ of $T$ to $Kl(B)$.)
\begin{remark}\label{rmk:fin-sem-BAlg} If $B\emptyset\not=0$ then the sequence
$(\widetilde{\T}^nF\emptyset)_{n<\omega}$ is the finitary part of the final
$\widetilde{\T}$-sequence in $B\mbox{-}\mathit{Alg}$. Moreover, it follows from
Remark~\ref{rmk:wtT-fb} that if $B$ is finitary, then the
$\omega$-limit $(\widetilde{\T}^\omega F\emptyset)$ of the final sequence is
the final $\widetilde{\T}$-coalgebra. To summarise, in addition to the
explanation of trace semantics as a final semantics in the
Kleisli-category as in
\cite{HasuoJacobsSokolova06GenericTraceTheory}, we can also give a
final semantics in the Eilenberg-Moore category. These two
approaches are slightly different, for example, the approach of
\cite{HasuoJacobsSokolova06GenericTraceTheory} works for $B=\P$ but
not for $B=\P_{\omega}$, whereas for us it is more natural to work with
$B=\P_{\omega}$ as we then have algebras with a finitary signature.
\end{remark}
\begin{example}\label{exle:B=Pom:2}
Consider $B=\P_{\omega}$, $T=\{\surd\}+\mathit{Act}\times\mathit{Id}$. Then
$\widetilde{\T}(FX)=F\{\surd\} + \mathit{Act}\cdot FX$. We can identify $F\emptyset$
with $\{\emptyset\}$ and $\widetilde{\T}^n(F\emptyset)$ with $\P_{\omega}(1+\mathit{Act}
+\ldots \mathit{Act}^n)$. Thus, elements of $\widetilde{\T}^n(F\emptyset)$ are finite
sets of finite words $\langle a_1\ldots a_i\rangle$, $i\le n$. As
$F\emptyset$ is initial and final in $B\mbox{-}\mathit{Alg}$, the
$\widetilde{\T}^n(F\emptyset)$ are part of the initial and of the final
$\widetilde{\T}$-sequence. The projections
$p^{n+1}_n:\widetilde{\T}^{n+1}(F\emptyset)\rightarrow\widetilde{\T}^{n}(F\emptyset)$ are
finite-union-preserving maps determined by acting as the identity on
singletons $\{\langle a_1\ldots a_i\rangle\}$ for $i\le n$ and
sending $\{\langle a_1\ldots a_{n+1}\rangle\}$ to $\emptyset$. The
embeddings $e^n_{n+1}:\widetilde{\T}^n(F\emptyset)\rightarrow\widetilde{\T}^{n+1}(F\emptyset)$
are given by the obvious inclusions. Note that $p^{n+1}_n\circ
e^n_{n+1}= \mathit{id}_n$. The colimit of the initial $\widetilde{\T}$-sequence
$(e^n_{n+1})_{n<\omega}$ is given by all finite subsets of
$\mathit{Act}^*=\coprod_{n<\omega}\mathit{Act}^n$. The limit of the final
$\widetilde{\T}$-sequence $(e^n_{n+1})_{n<\omega}$ is given by all subsets of
$\mathit{Act}^*$. Note that although all approximants $\widetilde{\T}^n(F\emptyset)$
are free algebras, the limit $\P(\mathit{Act}^*)$ is not free in $B\mbox{-}\mathit{Alg}$ and
hence does not appear in $\Kl(\P_{\omega})$.
\end{example}
\subsection{Logics for Finite $B$-Traces }
\label{MT-logic}
We develop logics for $(B,T)$-coalgebras with a semantic invariant
under trace equivalence in analogy to coalgebraic modal logic
for $T$-bisimulation.
Firstly we need a category carrying our logics. We have a number of
possible replacements for $\mathit{BA}$ in Diagram~(\ref{E:Duality}):
distributive lattices for positive logic, Heyting algebras for
intuitionistic logic, complete atomic Boolean algebras for infinitary
logic. The minimal choice (without propositional operators) is $\mathit{Set}$
itself as used for example by Klin in~\cite{klin:lics07}.
\begin{equation}\label{E:setsetop}
\xymatrix{
\mathit{Set}\ar@/^.5pc/[rr]^{2^{(-)}} &
\bot &
\mathit{Set}^{op}\ar@/^.5pc/[ll]^{2^{(-)}}
}
\end{equation}
In the above situation, $2$ takes the role of a schizophrenic
object. Analogously we may choose a $B$-algebra $\Omega$ to replace
$2$. In most examples we have considered, $F1$ is a suitable
choice, but for the moment we do not need to fix a choice.
\begin{notation}
If $B$ is a commutative monad, we write $Q$ for the contravariant
endofunctor $[-,\Omega]$ on $B\mbox{-}\mathit{Alg}$ where $\Omega$ is for now an
arbitrary but fixed object of `truth values'.
\begin{equation}\label{E:balgbalgop}
\xymatrix{
B\mbox{-}\mathit{Alg}\ar@/^.5pc/[rr]^{Q=[-,\Omega]}
&\bot&
\opB\mbox{-}\mathit{Alg}\ar@/^.5pc/[ll]^{Q=[-,\Omega]}
}
\end{equation}
$Q_0$ is the contravariant endofunctor
$U[F-,\Omega]=Set(-,U\Omega)$. We have $UQA=Q_0 UA$.
\end{notation}
\begin{example}\label{exle:one}
When $B=\P_{\omega}$, $B\mbox{-}\mathit{Alg}=\mathit{SLat}$ is the category of (join)
semi-lattices. For $\Omega$ we choose the two-element semi-lattice
$F1=\mathbbm{2}$, so that $[-,F1]$ takes a semi-lattice $A$ to the set of
`prime filters' over $A$. For future calculations, we record some
facts about semi-lattices. First, for finite $A$, there are
order-reversing bijections
\begin{equation}\label{exp-log}
\xymatrix{
A \ar@/^/[rr]^{\exp} & & [A,\mathbbm{2}] \ar@/^/[ll]^{\log}
}
\end{equation}
given by $\exp(a)=\lambda b. \neg(b\le a)$ and
$\log(\phi)=\bigvee\phi^-$ where $\neg:\mathbbm{2}\rightarrow\mathbbm{2}$ is negation
and $\phi^- = \{a\in A\mid \phi(a)=0\}$. Another description of $\log$
goes as follows. Since $\phi$ preserves joins it has a right adjoint
$\phi^\sharp$ and $\log(\phi)=\phi^\sharp(0)$. Second, if $A=FX$ with
$X$ not necessarily finite, we have the bijection
\begin{equation}\label{equ:UQFX}
UQFX=U[FX,\mathbbm{2}]\cong\mathit{Set}(X,2)\cong \P X
\end{equation}
which lifts to a semi-lattice isomorphism
\begin{equation}\label{equ:QFX}
QFX\cong (\P X, \emptyset, \cup)
\end{equation}
mapping $\phi\in[FX,\mathbbm{2}]$ to $\{x\in X\mid \phi(\{x\}) = 1\}$ and
$S\subseteq X$ to the unique $\phi$ with $\phi(x)=1\Leftrightarrow
x\in X$, or, equivalently, to $\lambda S'\in FX\,.\, S\cap
S'\not=\emptyset$ (where we use \eqref{equ:UQFX} to identify $S'$ with
a subset of $X$). Taking now $X=n$ finite again, we
obtain
\begin{equation}\label{setify0}
QFn\cong Fn.
\end{equation}
In this case it is more convenient to use $\exp$ and $\log$ to denote
the order-preserving bijections
\begin{equation}\label{exp-log}
\xymatrix{
Fn \ar@/^/[rr]^{\exp} & & [Fn,\mathbbm{2}] \ar@/^/[ll]^{\log}
}
\end{equation}
given by $\log(\phi)=\{i\in n\mid \phi(\{i\})=1\}$ and
$\exp(S)=\lambda S'. S\cap S'\not=\emptyset$ (where again we identify
elements $S, S'$ of $Fn$ with subsets $S,S'\subseteq n$).
One can check that $\exp(\exp(a)) = \lambda\phi.\phi(a)$. It follows
that $\exp\circ\exp: \mathit{Id}\rightarrow QQ$ is the unit of the adjunction
\eqref{E:balgbalgop}, and, moreover, that the unit is an isomorphism
on finite semi-lattices.\footnote{This also follows from the fact that
the adjunction \eqref{E:balgbalgop} restricts to an equivalence on
finite semi-lattices \cite{johnstone:stone-spaces}.} In case of
$Fn\rightarrow QQFn$ we have for $S\subseteq n$ that $\exp(\exp(S))(\phi) =
\log(\phi)\cap S\not=\emptyset$. The inverse $QQFn\rightarrow Fn$ of $Fn\rightarrow
QQFn$ maps $u:[Fn,\mathbbm{2}]\rightarrow\mathbbm{2}$ to $\log(\log(u))=n\setminus
\{i\in n\mid \exists\phi\,.\,u(\phi)=0 \,\&\, \phi(\{i\})=1\}$.
We will also use that for finite semi-lattices coproducts and products
coincide, with
\begin{equation}
\label{equ:co-prod}
\begin{array}{rcl}
A+B & \rightarrow & A\times B\\
a & \mapsto & (a,0)\\
b & \mapsto & (0,b)\\
a\vee b & \mapsfrom & (a,b)
\end{array}
\end{equation}
describing the isomorphism. \qed
\end{example}
In Section~\ref{S:FinTraceSem} we have defined the finite trace
semantics of $\mathit{Set}$-coalgebras $\gamma:X\rightarrow BTX$ as the
final coalgebra semantics of the lifted coalgebra $\gamma:FX\rightarrow\widetilde{\T} FX$
in $B\mbox{-}\mathit{Alg}$.
Secondly we need a functor $L$ providing the modalities for our logics,
as in the following diagram.
\begin{equation}
\label{E:balgbalgopTL}
\xymatrix{
B\mbox{-}\mathit{Alg}\ar@/^.5pc/[rr]^{Q}\POS!R(-.7)\ar@(ul,dl)_{\widetilde{\T}}
&\bot&
\opB\mbox{-}\mathit{Alg}\ar@/^.5pc/[ll]^{Q}\POS!R(.7),\ar@(ur,dr)^{L}
}
\end{equation}
In analogy to Section~\ref{S:CoalgLogic}, we develop finite trace
logics as the initial $L$-algebra $\L:LI\rightarrow I$ in $B\mbox{-}\mathit{Alg}$. Note that
under the assumptions of Remark~\ref{rmk:wtT-fb}, we have that $I$ is the
$\omega$-colimit of the initial $L$-sequence:
\begin{equation}
\xymatrix{
0\ar[r] & L0\ar[r] & L^20\ar[r] & \cdots
}
\end{equation}
\begin{definition}
A trace logic is given by a functor $L:B\mbox{-}\mathit{Alg}\toB\mbox{-}\mathit{Alg}$ and a natural
transformation $\delta: LQ\rightarrow Q\widetilde{\T}$. Formulas of the logic are
given by elements of the initial $L$-algebra. The semantics
$\sem{\cdot}_{\widetilde\gamma}$ wrt a $\widetilde{\T}$-coalgebra ${\widetilde\gamma}:FX\rightarrow\widetilde{\T} FX$
is given by initiality as in
\begin{equation}\label{equ:tr-semantics
\xymatrix{
LI\ar[d]_{L\sem{\cdot}_{\widetilde\gamma}}\ar[rr] && I\ar[d]^{\sem{\cdot}_{\widetilde\gamma}}\\
LQFX\ar[r]^{\delta_{FX}} & Q\widetilde{\T} FX \ar[r]^{Q{\widetilde\gamma}} & QFX
}
\end{equation}
This induces the semantics $\sem{\cdot}_\gamma$ wrt a coalgebra
$\gamma:X\rightarrow BTX$ via
\begin{equation}\label{equ:sem-set-coalg}
\xymatrix{
UI\ar[r]^{U\sem{\cdot}_{\widetilde\gamma}\ \ } & UQFX\ar[r]^\cong & Q_0 UFX \ar[r]^{Q_0\eta_X} & Q_0 X
}
\end{equation}
\end{definition}
For future reference, we record that the semantics in terms of
$\gamma$ and ${\widetilde\gamma}$ agree:
\begin{proposition}\label{prop:wtgamma}
Let ${\widetilde\gamma}:FX\rightarrow \widetilde{\T} FX$ be the $\widetilde{\T}$-coalgebra induced by the
$(B,T)$-coalgebra $\gamma:X\rightarrow BTX$, that is,
$\gamma=U{\widetilde\gamma}\circ\eta_X$ with $\eta_X:X\rightarrow BX$ the unit of the
monad $B$. Then $\sem{\phi}_\gamma(x)=
\sem{\phi}_{\widetilde\gamma}(\eta_X(x))$.
\end{proposition}
\begin{example}\label{exle:two}
Continuing from Example~\ref{exle:one}, in order to describe the
logic (\ref{equ:exle1-syntax}), we let $LA$ be the join-semilattice
which is freely generated by $\surd$ and $\diam{a}\phi$ for
$a\in\mathit{Act}$ and $\phi\in A$, quotienting by
(\ref{equ:exle1-axioms}). To describe $\delta_{FX}$ it is convenient
to note that $QFX$ can be identified with the set of subsets of $X$
as in \eqref{equ:QFX}
and $Q\widetilde{\T} FX= QFTX$ with the set of subsets of $TX$. It therefore
makes sense to define
\begin{align*}
\delta_{FX}:LQFX & \rightarrow Q\widetilde{\T} FX \label{equ:tr-delta-exle1}\\
\surd & \mapsto \{S\subseteq TX \mid *\in S\}\\
\diam{a}\phi &\mapsto \{S \subseteq TX \mid \exists x(x\in\phi\
\&\ (a,x)\in S)\}
\end{align*}
\end{example}
\begin{proposition}\label{prop:exle1}
$(L,\delta)$ of Example~\ref{exle:two}, together with
(\ref{equ:tr-semantics}), describes the same logic as (\ref{equ:exle1-syntax}) in
Section~\ref{sec:examples}.
\end{proposition}
\begin{proof}
For example, we calculate
$x\models\diam{a}\phi \ \Leftrightarrow \ %
\gamma(x)\in \{S \subseteq TX \mid \exists x'(x'\in\phi\ \&\ (a,x')\in
S)\}\Leftrightarrow \ %
\gamma(x)\in \delta_{FX}(\diam{a}\phi)\Leftrightarrow \ %
x\in QF\gamma(\delta_{FX}(\diam{a}\phi))\ \Leftrightarrow \ %
x\in\sem{\diam{a}\phi}$ where we use, respectively,
(\ref{equ:exle1-semantics-diam}), the definition of $\delta$, the
definition of $Q$, and (\ref{equ:tr-semantics}).
\end{proof}
\begin{theorem}\label{thm:main}
Consider a functor $T:\mathit{Set}\rightarrow\mathit{Set}$, a monad $B$, and a distributive
law $TB\rightarrow BT$. Any $(L,\delta)$ with $L:B\mbox{-}\mathit{Alg}\toB\mbox{-}\mathit{Alg}$ and
$\delta_K:LQK\rightarrow QK\overline{\T}$ gives rise to a logic for $BT$-coalgebras
invariant under $B$-trace semantics.
\end{theorem}
\begin{proof}
For a given $\gamma:X\rightarrow BTX$ and formula $\phi$, we have to show
that $\mathit{tr}(x)=\mathit{tr}(y)$ implies $x\Vdash\phi \Leftrightarrow
y\Vdash\phi$. Expressing this in $B\mbox{-}\mathit{Alg}$, this amounts to
$\widetilde\tr(\eta_X(x))=\widetilde\tr(\eta_X(y))$ only if $x\in\sem{\phi}_{\widetilde\gamma}
\Leftrightarrow y\in\sem{\phi}_{\widetilde\gamma}$. But this is immediate from
the initiality of the algebra of formulas as follows. Let
$(Z,\zeta)$ be the final $\widetilde{\T}$-coalgebra.
\begin{equation
\xymatrix{
LI\ar[d]_{L\sem{\cdot}_\zeta}\ar[rr] && I\ar[d]^{\sem{\cdot}_\zeta}\\
LQZ\ar[r]^{\delta_{Z}}\ar[d]_{LQ\widetilde\tr}
& Q\widetilde{\T} Z\ar[r]^{Q\zeta}\ar[d]_{Q\widetilde{\T}\widetilde\tr} & QZ\ar[d]_{Q\widetilde\tr}\\
LQFX\ar[r]^{\delta_{FX}} & Q\widetilde{\T} FX \ar[r]^{Q{\widetilde\gamma}} & QFX\\
}
\end{equation}
Since morphisms from the initial algebra $LI\rightarrow I$ are uniquely
determined, we must have
$\sem{\cdot}_{\widetilde\gamma}=Q\widetilde\tr\circ\sem{\cdot}_\zeta$.
\end{proof}
\hide{
\begin{remark}
As in \cite{KupkeKurzPattinson04RLC} the logic will be complete if
$\delta$ is in injective and expressive if $\delta$ is surjective
and $L$ respectively preserves these properties.
\end{remark}
\begin{remark}
there should be a remark analogous to Remark 1
\end{remark}
}
\subsection{Predicate Liftings}\label{sec:pred-lift}
\newcommand{\fp}{\mathsf{fp}} Whereas the previous section treats
logics from an abstract point of view, we are now going to see how to
describe them concretely using predicate liftings. First, we need to
extend the set-based notion of predicate lifting
\cite{pattinson:cml-j,schroeder:fossacs05} to coalgebras over $B\mbox{-}\mathit{Alg}$.
Suppose we have $L$ and $$LQ\rightarrow Q\wt T.$$ Using $\mathit{Id}\rightarrow QQ$ from the
adjunction \eqref{E:balgbalgop} this gives us $$L\rightarrow LQQ\rightarrow Q\wt T
Q.$$ We will see below that $Q\wt T Q$ gives us predicate liftings,
but first we are going to show how to recover $LQ\rightarrow Q\wt T$ from
$L\rightarrow Q\wt T Q$. Write $$J:\Kl_\omega B\rightarrow B\mbox{-}\mathit{Alg}$$ for the inclusion of
the category of finitely generated free algebras into $B\mbox{-}\mathit{Alg}$.
\begin{proposition}\label{prop:pred-lift}
Let $L$ be determined by finitely generated free algebras as in
Remark~\ref{rmk:wtT-fb}. Then there is a bijection between natural
transformations $LQ\rightarrow Q\widetilde{\T}$ and natural transformations $LJ\rightarrow
Q\widetilde{\T} QJ$.
\end{proposition}
\begin{proof}
Given $\delta:LQ\rightarrow Q\wt T$ we obtain $\rho:LJ\rightarrow Q\wt T QJ$ as
$\delta Q\circ L\eta$. Conversely, given $\rho$, we write $QA$ as a
colimit $\phi_i:Fn_i\rightarrow QA$, which is preserved by $L$, and obtain
$\delta$ via
\begin{equation}\label{equ:deltarho}
\xymatrix{
QA & L QA \ar[r]^{\delta_A} & Q\widetilde{\T} A\\
Fn_i\ar[u]^{\phi_i} & L Fn_i \ar[u]^{L \phi_i}\ar[r]^{\rho_{Fn_i}} & Q\widetilde{\T} QFn_i
\ar[u]_{Q\widetilde{\T}\check \phi_i} }
\end{equation}
where $\check \phi_i:A\rightarrow QFn_i$ is the adjoint transpose of $\phi_i$. To
check that these two assignments are inverse to each other, we first
note that the diagram \eqref{equ:deltarho} can be rewritten as
\begin{equation}
\xymatrix{
& L QA \ar[r]^{\delta_A} & Q\widetilde{\T} A\\
L Fn_i \ar[ur]^{L \phi_i} \ar[r]^{L\eta} &
L QQFn_i \ar[u]_{L Q \check \phi_i} \ar[r]_{\delta QFn_i} &
Q\widetilde{\T} QFn_i \ar[u]_{Q\widetilde{\T}\check \phi_i} }
\end{equation}
where the triangle commutes because of the adjunction
\eqref{E:balgbalgop} and the quadrangle commutes because of
naturality. It follows that starting from $\delta$ and defining
$\rho$, the original $\delta$ satisfies \eqref{equ:deltarho} and
therefore agrees with the $\delta$ defined from $\rho$. Conversely,
defining $\delta$ from $\rho$ in \eqref{equ:deltarho}, one can
choose $A=QFn$, $n_i=n$ and $\check\phi=\mathit{id}$, which shows that $\delta$
determines the $\rho$ it comes from uniquely.
\end{proof}
We can interpret the proposition as follows. An element of $$Q_0
A=UQA$$ is a predicate on $A$. An element of $$[n,Q_0 A]$$ is an
$n$-ary predicate on $A$. We have $[n,Q_0
A]\cong[Fn,QA]\cong[A,QFn]$ and find it useful to introduce the
following notation. We want to write $\phi$ for $n$-ary predicates and if
we want to make precise which of the three presentations we use, we
write
\begin{equation}\label{equ:notation-predicates}
\underline{\phi}\in [n,Q_0 A] \quad\quad \phi = \hat \phi \in
[Fn,QA]\quad\quad \check\phi \in [A,QFn].
\end{equation}
\medskip\noindent Next we show how elements $l\in LFn$ are $n$-ary
modal operators. Given an $n$-ary predicate $\phi$ on $A$, the `modal
operator' $l$ induces an predicate on $\widetilde{\T} A$ as follows.
\begin{equation}\label{equ:tr-pred-lift}
\xymatrix{
\widetilde{\T} A\ar[r]^{\widetilde{\T}(\check a)} & \widetilde{\T} QFn \ar[r]^{\rho_{Fn}(l)} & \Omega
}
\end{equation}
This shows that the meaning of the modal operator $l\in LFn$ is fully
determined by the image $\rho_{Fn}(l)\in Q\widetilde{\T} QFn$. We turn this
observation into a definition.
\begin{definition}
Elements of $Q\widetilde{\T} QFn$ are called $n$-ary predicate liftings. Each
$\lambda\in Q\widetilde{\T} QFn$ induces a natural transformation
\begin{equation}\label{equ:pred-lift-new}
\begin{array}{rl}
[Fn,QA] & \rightarrow Q\widetilde{\T} A\\
\phi & \mapsto \lambda\circ\widetilde{\T}(\check\phi)
\end{array}
\end{equation}
\end{definition}
\begin{example}\label{exle:pred-lift}
Consider $B=\P_{\omega}$, $T=\{*\}+\mathit{Act}\times\mathit{Id}$, $\widetilde{\T}(A)=F\{*\} +
\mathit{Act}\cdot A$. As in Example~\ref{exle:B=Pom:2}, we identify
$F\emptyset$ with $\{\emptyset\}$ and $\widetilde{\T}^n(F\emptyset)$ with
$\P_{\omega}(1+\mathit{Act} +\ldots \mathit{Act}^n)$. The initial and final $\widetilde{\T}$-algebras
are then $\P_{\omega}(\mathit{Act}^*)$ and $\P(\mathit{Act}^*)$,
respectively.
Recall that $QA=[A,F1]=[A,\mathbbm{2}]$ and we write
$0,1\in\mathbbm{2}$. Further note that, for finite $n$, there is a
bijection $UQFn=U[Fn,\mathbbm{2}]\cong\mathit{Set}(n,2)\cong Bn= UFn$ which
extends to a semi-lattice isomorphism $QFn\cong Fn$.
In order to obtain the clause for $\surd$, we instantiate
(\ref{equ:pred-lift-new}) with $n=\emptyset$ (because $\surd$ is a
constant) and let $\lambda_\surd$ be the unique isormorphism
\begin{equation}
\widetilde{\T} QF\emptyset\cong \widetilde{\T} F\emptyset = F\{*\}
+ \mathit{Act}\cdot F\emptyset \cong F\{*\} \longrightarrow \mathbbm{2}.
\end{equation}
Consider $A$ and $\phi:F\emptyset\rightarrow QA$ and $\check\phi:A\rightarrow
QF\emptyset\cong F\emptyset$. This gives us the semantics of $\surd$
as follows.
$\delta_A(\surd)\in Q\widetilde{\T} A$ as in \eqref{equ:deltarho} is the map
\begin{equation}\label{equ:sem-tick}
\xymatrix{
F\{*\} + \mathit{Act}\cdot A \ar[r]^{\quad\quad\quad\delta(\surd)}
\ar[d]_{F\{*\}+\mathit{Act}\cdot \check\phi}
& \mathbbm{2} \\
F\{*\} + \mathit{Act}\cdot F\emptyset \ar[ur]_{\lambda_\surd}
}
\end{equation}
Finally, putting this together with (\ref{equ:tr-semantics}) and
(\ref{equ:sem-set-coalg}) we find that, as expected,
$$x\Vdash \surd \ \Leftrightarrow *\in \gamma(x).$$
In order to obtain the clause for $\langle a \rangle\phi$, we
instantiate (\ref{equ:pred-lift-new}) with $n=1$ and let $\lambda_a$
be given by the map
\begin{equation} \widetilde{\T} QF1 \cong \widetilde{\T} F1 = F\{*\} +
\mathit{Act}\cdot F1 \longrightarrow \mathbbm{2}
\end{equation}
which sends all generators $*$ and $b\in A, b\not=a$ to $0$ and $a$ to
1. Consider $A$ and choose some $\phi:F1\rightarrow QA$. Note that
$\check\phi:A\rightarrow QF1\cong F1\cong\mathbbm{2}$. This gives us the semantics
of $\langle a \rangle\phi$ as follows.
$\delta(\langle a \rangle\phi)\in Q\widetilde{\T} A$
as in \eqref{equ:deltarho} is the map
\begin{equation}\label{equ:sem-actions}
\xymatrix{
F\{*\} + \mathit{Act}\cdot A \ar[r]^{\quad\quad\quad\delta(\langle a
\rangle\phi)}
\ar[d]_{F\{*\}+\mathit{Act}\cdot \check\phi }
& \mathbbm{2} \\
F\{*\} + \mathit{Act}\cdot F1 \ar[ur]_{\lambda_{a}}
}
\end{equation}
Finally, putting this together with (\ref{equ:tr-semantics}) and
(\ref{equ:sem-set-coalg}) we find that, as expected,
$$x\Vdash \langle a \rangle\phi \ \Leftrightarrow \ (a,x')\in\gamma(x)\ \textrm{and}
\ x'\Vdash\phi.$$
\end{example}
Every collection of predicate liftings defines a functor.
\begin{definition}
Given a collection of predicate liftings $\Lambda$ let $L_\Lambda
A=F\coprod_{\lambda\in\Lambda} [F(n_\lambda),A]$, where $n_\lambda$
is the arity of $\lambda$. The semantics $\delta_\Lambda$ acts on a
generator $(\lambda,\phi)\in Q\widetilde{\T} QFn\times [Fn,QA]$ as given by
\eqref{equ:pred-lift-new}.
\end{definition}
\begin{example}
Let $\Lambda=\{\lambda_\surd\}\cup\{\lambda_a\mid a\in\mathit{Act}\}$ as in
Example~\ref{exle:pred-lift}. Then $L_\Lambda A \cong F1 + \mathit{Act}\cdot
FUA$ and $\delta_\Lambda$ is given by \eqref{equ:sem-tick} and
\eqref{equ:sem-actions}.
\end{example}
It is possible to incorporate logical laws into the functor.
\begin{example}\label{exle:LambdaE}
Let $\Lambda=\{\lambda_\surd\}\cup\{\lambda_a\mid a\in\mathit{Act}\}$ as in
Example~\ref{exle:pred-lift} and consider the set $E$ of equations
given by \eqref{equ:exle1-axioms}. Then $L_{\Lambda E} \cong F1 +
\mathit{Act}\cdot\mathit{Id}$ and $\delta_{\Lambda E}$ is given by
\eqref{equ:sem-tick} and \eqref{equ:sem-actions}. Furthermore, we
have
\begin{equation}
\label{equ:deltaLE-kappa}
\vcenter{
\xymatrix@R=5pt{
F1+Act\cdot Q \ar[dd]_\cong\ar[dr]^{\kappa} & \\
& Q\widetilde{\T} \\
L_{\Lambda E}Q \ar[ur]_{\delta_{\Lambda E}}
}}
\end{equation}
where, on finite $A$, $\kappa_A$ is the isomorphism
\begin{equation}
\label{equ:kappa}
\xymatrix@R=5pt{
F1+Act\cdot QA \ar[r]
& F1\times\prod_\mathit{Act} QA \ar[r]
& Q(F1+Act\cdot A) \ar[r]^{}
& Q\widetilde{\T} A
}
\end{equation}
where the first iso comes from \eqref{equ:co-prod}, the second is due
to $Q$ being a hom-functor, and the third is from the definition of
$\widetilde{\T}$.
\end{example}
To summarise, we have extracted from the example in
Section~\ref{sec:examples} a general framework that allows to define
trace logics for general functors $T$ and monads $B$ satisfying
Assumption~\ref{ass:fin}.
\subsection{A generic trace logic}\label{sec:gen-logic}
In this section, we show how to define a logic $(L_T,\delta_T)$ for
general functors $T$ and monads $B$ satisfying
Assumption~\ref{ass:fin}. We show that the example from the previous
section arises in that way.
\begin{definition}\label{def:L}
The functor $L_{T}:B\mbox{-}\mathit{Alg}\toB\mbox{-}\mathit{Alg}$ is defined on finitely generated
free algebras $Fn$ as $L_{T}Fn=Q\widetilde{\T} QFn$. Since every $A\inB\mbox{-}\mathit{Alg}$
is a colimit of finitely generated free algebras, this extends
continuously to all $A\inB\mbox{-}\mathit{Alg}$.
\end{definition}
\begin{definition}\label{def:LT}
The semantics $\delta_T : L_{T} Q \rightarrow Q \widetilde{\T}$ is given by
considering $QA$ as a colimit $\phi_i:Fn_i\rightarrow QA$, which is, by
construction, preserved by $L_T$. More explicitly,
$(\delta_{T})_{X}$ is the unique arrow making the following diagram
\begin{equation}\label{equ:deltaT}
\xymatrix{
QA & L_{T} QA \ar[r]^{(\delta_T)_A} & Q\widetilde{\T} A\\
Fn_i\ar[u]^{\phi_i} & L_TFn_i \ar[u]^{L_T\phi_i}\ar[r]^{=} & Q\widetilde{\T} QFn_i
\ar[u]_{Q\widetilde{\T}\check \phi_i} }
\end{equation}
commute for each $i$; as in \eqref{equ:notation-predicates}, the
arrow $\check \phi_i$ comes from applying the isomorphism
$B\mbox{-}\mathit{Alg}(Fn_i, QA)\cong B\mbox{-}\mathit{Alg}(A,QF n_i)$ to $\phi_i$.
\end{definition}
To show that the example of the previous section is actually the
generic one, we need a lemma helping us to compare the two logics.
\begin{lemma}
Let $(L,\delta), (L',\delta')$ be two logics and $\rho, \rho'$ as in
\eqref{equ:deltarho}. If there is an isomorphism $\alpha:LJ\rightarrow L'J$
such that for all finite sets $n$ we have
\begin{equation}
\vcenter{
\xymatrix@R=5pt{
LFn \ar[dd]_{\alpha_n}\ar[dr]^{\rho} & \\
& Q\widetilde{\T} QFn \\
L'Fn \ar[ur]_{\rho'}
}}
\end{equation}
then this extends to an isomorphism $\beta:L\rightarrow L'$ of logics, ie,
$\beta$ satisfies
\begin{equation}
\vcenter{
\xymatrix@R=5pt{
LQ \ar[dd]_{\beta Q}\ar[dr]^{\delta} & \\
& Q\widetilde{\T} \\
L'Q \ar[ur]_{\delta'}
}}
\end{equation}
Moreover, $\beta_{Fn}=\alpha_n$.
\end{lemma}
Consequently, any collection of isomorphisms $Ln\rightarrow Q\widetilde{\T} QFn$,
$n\in\mathbb{N}$, defines the same logic, or, more precisely:
\begin{corollary}\label{cor:LT}
The generic logic $L_T$ is determined up to isomorphism, that is,
for any other logic $(L,\delta)$ with the $LFn\rightarrow Q\widetilde{\T} QFn$ as in
\eqref{equ:deltarho} being isos, there is a unique isomorphism $L\rightarrow L_T$
such that
\begin{equation}
\vcenter{
\xymatrix@R=5pt{
LQ \ar[dd]_{}\ar[dr]^{\delta} & \\
& Q\widetilde{\T} \\
L_TQ \ar[ur]_{\delta_T}
}}
\end{equation}
\end{corollary}
Finally, we can show that the generic logic of this subsection agrees
with the logic defined, in different ways, by
\eqref{equ:exle1-syntax}-\eqref{equ:exle1-axioms}, or again in
Example~\ref{exle:two} or in Example~\ref{exle:LambdaE}.
\begin{proposition}
Going back to Example~\ref{exle:LambdaE}, there is an isomorphism
such that
\begin{equation}
\xymatrix@R=5pt{
L_{\Lambda E}Q\ar[dd]_\cong\ar[dr]^{\delta_{\Lambda E}} & \\
& Q\widetilde{\T} \\
L_TQ \ar[ur]_{\delta_T}
}
\end{equation}
\end{proposition}
\begin{proof}
We write $(L,\delta)$ for $(L_{\Lambda E},\delta_{\Lambda E})$ and
$\rho$ for the natural transformation as in
\eqref{equ:deltarho}. According to Corollary~\ref{cor:LT}, it is
enough to show that $\rho_{Fn}:LFn\rightarrow Q\widetilde{\T} QFn$ is an
isomorphism. From the proof of Proposition~\ref{prop:pred-lift}, we
know that $\rho_{Fn}=\delta_{QFn}\circ L\eta$. Since $\eta$ is an
isomorphism for finite semi-lattices, the result now follows from
$\delta_{QFn}$ being iso, see Example~\ref{exle:LambdaE}.
\end{proof}
Finally, Definition~\ref{def:LT} does not depend on the choice of a
partiuclar $T$ or $B$, so we can summarise this section as follows.
\begin{theorem}
For every monad $B$ on $\mathit{Set}$ and functor $T:\mathit{Set}\rightarrow\mathit{Set}$ satisfying
Assumption~\ref{ass:fin} there is a generic trace logic.
\end{theorem}
Of course, given $B$ and $T$, the real work consists in finding a good
explicit description of the generic logic. We have illustrated this
for the moment only with one example.
We can apply the general framework to obtain results about generic
logics. For example, we have
\begin{theorem}
The logic of Example~\ref{exle:LambdaE} is expressive and complete.
\end{theorem}
\begin{proof}
We write $(L,\delta)$ for $(L_{\Lambda E},\delta_{\Lambda E})$. The
proof is straightforward due to the following facts: $B$ and $\widetilde{\T}$
preserve finite algebras and on finite algebras we have that
$\delta$ is an isomorphism. In detail:
Expressiveness means that any two non-trace equivalent states can be
separated by a formula. Consider a coalgebra $X\rightarrow BTX$ with
$x,x'\in X$ and suppose $x$ accepts trace $t$ and $x'$ does
not. Since the initial $L$-algebra is the free $B$-algebra over the
set of traces, $t$ can be considered as a formula and we have
$x\Vdash t$ and $x'\not\Vdash t$.
Completeness means that if $L$ does not prove $\phi=\phi'$, then
there must be a coalgebra $X\rightarrow BTX$ and $x\in X$ such that, wlog,
$x\Vdash\phi$ and $x\not\Vdash\phi'$. Since $\phi$ and $\phi'$ appear
at some stage $n$ in the initial algebra construction of $L$, the
semantics of $\phi$ and $\phi'$ is determined at stage $n$. Since
$\delta$ is an iso on finite algebras, the images
of $\phi$ and $\phi'$ in $Q\widetilde{\T}^nF\emptyset$ are different. It
follows from a standard argument that there is a $\widetilde{\T}$-coalgebra
${\widetilde\gamma}:\widetilde{\T}^nF\emptyset\rightarrow\widetilde{\T}(\widetilde{\T}^nF\emptyset)$ that refutes
the equation $\phi=\phi'$. In particular,
$\sem{\phi}_{\widetilde\gamma}\not=\sem{\phi'}_{\widetilde\gamma}$ are two different
morphisms $FT^n\emptyset=Q\widetilde{\T}^nF\emptyset\rightarrow\mathbbm{2}$, so they must
differ on some generator $\eta_X(x)$ where $\eta_X:X\rightarrow BX$ maps
elements $x$ to singletons $\{x\}$. It follows now from
Proposition~\ref{prop:wtgamma} that the $(B,T)$-coalgebra
$U\gamma\circ\eta_X:X\rightarrow BTX$ contains a state $x$ with
$x\Vdash\phi$ and $x\not\Vdash\phi$.
\end{proof}
|
\section{Introduction}
The number of degrees of freedom of a communication channel plays an
important role in evaluating the channel's Shannon capacity (see
e.g.~\cite{Somaraju2009b}). In a physical communication system,
because of constraints such as finite transmission power and noise at
the receiver, only finitely many (linearly independent) signals may be
exchanged between the transmitter and receiver. This physical
intuition may be captured using the concept of number of degrees of
freedom of the communication channel. The number of degrees of freedom
of a communication channel has been used in evaluating the Shannon
capacity for several physically realistic channels\footnote{The concept of number of degrees of freedom is used implicitly in evaluating
the capacity of Shannon's classical bandwidth limited, additive Gaussian white noise channel. In particular it can be shown that as the bandwidth becomes large the number of degrees of freedom at level $\epsilon$ approaches the well know constant $2WT$, for all noise levels $\epsilon$ (see Gallager~\cite[Ch. 8]{Gallager1968}).} (see
e.g.~\cite{Somaraju2009b,Shannon1948,Biglieri1998,Gallager1968}). This
concept has also been used in various other problem domains such as
multi-antenna communication~\cite{Migliore2006,Hanlen2006,Xu2006},
optics~\cite{Miller2000,Piestun2000} and electromagnetic field
sampling~\cite{Bucci1989}.
Kolmogorov numbers are a particular example of so called $s$-number
sequences. In the theory of $s$-numbers one associates with every
bounded linear operator $T$, mapping between any two Banach spaces,
a scalar sequence $s_1(T)\geq s_2(T)\geq \ldots\geq 0$ (see
e.g.~\cite{Pietsch1980,Pietsch1987}).
A classical example in the more restricted category of compact operators
mapping between Hilbert spaces is the sequence of singular values.
$s$-number sequences of various
types have primarily been used to classify operators based on the behaviour
of the sequence $s_n$ as $n\to \infty$. In particular, various
interesting operator ideals have been obtained based on the behaviour of these
sequences.
In this note we establish the connection between Kolmogorov numbers
and degrees of freedom of a communication channel. The remainder of
this note is organised as follows: in Section~\ref{sec:definitions}, we recall
the definitions of degrees of freedom of a communication channel and
briefly explain the physical intuition behind the definition. We also
recall the definitions of $s$-number sequences in general and Kolmogorov
numbers in particular. We present our main results, establishing the connection
between the number of degrees of freedom and Kolmogorov numbers, in
Section~\ref{sec:maiRes} and provide conclusions in the final
Section~\ref{sec:conclusion}.
\section{Definitions}\label{sec:definitions}
\subsection{Degrees of Freedom}
The number of degrees of freedom of a communication channel has for
example been defined in~\cite{Somaraju2009b}. The physical model used
in~\cite{Somaraju2009b} is as follows: there is a normed vector space
$X$ of transmitter functions and a normed vector space $Y$ of receiver
functions and a channel operator $T:X\to Y$ that is a compact linear
operator. Physically, the elements of $X$ can be thought of as signals
that a transmitter generates, and $T$ maps these signals onto elements
of $Y$, which is the space of signals a receiver can measure. It is
further assumed that there are two constraints on this channel: 1)
the power available for transmission is finite. This constraint is
modeled by assuming that the transmitter may only generate elements
of $X$ that have a norm less than or equal to $1$. 2) the receiver is noisy and
therefore small signals can not be measured. This constraint is
modeled by assuming that only received signals with norm greater than
some pre-specified small constant $\epsilon$ may be measured.
It was shown in~\cite{Somaraju2009b} that the following definition may be used for the number of degrees of freedom of a compact operator that models a communication channel.
\begin{definition}[Degrees of freedom at
level~$\epsilon$]\label{Ndef:banDof} Suppose $ X$ and $ Y$
are normed spaces with norms $\|\cdot\|_{ X}$ and $\|\cdot\|_{ Y}$, respectively, and $T: X\to Y$ is a compact operator. Also, let $\bar{B}_{r,X}(\theta)$ denote the closed ball of radius $r$ centered at $\theta$ in $X$.
Then the number of degrees of freedom $\cN(\epsilon)$ of $T$ at level~$\epsilon$ is the
smallest\footnote{$\bbZ_0^+$ denotes the set of non-negative integers.} $N\in \bbZ_0^+$ such that there exists a set of vectors
$\{\psi_1,\ldots,\psi_N\}\subset Y$ such that for all $x\in \bar{B}_{1,X}(0)$
\begin{eqnarray*}
&\inf_{a_1,\ldots,a_N} \left\|Tx-\sum_{i=1}^N a_i\psi_i\right\|_{ Y}
\leq \epsilon.&\\&&
\end{eqnarray*}
\end{definition}
It was shown in~\cite{Somaraju2009b} that the number of degrees of
freedom is well defined.
Physically, we interpret this definition as follows: if there is some constraint $\|\cdot\|_X\leq 1$ on
the space of transmitter functions\footnote{We normalised the
transmitter power constraint by assuming that available transmission power, $P=1$. Different
normalisations can e.g. be achieved by re-scaling the norm on $X$.} and if the receiver can only measure
signals that satisfy $\|\cdot\|_Y > \epsilon$, then the number of
degrees of freedom is the maximum number of linearly independent
signals that the receiver can distinguish under these constraints, where
the maximum is taken over all possible signal constellations.
The following theorem is a simple consequence of the above definition.
\begin{theorem}~\cite[Th. 3.2]{Somaraju2009b}\label{th:dofProp}
Suppose $ X$ and $ Y$ are normed spaces with norms $\|\cdot\|_{X}$ and $\|\cdot\|_{ Y}$, respectively, and $T: X\to Y$ is a compact operator. Let $\cN(\epsilon)$ denote the number of degrees of freedom of $T$ at level~$\epsilon$. Then
\begin{enumerate}
\item \label{dofProp2} $\cN(\epsilon) = 0$ for all $\epsilon\geq\|T\|$.
\item \label{dofProp5} Unless $T$ is identically zero, there
exists an $\epsilon_0>0$ such that $\cN(\epsilon) \geq 1$
for all $0<\epsilon<\epsilon_0$.
\item \label{dofProp3} $\cN(\epsilon)$ is a non-increasing,
upper semicontinuous function of $\epsilon$.
\item \label{dofProp4} In any finite interval
$(\epsilon_1,\epsilon_2)\subset\bbR$, with $0<\epsilon_1<
\epsilon_2$, $\cN(\epsilon)$ has only finitely many
discontinuities, i.e. $\cN(\epsilon)$ only takes finitely
many non-negative integer values in any finite $\epsilon$ interval.
\end{enumerate}
\end{theorem}
\subsection{$s$-numbers and Kolmogorov numbers}
The axiomatic characterisation of $s$-numbers for Banach space valued
operators is due to Pietsch, cf.~\cite{Pietsch1987}. In the remainder
of this section, we assume that $X,X',Y$ and $Y'$ are Banach spaces,
$\cB(X,Y)$ denotes the set of bounded linear operators $T:X\to Y$ and
$\|\cdot\|$ is the standard operator norm on $\cB(X,Y)$. Also, we denote by $I$ the
identity operator.
\begin{definition}
Let $s:T\mapsto (s_n(T))_{n=1}^\infty$ be a rule that assigns to every bounded operator $T$ on some Banach space a scalar sequence such that the following conditions are satisfied:
\begin{description}
\item[SN1] $\|T\| = s_1(T)\geq s_2(T)\geq \ldots\geq 0$.
\item[SN2] $s_n(S+T) \leq s_n(S) +\|T\|$, for $S,T\in \cB(X,Y)$ and $n=1,2,\ldots$.
\item[SN3] $s_n(BTA) \leq \|B\|s_n(T)\|A\|$, for $A\in \cB(X',X),T\in \cB(X,Y), B\in \cB(Y,Y')$ and $n=1,2,\ldots$.
\item[SN4] $s_n(I) = 1$, for $n=1,2,\ldots$.
\item[SN5] $s_n(T) = 0$ if $\mathrm{rank}(T) < n$.
\end{description}
Then $(s_n(T))_{n=1}^\infty$ is called an \emph{s-number sequence}.
\end{definition}
It turns out that in the compact, Hilbert space case, i.e. if we restrict the
above definition to the class of compact operators mapping between Hilbert
spaces, the above properties SN1--SN5 uniquely characterise the singular values of the operator (ordered in descending order).
For any linear subspace $S$ of $Y$ let $Q_S^Y$ denote the natural surjection from $Y$ onto the quotient space $Y/S$. The numbers
\eqs{d_n(T) \triangleq \inf\{\|Q_S^YT\|: \mathrm{dim}(S) < n\}}
for $T\in \cB(X,Y)$ define an $s$-number sequence~\cite{Pietsch1987},
and are called \emph{Kolmogorov numbers}.
\section{Main results}\label{sec:maiRes}
Let $X$ and $Y$ be Banach spaces with norms $\|\cdot\|_X$ and
$\|\cdot\|_Y$, respectively, let $T:X\to Y$ be compact\footnote{We
only consider compact $T$, because the number of degrees of freedom
is only defined for compact operators. One can conclude from the
arguments in~\cite{Somaraju2009b} that for physically realistic
models of communication channels, the channel operator must be
compact. However, it may be possible to generalise the theory
presented in the current paper to bounded operators that are not
necessarily compact.} and suppose $B(X)$ is the unit ball in
$X$. Let $\cN(\epsilon)$ denote the number of degrees of freedom of
$T$ at level $\epsilon$ and let $d_n = d_n(T)$ denote the $n^{th}$
Kolmogorov number.
For ease of notaion we set $\cN(\epsilon) = \infty$ for $\epsilon <0$ for the remainder of this document.
Now, consider the sequence $\{\sigma_n =
\sigma_n(T)\},n = 1,2,\ldots$ implicitely defined as follows
\begin{eqnarray*}
&\sup_{\epsilon > \sigma_n} \cN(\epsilon) = n-1& \textrm{and}\\
&\inf_{\epsilon < \sigma_n} \cN(\epsilon) = N \geq n.&
\end{eqnarray*}
Further, if $n<N$ then for all $m$ such that $n < m \leq N$, $\sigma_m \triangleq
\sigma_n$. The sequence $\{\sigma_n\}$ was defined
in~\cite{Somaraju2009b} and $\sigma_n$ was called the $n^{th}$ DOF
singular value of $T$ in~\cite{Somaraju2009b}. It should be noted that
in the Hilbert space situation, $\sigma_n$ is simply the $n^{th}$
singular value of $T$. Also, $\cN(\epsilon)$ is
discontinuous exactly at $\epsilon =\sigma_n$ for $n=1,2,\ldots$. We now show that $\sigma_n$ is in fact equal to $d_n$ for which we will need the following elementary lemma.
\begin{lemma}
\eqs{d_{n+1} = \inf_{\psi_1,\ldots,\psi_n\in Y} \sup_{x\in B(X)} \inf_{a_1,\ldots,a_n\in \bbR} \left\|Tx - \sum_{i=1}^n a_i\psi_i\right\|}
\end{lemma}
\begin{proof}
Let $S$ be any subspace of $Y$ such that $\mathrm{dim}(S)\leq n$. By the definition of $Q_S^Y$,
\eqs{\|Q_S^YT\| = \sup_{x\in B(X)}\inf_{y\in S} \|Tx - y\|_Y.}
The lemma now follows from two simple facts: for $\mathrm{dim}(S)\leq n$
there exists a set of vectors $\{\psi_1,\ldots,\psi_n\}$ such that
they span $S$, and the dimension of the span of $n$ vectors is not greater than $n$.
\end{proof}
\begin{theorem}\label{th:mainRes}
$\sigma_n = d_n.$
\end{theorem}
\begin{proof}
Because $\sigma_1 = d_1 = \|T\|$, we only need to prove that $\sigma_{n+1} = d_{n+1}$ for $n = 1,2,\ldots$.
Now suppose $\sigma_N = \sigma_{N-1} = \ldots = \sigma_{n+1} < \sigma_n$. We first prove that $\sigma_{n+1} \geq d_{n+1}$. Let $\delta > 0$ be some arbitrary number satisfying
$\delta < \sigma_n - \sigma_{n+1}$. Then for $\epsilon = \sigma_{n+1} + \delta$,
\begin{equation*}
\cN(\epsilon) = n.
\end{equation*}
Therefore, there exists a set $\{\psi_1,\ldots,\psi_n\}\in Y$ such that
\begin{equation*}
\sup_{x\in B(X)} \inf_{a_1,\ldots,a_n} \left\|Tx-\sum_{i=1}^n a_i\psi_i\right\|_{ Y}
\leq \epsilon = \sigma_{n+1} + \delta.
\end{equation*}
Because, the constant $\delta> 0$ can be made arbitrarily small,
\begin{eqnarray}
d_{n+1} &=& \inf_{\psi_1,\ldots,\psi_n\in Y} \sup_{x\in B(X)} \inf_{a_1,\ldots,a_n\in \bbR} \left\|Tx - \sum_{i=1}^n a_i\psi_i\right\|\nonumber\\ &\leq& \sigma_{n+1}\label{e1}.
\end{eqnarray}
Relabeling $N = n+1$, we get
\begin{equation*}
d_N \leq \sigma_N = \sigma_{n+1}.
\end{equation*}
To prove the converse inequality, let $\epsilon < \sigma_{N}$. Then, $\cN(\epsilon) > N$. Therefore, for all sets $\{\psi_1,\ldots,\psi_N\}\subset Y$, there exists an $x\in B(X)$ such that,
\begin{equation*}
\inf_{a_1,\ldots,a_N} \left\|Tx-\sum_{i=1}^{N} a_i\psi_i\right\|_{ Y}
> \epsilon > \sigma_{N} = \sigma_{n+1}.
\end{equation*}
Therefore,
\begin{eqnarray}
d_{n+1} &\geq& d_{N}\nonumber\\ &=& \inf_{\psi_1,\ldots,\psi_N\in Y} \sup_{x\in B(X)} \inf_{a_1,\ldots,a_N\in \bbR} \left\|Tx - \sum_{i=1}^{N} a_i\psi_i\right\|\nonumber\\ &\geq& \sigma_{N}\nonumber\\ &=& \sigma_{n+1}\label{e2}.
\end{eqnarray}
Here, we used the $s$-number property SN1 that $d_n$ is non-increasing in $n$. The result now follows from~\eqref{e1} and~\eqref{e2}.
\end{proof}
\begin{remark} The number of degrees of freedom of a communication
channel is defined for compact operators mapping between arbitrary normed
spaces, while the concept of Kolmogorov numbers is defined only
for operators mapping between Banach spaces.
However, if $X$ and $Y$ are normed spaces and $T:X\to Y$ is compact,
then consider the completions $\tilde{X}$ and $\tilde{Y}$ of $X$ and
$Y$, respectively, so that $\tilde{X}$ and $\tilde{Y}$ are Banach
spaces.
Let $\tilde{T}:\tilde{X}\to\tilde{Y}$ be the standard extension of
$T$ from $X$ to $\tilde{X}$.
Then $\sigma_n(T) = \sigma_n(\tilde{T})$. This fact follows directly
from the compactness of $T$ and the definition of $\cN(\cdot)$.
\end{remark}
We also restate below a useful result in~\cite{Somaraju2009b} that aids in the numerical computation of Kolgomorov numbers for compact operators.
\begin{theorem}~\cite[Th. 3.8]{Somaraju2009b} Suppose $X$ and $ Y$ are Banach spaces
and $T: X\to Y$ is a compact operator. Also suppose that $X$ has a
complete Schauder basis $\{\phi_1,\phi_2,\ldots\}$ and let
$S_m=\mathrm{span}\{\phi_1,\ldots,\phi_m\}$. Let $T_m = T|_{S_m}:S_m\to Y$,
$m \in \bbZ^+$. If $\sigma_n$, the $n^{th}$ DOF singular value of $T$, exists then for
$m$ large enough $\sigma_{n,m}$, the $n^{th}$ DOF singular value of
$T_m$, will exist and
\begin{equation*}
\lim_{m\to\infty}\sigma_{n,m} = \sigma_n.
\end{equation*}
If $\sigma_{n,m}$ exists then it is a lower bound for
$\sigma_n$.
\end{theorem}
We can therefore use finite dimensional approximations of $T$ to
numerically compute the Kolgomorov numbers.
\section{Conclusion}\label{sec:conclusion}
In this note we establish the connection between Kolgomorov numbers
and degrees of freedom of a communication channel. Specifically, we
show that the jump points (discontinuities) of the function relating
the number of degrees of freedom to the noise level $\epsilon$ are
equal to the Kolgomorov numbers.
This connection invites the question as to whether other $s$-number
sequences such as Gelfand numbers or approximation numbers play an
equally important role in communication systems.
|
\section{A general combinatorial fact}
Let $k$ be a field of characteristic zero, $n\in\mathbb{N}$ and $A=k[x_1,\dots,x_n]$ the polynomial co-algebra. Let $C(A)$ be its bar complex, computing $Tor_A(k,k)$. Since $A$ is Koszul it is well known that the homology $H(C(A))=Tor_A(k,k) \cong k[\epsilon_1,\dots, \epsilon_n]$ is the Koszul dual algebra, where $\epsilon_j$ are odd (degree one) variables.
On the complex $C(A)$ there is a $\mathbb{N}^n_0$-grading by the numbers of $x_1,\dots, x_n$ occuring. Let $C(A)^{(1,\dots,1)}$ be the degree $(1,\dots,1)$ subcomplex. For instance, if $n=3$ then $x_1\otimes x_2x_3$ is inside this subcomplex, while $x_1\otimes x_2x_1$ is not. It follows from the above that
\begin{prop}
\[
H(C(A)^{(1,\dots,1)})=
( k[\epsilon_1,\dots, \epsilon_n])^{(1,\dots,1)}=k\,\epsilon_1\dots \epsilon_n \cong k[-n].
\]
\end{prop}
\begin{rem}
As noted by V. Drinfeld \cite{drinfeld}, the complex $C(A)^{(1,\dots,1)}$
is the cosimplicial complex of an $n$ dimensional hypercube, relative to its boundary.
\end{rem}
Note further that $C(A)^{(1,\dots,1)} \cong \oplus_m (k^m)^{\otimes n}$, $m$ being the degree. Note also that the differential is invariant
under the $S_n$ action by permuting the factors. Since taking (co-)invariants under finite group actions commutes with taking homology, we obtain:
\begin{cor}
\label{cor:main}
Let $G\subset S_n$ be a subgroup and $M$ some right $G$-module. Then
\[
H( M \otimes_G C(A)^{(1,\dots,1)} ) = M \otimes_G \sgn_n
\]
where $\sgn_n$ is the one dimensional sign representation of $G$, concentrated in degree $n$.
\end{cor}
\begin{defi}
We call the complex $M \otimes_G C(A)^{(1,\dots,1)}$ the \emph{cubical complex} $\mathit{Cub}(G,M)$ of the pair $(G, M)$.
\end{defi}
\begin{rem}
There is a similar complex based on the Harrison complex $\mathit{Harr}(A)$, namely
\[
\mathit{Harr}(G,M) = M \otimes_G \mathit{Harr}(A)^{(1,\dots,1)},
\]
which we call the Harrison complex of the pair $(G,M)$.
It is well known that $H (\mathit{Harr}(A))\cong k^n[-1]$. Hence the following is also well known:
\[
H (\mathit{Harr}(G,M))\cong
\begin{cases}
0 &\text{for $n>1$}\\
M[-1] &\text{for $n=1$}.
\end{cases}
\]
\end{rem}
\begin{rem}
It is actually sufficient to consider the case $G=S_n$. If $G\subsetneq S_n$, one can equivalently take $G'=S_n$, $M'= M\otimes_G k[S_n]=Ind_G^{S_n}M$.
\end{rem}
\section{Examples}
Let us take $G=S_n$ and $M=\Lie(n)$ the $n$-ary operations in the Lie operad. By definition,
\begin{multline*}
\oplus_n\mathit{Cub}(S_n,\Lie(n))=
\oplus_n \Lie(n) \otimes_{S_n} C(A)^{(1,\dots,1)} =\\ =\oplus_{m,n} \Lie(n) \otimes_{S_n} (k^m)^{\otimes n}
= \oplus_m \FreeLie(k^m) =: \mathfrak{lie}
\end{multline*}
is a complex built from the free Lie algebras, which one can check is the same as the one considered by A. Alekseev and C. Torossian \cite{AT}.
One hence obtains:
\begin{cor}
\[
H(\mathfrak{lie}) = \oplus_n \Lie(n) \otimes_{S_n} \sgn_n = k[-1] \oplus k[-2].
\]
\end{cor}
\begin{proof}
The first equality is the previous Corollary, second follows from the fact that there are no fully antisymmetric elements in $\Lie(n)$ for $n\geq 3$ by the Jacobi identity.
\end{proof}
This cohomology was computed by M. Vergne \cite{vergne} (the second part of Thm 2.3 in \cite{vergne}).
A simpler and very similar example is $G=S_n$ and $M=k[S_n]=\Ass(n)$, the space of $n$-ary operations in the associative operad. In this case $\oplus_n\mathit{Cub}(S_n,k[S_n])$ is a complex buit from free associative algebras and its cohomology is $\oplus_n k[S_n]\otimes_{S_n}\sgn_n=\oplus_n k[-n]$.
This is the first part of Theorem 2.3 in \cite{vergne}.
Next, take again $G=S_{n}$ and $M=k[S_{n}]_{C_n}$ where $C_n\subset S_n$ is the cyclic group. Equivalently, $M=\Ass(n-1)$, where we view $\Ass$ as a cyclic operad. In this case we get the complex $\mathfrak{tr}$ defined by Alekseev and Torossian \cite{AT}
\begin{align*}
\mathfrak{tr} &= \oplus_n k[S_n]_{C_n}\otimes_{S_n} C(A)^{(1,\dots,1)}\cong \oplus_{m,n} k[S_n]_{C_n}\otimes_{S_n} (k^m)^{\otimes n} = \oplus_{m,n}( (k^m)^{\otimes n})_{C_n} \\
&= \oplus_{m} \FreeAss(k^m) / [ \FreeAss(k^m) , \FreeAss(k^m)].
\end{align*}
Applying Corollary \ref{cor:main} we get
\begin{cor}
\[
H(\mathfrak{tr}) = \oplus_n k[S_n]_{C_n}\otimes_{S_n} \sgn_n = \oplus_n k \otimes_{C_n} \sgn_n \cong \oplus_{n\,\text{odd}} k[-n]
\]
\end{cor}
We can also take $G=S_{n}$ and $M=\Lie(n-1)$. This is possible since $\Lie$ (-as a suboperad of $\Ass$-) is a cyclic operad. Consider
\[
\oplus_n \Lie(n) \otimes_{S_{n+1}} C(A)^{(1,\dots,1)} = \oplus_{m,n} \Lie(n) \otimes_{S_{n+1}} (k^m)^{\otimes n}
= \oplus_m \mathfrak{sder}_m =: \mathfrak{sder}.
\]
Note that this space coincides with the total space of the operad of Lie algebras $\mathfrak{sder}$ defined in \cite{AT}, which in turn is the same as the $\mathcal{F}$ defined in \cite{drinfeld}. By a similar argument as before, we hence obtain:
\begin{cor}
\[
H(\mathfrak{sder}) = \oplus_n \Lie(n) \otimes_{S_{n+1}} \sgn_{n+1} \cong k[-3].
\]
\end{cor}
\begin{rem}
All examples above have analogs for their respective Harrison complexes. Concretely, the cohomology in this case is always given by the $n=1$-part of the cohomology computed in the above propositions. We leave the details to the reader.
\end{rem}
\section{Generalizations and further examples in the literature}
There is a generalization to the above trick. Let $\mathcal P$ be any cooperad and $A$ the free $\mathcal P$-coalgebra on $k^n$. Then $A$ is endowed with a $\mathbb{Z}^n$-grading, along with an $S_n$ action. Any ``sensible'' homology theory of $\mathcal P$-coalgebras will produce a complex $C(A)$, inheriting this grading and $S_n$ action. For any right $S_n$ module $M$ one can hence form the complex
\[
M\otimes_{S_n} C(A)^{(1,\dots,1)}
\]
with cohomology $M \otimes_{S_n} H(C(A))^{(1,\dots,1)}$. This situation occurs frequently in nature, for example, taking for $\mathcal P$ the coassociative cooperad and $C(A)$ the cyclic complex, one can produce a proof of Theorem 8 in \cite{barnatan}.
Taking $P=e_2^*\{2\}$ the co-Gerstenhaber cooperad (up to degree shift), and for $C(A)$ the Gerstenhaber cohomology, one can produce a shorter proof of Proposition 21 in \cite{thomas}.
We want to emphasize that the combinatorial trick described here is not novel, and has been used many times in the mathematical literature, in various forms and by various authors.\footnote{In particular, D. Bar-Natan made us aware of his article \cite{barnatan2}, where the case of general $M$ is discussed as well.}
This paper was written merely to advertise a clean version of the trick and point out some further applications.
\bibliographystyle{plain}
|
\section{Introduction}
The search for supersymmetry and its breaking, in addition to the direct searches at LEP, B-factories, Tevatron and the
Large Hadron Collider (LHC), is actively pursued using the WMAP limits on the relic density constraints.
However the sensitivity of the lightest
supersymmetric particle relic density calculation to the variation of the cosmological expansion rate before Big-Bang
Nucleosynthesis (BBN), even if modest and with no consequences on the cosmological observations, can modify considerably
the relic density, and therefore change the constraints on the supersymmetric parameter space \cite{Arbey:2008kv,Arbey:2009gt}.
In the standard cosmology the dominant component before BBN is radiation, however energy density and entropy content can be
modified. In the following we
consider the impact of different scenarios of alternative cosmologies. The precision of the WMAP data should therefore not make
us forget the hypothesis which are implied by the use of standard cosmology. We discuss in the following the implications of
precision B-physics, direct searches and cold dark matter relic abundance for the case of anomaly mediated models, from a
minimal anomaly mediation supersymmetry breaking \cite{amsb}, to mixed moduli-anomaly
mediated \cite{Choi:2005uz} and to hypercharge anomaly mediation \cite{Dermisek:2007qi}. We also discuss similar
supersymmetry breaking scenarios in the case of the next-to-minimal supersymmetric standard model.
In section \ref{sec:models} we discuss the different anomaly mediated supersymmetry breakings in the Minimal Supersymmetric
Standard Model (MSSM) in terms of the parameter spaces for these models and discuss the limits which can be applied due
to the present data from particle physics experiments and from relic dark matter density assuming the standard
model of cosmology. In section \ref{sec:models2} similar scenarios are considered in the Next-to-Minimal Supersymmetric
Standard Model (NMSSM) and the corresponding particle and cosmological bounds are discussed.
In section \ref{sec:cosmologies} we discuss how alternative cosmological models can affect dramatically the bounds on
the parameter space of the models we have considered while letting unchanged the observable cosmology. Four different
alternatives to the standard cosmology are discussed which share this behaviour.
Section \ref{sec:constr} and section \ref{sec:lhc} discuss respectively the constraints implied by these different cosmological
scenarios and the perspectives at the LHC for a list of benchmark points which are representative of the available parameter space for
these AMSB models. Our conclusions are given in section \ref{sec:conclusion}.
\section{Anomaly mediated symmetry breaking in the MSSM}
\label{sec:models}
The superconformal Anomaly Mediated Supersymmetry Breaking (AMSB) mechanism \cite{amsb} is one of the most well-known and
attractive set-ups for supersymmetry breaking. Supersymmetry breaking effects in the observable
sector have in this framework a gravitational origin.
Superconformal symmetry is classically preserved in theories without dimensional parameters and it is in general
broken by the quantum effects. As anomalies only depend on the low-energy effective theory, the same will be true for
the soft terms. Usually the AMSB scenario cannot be applied to the MSSM, as it leads to tachyonic sleptons. However the
presence of an intermediate threshold can displace the soft terms and avoid this problem.
In superconformal gravity one introduces a chiral superfield playing the role of the compensating multiplet for super-Weyl
transformations, called the Weyl or conformal compensator. The F-term vacuum expectation value of the conformal
compensator is turned on by the supersymmetry breaking in the hidden sector and the soft breaking of supersymmetry in the
visible sector appears through the chiral anomaly supermultiplet. As the soft SUSY breaking terms arise from the anomaly, the
supersymmetry breaking terms do not dominate at tree-level. Several soft SUSY breaking scenarios
can be realised starting from this setup. We discuss in the following some of these realisations.
\subsection{Minimal AMSB}
The minimal AMSB (mAMSB) scenario \cite{amsb} has very attractive properties, since the soft SUSY breaking terms are calculated in terms
of one single parameter, namely the gravitino mass $m_{3/2}$, and the soft terms are renormalisation group invariants which can
be calculated for any scale choice. However, the AMSB scenarios suffer from the problem that slepton squared masses are found
to be negative, leading to tachyonic states. A solution to this problem is to consider that the scalar particles acquire a universal mass $m_0$ at the GUT scale, which when added to the AMSB soft SUSY breaking terms, makes them positive. Therefore, the mAMSB model
relies on only four parameters:
\begin{equation}
m_0, m_{3/2}, \tan \beta, \mbox{sgn}(\mu) \;.
\end{equation}
This scenario has been thoroughly studied in the literature, but is known to have cosmological consequences incompatible with the
WMAP observations of the dark matter density \cite{Baer:2010kd}. We perform here an updated analysis of the mAMSB parameter space constraints
from flavour physics and cosmological relic density. For this study, we generate mass spectra and couplings using Isajet 7.80
\cite{isajet}. The calculation of flavour observables and the computation of the relic density are performed with SuperIso Relic v3.0 \cite{superiso,superiso_relic}. We use the constraints given in Table~9 of the latest version of the SuperIso manual.
The first flavour observable that we consider here is the branching ratio of $ B \to X_s \gamma$, which has been been thoroughly
studied in the literature and is still under scrutiny. This observable is very interesting, as its SM contributions only appear at loop
level, and its theoretical uncertainties as well as the experimental errors are now under control. It provides strong constraints on
the supersymmetric parameter space, especially for large $\tan \beta$, where it receives large enhancements from its
supersymmetric contributions. We use the following interval at 95\% C.L. \cite{superiso,Mahmoudi:2007gd}
:
\begin{equation}
2.16 \times 10^{-4} < \mbox{BR}(B \to X_s \gamma) < 4.93 \times 10^{-4} \;.
\end{equation}
Another interesting observable is the branching fraction of $B_s \to \mu^+ \mu^-$, which is also a loop level observable, and which
can receive extremely large contributions from SUSY at large $\tan\beta$, and can receive an enhancement of several orders of
magnitude compared to the SM branching ratio. This decay mode has not yet been observed, and we have at 95\% C.L. \cite{superiso,CDF_bsmumu}
:
\begin{equation}
\mbox{BR}(B_s \to \mu^+ \mu^-) < 4.7 \times 10^{-8} \;.
\end{equation}
We also consider a set of tree-level observables which are very sensitive to the charged Higgs mass as well as $\tan\beta$, and
we use the following 95\% level intervals, which include the theoretical and experimental errors \cite{superiso,Akeroyd:2009tn,Antonelli:2008jg}:
\begin{eqnarray}
0.56 < &\dfrac{\mbox{BR}(B \to \tau \nu)}{\mbox{BR}_{SM}(B \to \tau \nu)}& < 2.70 \;,\\
4.7 \times 10^{-2} < &\mbox{BR}(D_s \to \tau \nu )& < 6.1 \times 10^{-2} \;,\\
0.151 < &\dfrac{\mbox{BR}(B \to D^0 \tau \nu)}{\mbox{BR}(B \to D^0 e \nu)}& < 0.681 \;,\\
0.982 < &\mbox{R}_{\ell23}(K \to \mu \nu) & < 1.018 \;.
\end{eqnarray}
The observable $\mbox{R}_{\ell23}(K \to \mu \nu)$ is related to the decay of $K \to \mu \nu$ and is detailed
in \cite{Antonelli:2008jg}.
\\
For the relic density constraint, we use the WMAP constraints \cite{Komatsu:2010fb} increased by 10\% of theoretical error to account for the uncertainties in the calculation of the relic density:
\begin{equation}
0.088 < \Omega_{DM} h^2 < 0.123 \;.
\end{equation}
In the following, we disregard the case of negative $\mbox{sgn}(\mu)$ since it is disfavoured by the muon
anomalous magnetic moment
constraint, and we scan over the intervals $m_0 \in [0, 2000]$ GeV, $m_{3/2} \in [0, 100]$ TeV and $\tan \beta \in [0, 60]$. Figure
\ref{fig_amsb} presents projection plots of the parameter space into the possible different planes. The green region in the plots
corresponds to the parameter zone which is not excluded by flavour constraints or mass limits. The red stars corresponds to points
leading to a favoured relic density but excluded by other constraints, whereas black stars are favoured by all the presented
constraints, including the relic density constraint. As can be seen, no black star is visible in these plots, and the whole parameter space presented here is disfavoured either by flavour or direct constraints, or by the relic density constraint which tends to favour the low $m_{3/2}$ region. Disregarding the relic density constraint, a large zone at low $m_{3/2}$ is excluded.
\begin{figure}[p!]
\begin{center}
\vspace*{-2.cm}\includegraphics[width=16cm]{amsb_m32_m0.png}\\
\includegraphics[width=16cm]{amsb_m0_tanb.png}\\
\includegraphics[width=16cm]{amsb_m32_tanb.png}
\end{center}
\caption{\label{fig_amsb} Constraints on the minimal AMSB parameter space. The exclusion regions are plotted in the order given
in the legend. The red zones are excluded by the inclusive branching ratio of $B \to X_s \gamma$, the yellow ones correspond to
charged LSP, the olive-green areas are excluded by direct collider constraints, the light blue zones are excluded by
BR($B\to\tau \nu$), the dark blue zones by BR($B_s\to\mu^+ \mu^-$), the magenta zones by $R_{\ell 23}$, the orange zones by
BR($B\to D \tau \nu$) and the grey zones by BR($D_s\to \tau \nu$). The green area are in agreement with all the previously
mentioned constraints. The stars are points favoured by the relic density observable, in red if disfavoured by any other constraints
and in black if in agreement with all the constraints simultaneously.}
\end{figure}
In Fig. \ref{fig_amsb_omega}, we show the relic density values as a function of the AMSB parameters. The green zones
correspond to regions favoured by the flavour and direct constraints, whereas the other points are either excluded by these
constraints or by cosmological considerations (charged relic or sneutrino relic, which interact therefore strongly). Two green
zones clearly appear on the $m_0$ and $\tan\beta$ plots, for $\Omega h^2$ around $10^{-4}$ and $10^{-9}$. These areas
are far from the WMAP dark matter allowed interval, making the mAMSB scenario disfavoured by the standard cosmology.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{amsb_omega_m0.png}\\
\includegraphics[width=16cm]{amsb_omega_m32.png}\\
\includegraphics[width=16cm]{amsb_omega_tanb.png}
\end{center}
\caption{\label{fig_amsb_omega}Relic density in function of the AMSB parameters. The green points are favoured by all the
constraints, the yellow points corresponds to a charged LSP, the blue points correspond to left sneutrino LSP, and the red points
are excluded by the other constraints (flavour and direct limits).}
\end{figure}
\subsection{HCAMSB}
Another possibility to solve the negative slepton squared masses of the original AMSB scenario has been proposed: the hypercharge
anomaly mediated supersymmetry breaking (HCAMSB) scenario \cite{Dermisek:2007qi}, in which the MSSM resides on a D-brane and the hypercharge
gaugino mass is generated in a geometrically separated hidden sector \cite{Baer:2009wz}. In this way, additional contribution to
the gaugino mass $M_1$ is generated, and the large value of $M_1$ then increases the weak scale slepton masses beyond
tachyonic values, solving the generic AMSB problem \cite{Dermisek:2007qi}.
The HCAMSB scenario has four parameters:
\begin{equation}
\alpha = \frac{\tilde{M}_1}{m_{3/2}}, m_{3/2}, \tan \beta, \mbox{sgn}(\mu) \;.
\end{equation}
where $\tilde{M}_1$ is the HCAMSB contribution to $M_1$.
In order to study the parameter space, we generate mass spectra and couplings using Isajet 7.80 \cite{isajet} and
compute the flavour observables and relic density with SuperIso Relic v3.0 \cite{superiso,superiso_relic}.
We use the constraints described in the previous subsection.
In Fig. \ref{fig_hcamsb}, we scan over the whole parameter space, and project the results in two-dimensional planes. The results are somehow similar to those of the mAMSB scenario: the constraints exclude low $m_{3/2}$ values. A large part of the parameter space is favoured by flavour and direct constraints, but unfortunately no part of the parameter space respects at the same time the relic density and the other constraints, so that HCAMSB is also disfavoured by the standard cosmology.
In Fig. \ref{fig_hcamsb_omega}, we show the relic density values as a function of the HCAMSB parameters.
Again, the results are similar to those of the mAMSB scenario, with two distinct zones which are not excluded by
flavour and direct constraints, corresponding to a relic density $\Omega h^2$ around $10^{-4}$ and $10^{-8}$.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{hcamsb_m32_alpha.png}\\
\includegraphics[width=16cm]{hcamsb_alpha_tanb.png}\\
\includegraphics[width=16cm]{hcamsb_m32_tanb.png}
\end{center}
\caption{\label{fig_hcamsb}Constraints on the HCAMSB parameter space. The colour codes are the same as
in Fig. \ref{fig_amsb}.}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{hcamsb_omega_alpha.png}\\
\includegraphics[width=16cm]{hcamsb_omega_m32.png}\\
\includegraphics[width=16cm]{hcamsb_omega_tanb.png}
\end{center}
\caption{\label{fig_hcamsb_omega}Relic density in function of the HCAMSB parameters. The colour codes are the same as
in Fig. \ref{fig_amsb_omega}.}
\end{figure}
\subsection{MMAMSB}
Contrary to the two previous AMSB scenarios, the Mixed Modulus Anomaly mediated SUSY breaking (MMAMSB) scenario \cite{Choi:2005uz} provides viable dark matter candidates, in addition to solving the negative slepton mass problem naturally. This scenario is based on type-IIB
superstrings with stabilised moduli \cite{Baer:2006id}. In this scenario, an interesting result is that the soft SUSY breaking terms
receive comparable contributions from both anomaly and modulus, resulting in positive slepton masses. We examine
here the minimal MMAMSB scenario\footnote{{We note that in the general MMAMSB, there are two more
parameters, $n_{i}$ and $l_{\alpha}$, which represent respectively the modular weights of visible sector of the matter fields and
the gauge kinetic function, and which can modify the mass spectra.}}, which relies on four parameters:
\begin{equation}
\alpha , m_{3/2}, \tan \beta, \mbox{sgn}(\mu) \;.
\end{equation}
$\alpha$ here parametrises the relative contributions of modulus mediation and anomaly mediation to the soft breaking terms: the
largest $\alpha$ is, the more mediation comes from modulus \cite{Choi:2005uz}.
In Fig. \ref{fig_mmamsb}, the parameter space is scanned over, and it is projected in two-dimensional planes.
The resulting plots are different from those of mAMSB and HCAMSB. Indeed, a large part of parameter space escapes
the flavour and direct constraints, and zones around $\alpha \sim 0$ or at low $m_{3/2}$ also fulfil the relic density constraint.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{mmamsb_m32_alpha.png}\\
\includegraphics[width=16cm]{mmamsb_alpha_tanb.png}\\
\includegraphics[width=16cm]{mmamsb_m32_tanb.png}
\end{center}
\caption{\label{fig_mmamsb}Constraints on the MMAMSB parameter space. The colour codes are the same as
in Fig. \ref{fig_amsb}.}
\end{figure}
In Fig. \ref{fig_mmamsb_omega}, we present the relic density values in function of the MMAMSB parameters. The relic density of points
favoured by flavour and direct constraints takes values between $10^{-9}$ and $10^3$. We can notice however that most of the
points are in the interval $[1,10^3]$, and some points fit in the WMAP interval. For this reason, MMAMSB is a scenario which can
appear as attractive, as it fulfils simultaneously the standard cosmology and particle physics constraints.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{mmamsb_omega_alpha.png}\\
\includegraphics[width=16cm]{mmamsb_omega_m32.png}\\
\includegraphics[width=16cm]{mmamsb_omega_tanb.png}
\end{center}
\caption{\label{fig_mmamsb_omega}Relic density in function of the MMAMSB parameters. The colour codes are the same
as in Fig. \ref{fig_amsb_omega}.}
\end{figure}
\section{Anomaly mediated symmetry breaking in the NMSSM}
\label{sec:models2}
An interesting extension of the MSSM is the NMSSM, which brings a solution to the $\mu$-problem \cite{Maniatis:2009re}. It has an extended Higgs
sector involving additional Higgs bosons, modifying the relic density and flavour physics constraints.
Moreover, the couplings being modified, the NMSSM can escape direct constraints, and new parameter zones can be allowed by
the constraints used in the previous section.
We consider here the simplest version of the NMSSM, where the term $\mu \hat H_u \cdot \hat H_d$ of the MSSM superpotential
is replaced by
\begin{equation}
\lambda \hat H_u \cdot \hat H_d \hat S + \frac\kappa 3 \hat S^3 \;,
\end{equation}
in order for the superpotential to be scale invariant. The soft breaking terms
\begin{equation}
m_{H_u}^2 |H_u|^2 + m_{H_d}^2 |H_d|^2 + m_S^2 |S|^2 + \bigl(\lambda A_\lambda S H_u\cdot H_d
+ \frac13 \kappa A_\kappa S^3 + \mbox{h.c.} \bigr) \;,
\end{equation}
are {\it a priori} independent. Using the minimisation conditions for the potential, the scalar mass parameters $m_{H_{u,d}}$
can be replaced by the vacuum expectation values of the doublet $v_u$ and $v_d$, with
\begin{equation}
v_u^2 + v_d^2 = v^2 \approx (174 \mbox{ GeV})^2 \;\;, \qquad \tan\beta = \frac{v_u}{v_d} \;.
\end{equation}
The singlet field mass parameter can also be replaced by the singlet expectation value $v_s$. Expanding the singlet field $S$
around $v_s$ gives rise to an effective parameter $\mu_{eff}= \lambda v_s$.
One can also define an effective doublet mass such as
\begin{equation}
m_A^2 \equiv \frac{\lambda v_s}{\sin\beta \cos\beta} (A_\lambda + \kappa v_s) \;.
\end{equation}
Once the MSSM-like parameters (and in particular $\mu_{eff}$) have been fixed by the specification of the AMSB scenario, we
are left with four additional independent NMSSM-specific parameters:
\begin{equation}
\lambda, \kappa, A_\kappa, M_A \;.
\end{equation}
We scan over the intervals $\lambda \in [-0.7, 0.7]$ GeV, $\kappa \in [-0.7, 0.7]$, $A_\kappa \in [-2000, 2000]$ GeV and
$M_A \in [5,1000]$. We review in the following the differences between the different AMSB scenarios when applied to the
NMSSM.
\subsection{mNAMSB}
Minimal NMSSM-AMSB (mNAMSB) parameter points are generated here using NMSSMTools 2.3.4 \cite{NMSSMTools},
and flavour constraints and relic density are computed with SuperIso Relic v3.0 \cite{superiso,superiso_relic}.
In Fig. \ref{fig_namsb}, the parameter space of mNAMSB is presented. As in Fig. \ref{fig_amsb}, no point satisfy simultaneous
flavour, direct and relic density constraints. The direct limits are less constraining than in the MSSM, but the flavour constraints
are stronger and exclude a larger part of the NAMSB parameter space in comparison with AMSB.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmamsb_m32_m0.png}\\
\includegraphics[width=16cm]{nmamsb_m0_tanb.png}\\
\includegraphics[width=16cm]{nmamsb_m32_tanb.png}
\end{center}
\caption{\label{fig_namsb}Constraints on the mNAMSB parameter space. The colour codes are the same as in
Fig. \ref{fig_amsb}.}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmamsb_omega_m0.png}\\
\includegraphics[width=16cm]{nmamsb_omega_m32.png}\\
\includegraphics[width=16cm]{nmamsb_omega_tanb.png}
\end{center}
\caption{\label{fig_namsb_omega}Relic density in function of the mNAMSB parameters. The colour codes are the same as
in Fig. \ref{fig_amsb_omega}.}
\end{figure}
Fig. \ref{fig_namsb_omega} reveals more differences between the mAMSB and mNAMSB models. First, the zone not
excluded by flavour and direct constraints having a relic density around $10^{-9}$ does not exist in the NAMSB, a new
green zone appears for $m_{3/2} > 60$, and its relic density around $10^{-2}$ is much closer to the WMAP constraint.
However, as in the mAMSB scenario, the mNAMSB model is globally disfavoured by the standard cosmology.
\subsection{NHCAMSB}
We generate the NMSSM-HCAMSB (NHCAMSB) parameter points using NMSSMTools 2.3.4 \cite{NMSSMTools}.
The obtained constraints are shown in Fig. \ref{fig_nhcamsb}. Again, we see that no parameter point satisfies at the
same time the relic density constraint and the direct and flavour constraints. Similarly to the mNAMSB scenario,
NHCAMSB is less constrained by the direct mass limits, but is more excluded by the flavour constraints.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmhcamsb_m32_alpha.png}\\
\includegraphics[width=16cm]{nmhcamsb_alpha_tanb.png}\\
\includegraphics[width=16cm]{nmhcamsb_m32_tanb.png}
\end{center}
\caption{\label{fig_nhcamsb}Constraints on the NHCAMSB parameter space. The colour codes are the same as in
Fig. \ref{fig_amsb}.}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmhcamsb_omega_alpha.png}\\
\includegraphics[width=16cm]{nmhcamsb_omega_m32.png}\\
\includegraphics[width=16cm]{nmhcamsb_omega_tanb.png}
\end{center}
\caption{\label{fig_nhcamsb_omega}Relic density in function of the NHCAMSB parameters. The colour codes are the
same as in Fig. \ref{fig_amsb_omega}.}
\end{figure}
Fig. \ref{fig_nhcamsb_omega} shows the relic density in function of the different parameters. The green region that exists
in the HCAMSB scenario for a relic density around $10^{-8}$ disappears in the NHCAMSB, and a new region opens up
around $10^{-2}$. However, as the mNAMSB scenario, the NHCAMSB scenario remains also disfavoured by the standard cosmology.
\subsection{NMMAMSB}
The NMSSM-MMAMSB (NMMAMSB) scenario leads to similar results as the MMAMSB scenario. As can be seen in
Fig. \ref{fig_nmmamsb}, there exists many points satisfying all the constraints, including relic density, especially for
values of $\alpha$ near to 0.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmmmamsb_m32_alpha.png}\\
\includegraphics[width=16cm]{nmmmamsb_alpha_tanb.png}\\
\includegraphics[width=16cm]{nmmmamsb_m32_tanb.png}
\end{center}
\caption{\label{fig_nmmamsb}Constraints on the NMMAMSB parameter space. The colour codes are the same as in
Fig. \ref{fig_amsb}.}
\end{figure}
Fig. \ref{fig_nmmamsb_omega} reveals a difference, as the calculated relic density takes values between $10^{-4}$ and
$10^{3}$, which is more restrictive in comparison to the MSSM case.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmmmamsb_omega_alpha.png}\\
\includegraphics[width=16cm]{nmmmamsb_omega_m32.png}\\
\includegraphics[width=16cm]{nmmmamsb_omega_tanb.png}
\end{center}
\caption{\label{fig_nmmamsb_omega}Relic density as a function of the NMMAMSB parameters. The colour codes are the
same as in Fig. \ref{fig_amsb_omega}.}
\end{figure}
\section{AMSB and relic density in alternative cosmology}
\label{sec:cosmologies}
We have seen in the previous section that the relic density imposes severe constraints to the parameter spaces, excluding a
major part of the AMSB scenarios. However, the relic density calculation is generally based on the simplistic standard model of
cosmology. In this section, we reinterpret the previous results by considering four different alternatives to the cosmological
standard scenario. For this study, we focus our interest on six different benchmark points, which are described in Table \ref{points}
and are representative of the allowed parameter space in the different models.
The mass spectra associated to these points are shown in Fig. \ref{fig_spectra}.
\begin{table}[!t]
\hspace*{0.4cm}
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
Point & Model & $\Omega_{DM} h^2$ & $m_0$ (GeV) & $\alpha$ & $m_{3/2}$ (TeV) & $\tan\beta$ & $M_A$(GeV)\\[1mm]
\hline\hline
A & mAMSB & $3.33 \times 10^{-4}$ & 1000 & n/a & 80 & 30 & 1060.5\\[1mm]
\hline
B & mAMSB & $4.63 \times 10^{-10}$ & 2000 & n/a & 20 & 40 & 1322.8\\[1mm]
\hline
C & HCAMSB & $3.24 \times 10^{-4}$ & n/a & 0.1 & 80 & 10 & 1931.3\\[1mm]
\hline
D & MMAMSB & 5.98 & n/a & 10 & 20 & 30 & 1904.4\\[1mm]
\hline
E & MMAMSB & $6.95 \times 10^{2}$ & n/a & 20 & 100 & 10 & 2320.5\\[1mm]
\hline
F & mNAMSB & $1.21 \times 10^{-2}$ & 1300 & n/a & 70 & 20 & 770\\[1mm]
\hline
\end{tabular}
\caption{Benchmark points for testing alternative cosmology scenarios. All these points have $\mu > 0$ and are in agreement with
all the flavour and direct search constraints but would be excluded by WMAP constraints based on the standard cosmology.
For the point F extra parameters are needed to specify a point in the parameter space, which are chosen to be: $\lambda=-0.1$,
$\kappa=0.5$ and $A_\kappa= 1500$ GeV.
\label{points}}
\end{table}
\begin{figure}[p!]
\begin{center}
\hspace*{-4.8cm}Point A \hspace*{6.3cm} Point B\\
\includegraphics[width=8cm]{amsb_10m4.pdf}\includegraphics[width=8cm]{amsb_10m9.pdf}\\
\hspace*{-4.7cm}Point C \hspace*{6.5cm} Point D\\
\includegraphics[width=8cm]{hcamsb_10m4.pdf} \includegraphics[width=8cm]{mmamsb_10p1.pdf}\\
\hspace*{-4.8cm}Point E \hspace*{6.5cm} Point F\\
\includegraphics[width=8cm]{mmamsb_10p3.pdf} \includegraphics[width=8cm]{nmamsb_10m2.pdf}\\
\end{center}
\caption{\label{fig_spectra}Mass spectra of the six benchmark points. Note that the scales are not identical for all spectra.}
\end{figure}
\subsection{Alternative cosmological scenarios}
In the following, we consider that dark matter is composed of exclusively one particle produced thermally.\\
The density number of supersymmetric particles is determined by the Boltzmann equation \cite{relic_calculation}:
\begin{equation}
\frac{dn}{dt} = - 3 H n - \langle \sigma v \rangle (n^2 - n^2_{eq})\;,\label{boltzmann}
\end{equation}
where $n$ is the number density of supersymmetric particles, $\langle \sigma v \rangle$ is the thermally averaged annihilation
cross-section, $H$ is the Hubble parameter, $n_{eq}$ is the relic particle equilibrium number density. The expansion rate $H$ is
determined by the Friedmann equation:
\begin{equation}
H^2=\frac{8 \pi G}{3}\rho_{rad} \;,\label{friedmann}
\end{equation}
where $\rho$ is the total energy density of the Universe. The entropy evolution reads:
\begin{equation}
\frac{ds}{dt} = - 3 H s\label{entropy_evolution} \;,
\end{equation}
where $s$ is the total entropy density. Solving and evolving simultaneously Eqs. (\ref{boltzmann}), (\ref{friedmann}) and
(\ref{entropy_evolution}) enable to compute the relic density in our present Universe. In the standard cosmology, the dominant
component before Big-Bang Nucleosynthesis is considered to be radiation, which is constituted of all relativistic particles. This
assumption is however relaxed in many cosmological scenarios. The last two equations can indeed be written more generally as \cite{Arbey:2009gt}
\begin{eqnarray}
H^2 &= &\frac{8 \pi G}{3} (\rho_{rad} + \rho_D) \;,\label{friedmannmod}\\
\frac{ds}{dt} &=& - 3 H s + \Sigma_D \label{entropy_evolutionmod} \;,
\end{eqnarray}
$\rho_D$ parametrises a modified evolution of the total density of the Universe, beyond radiation density $\rho_{rad}$.
$\Sigma_D$ parametrises here effective entropy fluctuations due to unknown properties of the Early Universe. The radiation
energy and entropy density evolutions are known and can be written as usual:
\begin{equation}
\rho_{rad}=g_{\rm{eff}}(T) \frac{\pi^2}{30} T^4 \;, \qquad\qquad s_{rad} = h_{\rm{eff}}(T) \frac{2\pi^2}{45} T^3 \;. \label{srad}
\end{equation}
In the following, we consider two scenarios in which the energy density is modified, and two scenarios with a modified entropy
content.
\subsection{Quintessence}
The quintessence model is one of the most well-known models for dark energy. It is based on a cosmological scalar field which has
presently a negative constant pressure $P$ and a positive constant energy density $\rho$ such as $P \approx - \rho$
\cite{quintessence2}. This behaviour is achieved when the kinetic term of the scalar field equilibrates the potential. However, in the
early Universe, the scalar field has a dominating kinetic term, leading to a positive pressure such as $P \approx \rho$. During this
period, the energy density was varying very quickly, such as $\rho \propto T^6$. We study here a quintessence scenario in which
the quintessence field before Big-Bang Nucleosynthesis was dominating the expansion of the Universe. In this case \cite{Arbey:2008kv}:
\begin{equation}
\rho_D(T) = \kappa_\rho \rho_{rad}(T_{BBN}) \left(\frac{T}{T_{BBN}}\right)^6 \;,
\end{equation}
where $\kappa_\rho$ is the proportion of quintessence to radiation at the BBN temperature ($\sim$1 MeV). We consider that
$\kappa_\rho$ is a free parameter, which can be constrained using the BBN abundance constraints. To compute the abundance
of the elements produced during the primordial nucleosynthesis, we use the code AlterBBN \cite{alterbbn} integrated into SuperIso Relic. Comparing the abundances to the observational constraints, we obtain limits on $\kappa_\rho$.
\subsection{Late decaying inflaton}
The second scenario we consider here is a late decay of an inflaton field. The inflaton field is a scalar field which is considered to
be responsible for the rapid inflation of the early Universe. Generally, the inflaton is considered to decay into standard particles
much before the relic decoupling from the primordial soup. However, several models evoke the possibility of a late decay of the
inflaton, around the time of BBN. From \cite{Gelmini:2006pw, Gelmini:2006pq}, there exist cosmological models in
which the late decay of a scalar field reheats the Universe to a low reheating temperature , which can be smaller than
the freeze-out temperature, without spoiling primordial nucleosynthesis. The decay of this scalar field into radiation increases the
entropy and modifies the expansion rate of the Universe. We consider here such a scenario in which we neglect the eventual
entropy production, and we takes \cite{Arbey:2008kv}
\begin{equation}
\rho_D(T) = \kappa_\rho \rho_{rad}(T_{BBN}) \left(\frac{T}{T_{BBN}}\right)^8 \;.
\end{equation}
The exponent is here increased from 6 to 8 in comparison to the quintessence field, as mentioned also in
\cite{reheating5,reheating6}. Such a modification of the expansion rate can be also achieved in a Universe with
extra-dimensions modifying the Friedmann equations \cite{Okada:2004nc}.
\subsection{Primordial entropy production}
In this third scenario, we assume that a primordial entropy production due to an unknown component occurs. In general, such an
entropy production should be accompanied with energy production, but to better estimate the deviation of the relic density in such
a cosmological scenario, we neglect here the energy production, and we consider that the Universe has, in addition to the
standard radiation entropy density, a dark entropy density evolving like \cite{Arbey:2009gt}
\begin{equation}
s_D(T) = \kappa_s s_{rad}(T_{BBN}) \left(\frac{T}{T_{BBN}}\right)^3 \;,
\end{equation}
where $\kappa_s $ is the ratio of effective dark entropy density over radiation entropy density at the time of BBN.
The corresponding entropy production is related to $s_D$ by the relation
\begin{equation}
\Sigma_D = \sqrt{\frac{4 \pi^3 G}{5}} \sqrt{1 + \tilde{\rho}_D} T^2 \left[\sqrt{g_{eff}} s_D - \frac13 \frac{h_{eff}}{g_*^{1/2}} T \frac{ds_D}{dT}\right] \;, \label{SigmaD}
\end{equation}
with
\begin{equation}
g_*^{1/2} = \frac{h_{eff}}{\sqrt{g_{eff}}}\left(1+\frac{T}{3 h_{eff}} \frac{dh_{eff}}{dT}\right) \;,
\end{equation}
and
\begin{equation}
\tilde{\rho}_D=\frac{\rho_{D}}{\rho_{rad}} \;.
\end{equation}
\subsection{Late reheating}
In this last scenario, the reheating temperature $T_{RH}$ is smaller than the neutralino
freeze out temperature ($T_{f.o.} \simeq m_{\chi}/20$ GeV) \cite{Gelmini:2006pq}, $T_{RH}$ should be considered as a cosmological parameter that
can take any value around a few MeV and then the neutralinos decoupled from the plasma before the end of the reheating process
so their relic density will differ from its standard value. We consider that the inflaton decays around the time of Big-Bang
nucleosynthesis, generating entropy. From the standard late reheating scenarios, we assume that the entropy production evolves like \cite{Arbey:2009gt}
\begin{equation}
\Sigma_D(T) = \kappa_\Sigma \Sigma_{rad}(T_{BBN}) \left(\frac{T_{BBN}}{T}\right)
\end{equation}
for $T > 1$ MeV and that this entropy production stops at the time of BBN. We again neglect the energy production or non-thermal
production of particles in order to better understand the effects of a reheating entropy production on the relic density. In term of
entropy density, we have
\begin{equation}
s_D(T) = 3 \sqrt{\frac{5}{4 \pi^3 G}} h_{eff} T^3 \int_0^T dT' \frac{g_*^{1/2}\Sigma_D(T')}{\sqrt{1+\dfrac{\rho_D}
{\rho_{rad}}}h_{eff}^2 T'^6}\;.
\end{equation}
\subsection{BBN constraints and modified relic density}
The different scenarios do not have impact on the cosmological observations, but they can modify the abundance of the
elements. We compute with AlterBBN \cite{alterbbn} the abundance of the elements in the different scenarios for each benchmark
points, varying the single parameter for a given scenario, and we apply the following conservative constraints \cite{jedamzik06}:
\begin{eqnarray}
0.240 < Y_p < 0.258\;, \qquad 1.2 \times 10^{-5} < \!~^2\!H/H < 5.3 \times 10^{-5}\;,\qquad\qquad\label{BBNconstraints}\\
\nonumber 0.57 < \!~^3\!H/\!~^2\!H < 1.52\;, \qquad \!~^7\!Li/H > 0.85 \times 10^{-10}\;,\qquad \!~^6\!Li/\!~^7\!Li < 0.66\;,
\end{eqnarray}
for the helium abundance $Y_p$ and the primordial $^2\!H/H$, $^3\!H/\!~^2\!H$, $^7\!Li/H$ and $^6\!Li/\!~^7\!Li$ ratios.
In Fig. \ref{fig_bbn}, we consider the relic density calculated for each of the benchmark points (A-F, from top to bottom) and for different cosmological scenarios (from left to right).
\begin{figure}[p!]
\begin{center}
\includegraphics[width=4.3cm]{quintessenceA.png}\includegraphics[width=4.3cm]{inflatonA.png}\includegraphics[width=4.3cm]{entropyA.png}\includegraphics[width=4.3cm]{reheatA.png}\\
\includegraphics[width=4.3cm]{quintessenceB.png}\includegraphics[width=4.3cm]{inflatonB.png}\includegraphics[width=4.3cm]{entropyB.png}\includegraphics[width=4.3cm]{reheatB.png}\\
\includegraphics[width=4.3cm]{quintessenceC.png}\includegraphics[width=4.3cm]{inflatonC.png}\includegraphics[width=4.3cm]{entropyC.png}\includegraphics[width=4.3cm]{reheatC.png}\\
\includegraphics[width=4.3cm]{quintessenceD.png}\includegraphics[width=4.3cm]{inflatonD.png}\includegraphics[width=4.3cm]{entropyD.png}\includegraphics[width=4.3cm]{reheatD.png}\\
\includegraphics[width=4.3cm]{quintessenceE.png}\includegraphics[width=4.3cm]{inflatonE.png}\includegraphics[width=4.3cm]{entropyE.png}\includegraphics[width=4.3cm]{reheatE.png}\\
\includegraphics[width=4.3cm]{quintessenceF.png}\includegraphics[width=4.3cm]{inflatonF.png}\includegraphics[width=4.3cm]{entropyF.png}\includegraphics[width=4.3cm]{reheatF.png}
\end{center}
\caption{\label{fig_bbn}Relic density in function of the cosmological model parameters. The red lines correspond to points excluded
by the BBN constraints in Eq. (\ref{BBNconstraints}), whereas the blue lines correspond to a region with a correct abundance of the elements. The green squares correspond to points allowed by the BBN constraints and giving a relic density in agreement with the WMAP dark matter interval.}
\end{figure}%
These plots reveal that the quintessence and inflaton scenarios globally increase the relic density, while the entropy and reheating scenarios decrease it. The comparison with the BBN constraints is also represented, and the red part of the curves is excluded at 95\% C.L., while the blue part gives a correct abundance of the elements. As general features, the quintessence and inflaton scenarios can increase the relic density by three orders of magnitudes without interfering with the BBN constraints, and the entropy and reheating scenarios can decreased to a factor of $10^6$. Therefore, apart from point B which has an extremely low relic density value in the standard cosmological scenario, all the other benchmark points can have a relic density value compatible with the cosmological observations if a non minimal cosmological scenario is considered.
\section{Generalised relic density constraints}
\label{sec:constr}
We have shown with different well-known cosmological scenarios, that the relic density constraints can be very strongly relaxed. Therefore, we propose to compare the relic density calculated in the standard model of cosmology to the following interval
\begin{equation}
10^{-4} < \Omega_{DM} h^2 < 10^5
\label{newWMAP}
\end{equation}
to take into consideration the fact that it is possible to increase any relic density calculated in the standard cosmology by three
orders of magnitude, and to decrease it by six orders, with non-standard cosmological scenarios in agreement with the current cosmological data. We can re-apply the relic density constraints, and the results are shown
in Figs. \ref{fig_amsb_reloaded}-\ref{fig_nmmamsb_reloaded}. As in the figures of section \ref{sec:models}, the green zones correspond to regions in agreement with all flavour and direct constraints, but not necessarily with the relic density constraint. We added in the figures black points, which correspond to regions also in agreement with the new dark matter interval. It is clear that the allowed regions are therefore much larger than with the initial relic density interval, but a surprising result is that even with the very large interval we use here for the relic density, the relic density constraint still excludes large part of the parameter spaces. In particular, in the mAMSB and HCAMSB scenarios and their NMSSM counterparts, the relic density constraints clearly exclude the region $m_{3/2} \lesssim 40$ TeV. The MMAMSB scenario however is not constrained anymore when using the new dark matter interval.
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{amsb_m32_m0_b.png}\\
\includegraphics[width=16cm]{amsb_m0_tanb_b.png}\\
\includegraphics[width=16cm]{amsb_m32_tanb_b.png}
\end{center}
\caption{\label{fig_amsb_reloaded}Constraints on the minimal AMSB parameter space. The exclusion regions are plotted in the
order given in the legend. The red zones are excluded by the inclusive branching ratio of $B \to X_s \gamma$, the yellow ones
correspond to charged LSP, the olive green area is excluded by direct collider constraints, the light blue zones are excluded by
BR($B\to\tau \nu$), the dark blue zones by BR($B_s\to\mu^+ \mu^-$), the magenta zones by $R_{\ell 23}$, the orange zones by
BR($B\to D \tau \nu$) and the grey zones by BR($D_s\to \tau \nu$). The green areas are in agreement with all the previously
mentioned constraints. The black area corresponds to parameters in agreement with all constraints, including the revised relic
density interval given in Eq. (\ref{newWMAP}).}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{hcamsb_m32_alpha_b.png}\\
\includegraphics[width=16cm]{hcamsb_alpha_tanb_b.png}\\
\includegraphics[width=16cm]{hcamsb_m32_tanb_b.png}
\end{center}
\caption{\label{fig_hcamsb_reloaded}Constraints on the HCAMSB parameter space with the revised relic density interval given in Eq. (\ref{newWMAP}).
The colour codes are the same as in Fig. \ref{fig_amsb_reloaded}.}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{mmamsb_m32_alpha_b.png}\\
\includegraphics[width=16cm]{mmamsb_alpha_tanb_b.png}\\
\includegraphics[width=16cm]{mmamsb_m32_tanb_b.png}
\end{center}
\caption{\label{fig_mmamsb_reloaded}Constraints on the MMAMSB parameter space with the revised relic density interval given in Eq. (\ref{newWMAP}). The colour codes are the same as in Fig. \ref{fig_amsb_reloaded}.}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmamsb_m32_m0_b.png}\\
\includegraphics[width=16cm]{nmamsb_m0_tanb_b.png}\\
\includegraphics[width=16cm]{nmamsb_m32_tanb_b.png}
\end{center}
\caption{\label{fig_namsb_reloaded}Constraints on the mNAMSB parameter space with the revised relic density interval given in Eq. (\ref{newWMAP}).
The colour codes are the same as in Fig. \ref{fig_amsb_reloaded}.}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmhcamsb_m32_alpha_b.png}\\
\includegraphics[width=16cm]{nmhcamsb_alpha_tanb_b.png}\\
\includegraphics[width=16cm]{nmhcamsb_m32_tanb_b.png}
\end{center}
\caption{\label{fig_nhcamsb_reloaded}Constraints on the NHCAMSB parameter space with the revised relic density interval given in Eq. (\ref{newWMAP}).
The colour codes are the same as in Fig. \ref{fig_amsb_reloaded}.}
\end{figure}
\begin{figure}[p!]
\begin{center}
\includegraphics[width=16cm]{nmmmamsb_m32_alpha_b.png}\\
\includegraphics[width=16cm]{nmmmamsb_alpha_tanb_b.png}\\
\includegraphics[width=16cm]{nmmmamsb_m32_tanb_b.png}
\end{center}
\caption{\label{fig_nmmamsb_reloaded}Constraints on the NMMAMSB parameter space including the revised relic
density interval. The colour codes are the same as in Fig. \ref{fig_amsb_reloaded}.}
\end{figure}
\section{LHC phenomenology}
\label{sec:lhc}
The benchmark points selected for testing alternative cosmology scenarios are in agreement with precision flavour and
direct search constraints. It turns out that the phenomenology expected at the LHC is quite peculiar. The mass spectra for the
benchmark points we considered show that the lightest neutralino is the LSP and the lightest chargino
is in most cases very close in mass to the neutralino (points A, B, C, F). This will give rise to peculiar signatures due to the very
limited number of open channels for the decay modes. We analyse in more detail in the following each of the
six benchmark points previously selected. Production of charginos and neutralinos takes place at the LHC via cascade
decays of squark and gluinos and via the direct production channels
\begin{equation}
pp \to {\tilde{\chi}}_i {\tilde{\chi}}_j + X
\end{equation}
where the s-channel exchange of an off-shell W or Z or photon, and the contribution of SUSY-QCD diagrams are important.
Indeed these cross-sections receive important SUSY-QCD corrections, typical values are given in \cite{Beenakker:1999xh} or can be
obtained using a Monte Carlo program including the relevant K-factors or the detailed next-to-leading matrix elements.
\subsection{mAMSB point A}
The minimal AMSB scenario is disfavoured by the standard cosmology as very far from the WMAP dark matter allowed interval. The
region allowed in our more general analysis typically favours points in which the lightest chargino and neutralino are very close in
mass and not so heavy. For point A, $m_{{\tilde{\chi}}_1^0}=231.76$ GeV and $m_{{\tilde{\chi}}_1^+}=231.93$ GeV so that
the mass splitting is only 170 MeV. Due to this small mass splitting, the open decay modes for the ${\tilde{\chi}}_1^+$ are
(neglecting very small modes below the $10^{-3}$ level) ${\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 l \nu$, where
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 \mu^+ \nu_\mu) &\simeq& 1.87 \times 10^{-2} \\
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 e^+ \nu_e) &\simeq& 1.87 \times 10^{-2} \\
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 \pi^+ \to
{\tilde{\chi}}_1^0 e^+ \nu_e ) &\simeq& 0.96 \;,
\end{eqnarray}
so that $l=e$ and $l=\mu $ have the same branching through the three body decays while most of the signal
with ${\tilde{\chi}}_1^0 e^+ \nu_e$ is through the two body production of a charged pion.
The next lightest particle is the ${\tilde{\chi}}_2^0$, decaying mainly to ${\tilde{\chi}}_1^+$ and W or ${\tilde{\chi}}_1^0$ and SM-like Higgs
as for the light Higgs for this point (118.15 GeV) the decay is kinematically allowed :
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^\pm W^\mp ) &\simeq& 0.75 \\
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 h^0 ) &\simeq& 0.19\\
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 l^+ l^- ) &\simeq& 5.5 \times 10^{-3}\;,
\end{eqnarray}
where $l=e$, $\mu$.
Since the branching of the mode ${\tilde{\chi}}_1^\pm \to {\tilde{\chi}}_1^0 l^\pm \nu$ is 100\% and the mode
${\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 l^+ l^-$ is non-negligible one can study the clean trilepton signal usually
suggested at hadron colliders \cite{Baer:1992dc,Baer:1994nr}.
\subsection{mAMSB point B}
The benchmark point B is in the minimal AMSB scenario too. The situation is similar to the previous point with in this case
the lightest chargino and neutralino lighter than the SM-like Higgs boson. For point B, $m_{{\tilde{\chi}}_1^0}=52.70$ GeV and
$m_{{\tilde{\chi}}_1^+}=53.07$ GeV so that the mass splitting is only 63 MeV. The decay modes for the lightest chargino are
similar to the previous case
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 \mu^+ \nu_\mu) &\simeq& 5.26 \times 10^{-2} \\
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 e^+ \nu_e) &\simeq& 5.26 \times 10^{-2} \\
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 \pi^+ \to
{\tilde{\chi}}_1^0 e^+ \nu_e ) &\simeq& 0.89\;,
\end{eqnarray}
and for the second neutralino
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^\pm W^\mp ) &\simeq& 0.8 \\
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 h^0 ) &\simeq& 0.1\\
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 Z^0 ) &\simeq& 9.6 \times 10^{-2}\;.
\end{eqnarray}
The SM-Higgs boson (with a mass of 115.7 GeV) decays with a sizeable branching to pairs of lightest charginos and pairs of
lightest neutralinos (branching $1.1 \times 10^{-1}$ and $7.9 \times 10^{-2}$ respectively) while the largest mode is
$h^0 \to b{\bar b}$ with a branching of $6.5 \times 10^{-1}$.
\subsection{HCAMSB point C}
The HCAMSB scenario is disfavoured by standard cosmology and excluded by the WMAP dark matter allowed interval as
for the minimal AMSB scenario discussed above. Also in this case the lightest chargino and neutralino are very close in
mass. The masses are $m_{{\tilde{\chi}}_1^0}=229.41$ GeV and $m_{{\tilde{\chi}}_1^+}=229.58$ GeV with a splitting of only
17 MeV. In this situation the lightest chargino decays are very close to the numbers for point A. The decay modes for the
lightest chargino are :
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 \mu^+ \nu_\mu) &\simeq& 1.72 \times 10^{-2} \\
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 e^+ \nu_e) &\simeq& 1.72 \times 10^{-2} \\
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 \pi^+ \to
{\tilde{\chi}}_1^0 e^+ \nu_e ) &\simeq& 0.96\;,
\end{eqnarray}
and for the second neutralino
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^\pm W^\mp ) &\simeq& 0.67 \\
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 h^0 ) &\simeq& 7.3 \times 10^{-2}\\
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 Z^0 ) &\simeq& 0.26
\end{eqnarray}
In this case the trilepton mode $pp \to {\tilde{\chi}}_1^\pm {\tilde{\chi}}_2^0 \to 3 l + {\sla{E}}_T$ is especially interesting
as the branching fraction ${\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 l \nu$ is 100\% and the one for
${\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 Z^0$ is 26\%.
\subsection{MMAMSB point D}
The Mixed Modulus AMSB supersymmetry breaking scenario allows for viable dark matter candidates. Benchmark point D has the neutralino LSP with a mass $m_{{\tilde{\chi}}_1^0}=736$ GeV. The next-to-lightest
supersymmetric particle is the stau with a mass $m_{{\tilde{\tau}}}=860$ GeV while the lightest chargino and the second lightest neutralino are heavier and almost degenerate with a mass of 1095 and 1098 GeV respectively. The stau decays to
tau and the lightest neutralino
\begin{equation}
{\mathrm {BR}}({\tilde{\tau}} \to {\tilde{\chi}}_1^0 \tau ) = 1\;,
\end{equation}
while the lightest chargino decays mainly to
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\tau}} \nu_\tau ) &\simeq& 0.84 \\
{\mathrm {BR}}({\tilde{\chi}}_1^+ \to {\tilde{\chi}}_1^0 W ) &\simeq& 0.16\;.
\end{eqnarray}
The second lightest neutralino decays to stau and tau while as the SM-like Higgs is light enough
to be produced (124.3 GeV), the decay $ {\tilde{\chi}}_2^0\to {\tilde{\chi}}_1^0 h^0$ is also kinematically open
\begin{eqnarray}
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\tau}} \tau ) &\simeq& 0.83 \\
{\mathrm {BR}}({\tilde{\chi}}_2^0 \to {\tilde{\chi}}_1^0 h^0 ) &\simeq& 0.16\;.
\end{eqnarray}
\subsection{MMAMSB point E}
For the same model as in the previous paragraph we have also an example of a higher mass spectrum in which all supersymmetric partners are quite heavy with the neutralino LSP above 7 TeV and all other particles above 10 TeV.
\subsection{mNAMSB point F}
NMSSM with minimal AMSB scenario is one of the models considered in the previous sections. It is disfavoured by the relic
density constraints if standard cosmology is assumed. Point F is in the allowed zone for flavour, low energy and direct
search constraints and is allowed if standard cosmology constraints are relaxed as discussed in the previous sections.
From the point of view of LHC searches, this point is similar to point A. For point F, $m_{{\tilde{\chi}}_1^0}=230.03$ GeV and
$m_{{\tilde{\chi}}_1^+}=231.02$ GeV so that the mass splitting is 99 MeV.
Apart from these two particles, the next
supersymmetric particle (in terms of mass) is the second neutralino (see Fig. \ref{fig_spectra}) with a mass of 721 GeV.
\section{Conclusions and perspectives}
\label{sec:conclusion}
We have considered in details the constraints on different possible realisations of superconformal anomaly mediation breaking
mechanisms in supersymmetry. These constraints include the usual LEP, B-factories, Tevatron and LHC searches, but also
precision constraints of cosmological origin, namely the WMAP limits on the relic density of cold dark matter. We have discussed
the standard cosmological approach and also alternative cosmological scenarios which do not change the cosmological observations
but which can affect strongly the constraints on the parameter space of these supersymmetric models based on the relic
abundance of dark matter. We therefore show how the dark matter constraints can be weakened in order to avoid strong model
dependent assumptions in the choice of the cosmological model. Based on different benchmark points for AMSB models, we
performed a detailed analysis of the constraints imposed by particle data and cosmology (both standard and alternative) and
finally we gave the typical mass spectra and decay modes relevant for the LHC searches. The main lesson that can be learnt in
such an exercice is that usual bounds on the parameter space of these models are too restrictive and bear a strong hidden
cosmological model dependence in the assumption of the standard cosmological scenario. Concerning the LHC searches,
points which are excluded in the standard analysis but permitted in this more general approach, may be quite
relevant for testing not only the particle theory models themselves but also alternative cosmological scenarios at the LHC, as in
many cases relatively low mass supersymmetric particles are allowed with a peculiar spectrum where the lightest neutralino and chargino are very close in mass.
\subsection*{Acknowledgements}
We would like to thank Nazila Mahmoudi for useful discussions and comments on the manuscript.
|
\section{Introduction}
Investigations into the interactions of circumstellar discs and embedded protoplanets were first conducted by \cite{GolTre1980}. They found that the exchange of angular momentum between the two should lead to migration of the protoplanet (latterly known as Type I migration), but their work did not suggest in which direction this migration would be. The analysis of \cite{War1986} for a non-self gravitating, two-dimensional disc, showed that if the disc had a negative temperature gradient (i.e. temperature reduces with increasing heliocentric radius) then an embedded protoplanet should migrate inwards. When Hot Jupiters were discovered \citep{MayQue1995} and the difficulties with their in situ formation became apparent, this inward migration seemed to be an ideal mechanism by which to reconcile existing formation scenarios with these planets' small final orbital radii. However, the timescales of growth and migration did not at first compare favourably. Migration timescales were so short that protoplanets should plummet into their stars before they were able to grow to any considerable mass. Only by reaching masses sufficient to open disc gaps can protoplanets move from fast Type I migration to slower Type II migration, which proceeds at the disc's viscous evolution timescale. Once slowed, the planets only have to survive until the disc gas dissipates, thought to occur at a stellar age of less than 6 Myrs \citep*{HaiLadLad2001}, at which point migration due to planet-gas interactions necessarily ceases.
Since \cite{War1986} further analytical descriptions (i.e. \citealt*{War1997, TanTakWar2002}), and numerical modelling (\citealt{KorPol1993}; \citealt{NelPapMasKle2000}; \citealt{Mas2002}; \citealt*{DAnHenKle2002}; \citealt{BatLubOgiMil2003}; \citealt*{DAnKleHen2003}; \citealt*{AliMorBen2004}; \citealt*{DAnBatLub2005}; \citealt{KlaKle2006}; \citealt{DAnLub2008}; \citealt{LiLubLiLin2009}) of planet migration have continued to find fast inward rates in the Type I regime ($<$ 100 \earthmass). However, these works have generally considered locally-isothermal conditions. A large number of works, both analytical and numerical, have now been devoted to exploring the impact of more complex thermodynamics upon planet migration. The first of such works was published by \cite{MorTan2003}, who performed local calculations focussing on a low-mass planet's interaction with its parent disc. They found that radiative cooling led to a non-axisymmetric mass distribution about the planet. This resulted in an additional torque beyond the commonly considered differential Lindblad torques, and altered the migration rate relative to a purely isothermal calculation. \cite{MenGoo2004} found that a planet's migration rate was sensitive to the temperature and opacity structure of the disc through which it travelled, and under certain conditions the migration rate could be very much slower than that achieved in an isothermal model. This was closely followed by \cite{JanSas2005} who also identified a slow down in Type I migration rates upon the introduction of radiative transfer.
It was \cite{PaaMel2006} who first found cases of outward migration due to a coorbital torque in their three-dimensional radiative transfer models. They found that this coorbital torque could be prevented from saturating if the radiation diffusion timescale was shorter than the libration timescale of gas in the horseshoe region. This enabled temperature asymmetries to be maintained beyond a single libration period. They also showed that reducing the opacity (further shortening the radiation diffusion timescale) of their disc could lead to results in agreement with previous isothermal calculations. This was further described by \cite{BarMas2008}, who make clear that the radiation diffusion timescale must be greater than the period of a single horseshoe orbit to avoid returning to an isothermal like migration, and that outward migration can be related to the disc's radial entropy gradient. Many numerical models have since also found evidence of outward migration in the Type I regime (\citealt{PaaMel2008}; \citealt{PaaPap2008}; \citealt{KleCri2008}; \citealt*{KleBitKla2009}; \citealt{AylBat2010}; \citealt{YamIna2011}). Periods of outward migration can help to increase the overall migration timescale of a forming planet embedded in a disc, as is required by synthesis models to explain the population of exoplanets that has been observed \citep*{IdaLin2008, MorAliBen2009}.
More recently there has been an extension of analytical descriptions of the non-linear coorbital torque, called the horseshoe drag, to describe it in both its unsaturated and saturated states (\citealt{PaaBarCriKle2010}; \citealt*{PaaBarKle2011}; \citealt{MasCas2010}). These works give expressions for the total torque acting on a planet due to both Lindblad torques and the coorbital component. As such they can describe planet migration for a large range of scenarios, taking into account the disc's density and temperature profiles, as well as its thermal diffusivity. Outward migration is fastest in discs with steep radial temperature profiles, becoming more marginal at typical gradients such as $T \propto r^{-1}$.
This paper is intended to explore protoplanet migration in a series of discs with radial temperature gradients from $r^{0}$ to $r^{-2}$. We conduct three dimensional global disc models, using smoothed particle hydrodynamics (SPH), that include self-gravity and which use a planetary surface to allow modelling of gas flow to well within the Hill sphere and the self-consistent formation of an atmosphere. We conduct a few isothermal models to investigate whether or not the temperature profile alone has an impact on the migration rate, but otherwise our calculations all include radiative transfer using a flux-limited diffusion approximation. In Section~\ref{sec:setup}, we describe our computational method, in Section \ref{sec:calc} we explain how we obtained our results which are then presented in Section~\ref{sec:results}. Section~\ref{sec:discussion} discusses these results, whilst a summary and our conclusions are given in Section~\ref{sec:summary}.
\section{Computational Method}
\label{sec:setup}
The calculations described herein have been performed using a three-dimensional SPH code. This SPH code has its origins in a version first developed by \citeauthor{Ben1990} (\citeyear{Ben1990}; \citealt{BenCamPreBow1990}) but it has undergone substantial modification in subsequent years. Energy and entropy are conserved to timestepping accuracy by use of the variable smoothing length formalism of \cite{SprHer2002} and \cite{Mon2002} with our specific implementation being described in \cite{PriBat2007}. Gravitational forces are calculated and neighbouring particles are found using a binary tree. Radiative transfer is modelled in the two temperature (gas, $T_{\rm g}$, and radiation, $T_{\rm r}$) flux-limited diffusion approximation using the method developed by \citet*{WhiBatMon2005} and \citet{WhiBat2006}. Integration of the SPH equations is achieved using a second-order Runge-Kutta-Fehlberg integrator with particles having individual timesteps \citep{Bat1995}. The code has been parallelised by M. Bate using OpenMP.
\subsection{Equation of state and radiative transfer}
\label{sec:RT}
We present a few calculations performed using a locally-isothermal equation of state, as well as many more calculations which include radiative transfer. In locally-isothermal models the temperature of the gas in the disc remains as a function of radius throughout the calculations. For the radiation hydrodynamical calculations we use the ideal gas equation of state, $p=\rho T_{g} R_{g}/\mu$ where $R_{g}$ is the gas constant, $\rho$ is the density, $T_{g}$ is the gas temperature, and $\mu$ is the mean molecular mass. The equation of state takes into account the translational, rotational, and vibrational degrees of freedom of molecular hydrogen (assuming a 3:1 mix of ortho- and para-hydrogen; see \citealt{BolHarDurMic2007}). It also includes the dissociation of molecular hydrogen, and the ionisations of hydrogen and helium. The hydrogen and helium mass fractions are $X=0.70$ and $Y=0.28$, respectively, whilst the contribution of metals to the equation of state is neglected. More details on the implementation of the equation of state can be found in \cite{WhiBat2006}.
The radiative transfer in these calculations is performed using the flux-limited diffusion approximation, as implemented by \cite{WhiBatMon2005} and \cite{WhiBat2006}, in which work and artificial viscosity (including both bulk and shear components) increase the thermal energy of the gas. Work done on the radiation field increases the radiative energy which can be transported via flux-limited diffusion. The energy transfer between the gas and radiation fields is dependent upon their relative temperatures, the gas density, and the opacity, $\kappa$. Energy is lost from the system by the radiation field into the vacuum surrounding the disc; this is performed numerically through the use of a radiation boundary. This boundary is positioned at a height from the disc midplane at which the radiation path through the gas from this height outwards has an optical depth of $\tau_{\rm op} \approx 1$; i.e. the radiation can be expected to reach the vacuum. As a result, two layers of particles, one above and one below the disc, are made to comprise the radiation boundary. These particles are compelled to follow the initial temperature profile of the disc, causing them to function as energy sinks which impose a minimum temperature to which the disc can cool \citep[see][for a fuller description]{AylBat2010}.
The use of steep radial temperature profiles in some of our models in this work leads to very high temperatures towards the inner boundary of these circumstellar discs. This material forms an unrealistically hot annulus at the inner boundary, from which radiation diffuses out through the disc, changing its structure out to the environs of the embedded protoplanet. To remedy this situation, the disc out to 2 au is compelled to maintain its initial temperature profile throughout the simulation, which quashes any diffusion in this region. This rule was applied in all the calculations for consistency, including the shallower temperature profile calculations where it has no discernible effect.
The opacities used here are those of \cite{PolMcKChr1985} and \cite{Ale1975} (the IVa King model), with the former providing the grain opacities and the latter the gas opacities at temperatures beyond the grain sublimation point. In some calculations the grain opacity is reduced by a factor of 100 to emulate possible modification of the population due to agglomeration processes \citep[see][for further details]{AylBat2009}.
\subsection{Disc models}
We model a protoplanetary disc with radial bounds of 0.1 - 3 \rorbit \ (0.52 - 15.6 au), where \rorbit \ is the initial orbital radius of our embedded protoplanets, taking a value of 5.2 au. The disc is represented by 2 million SPH particles, a number found to deliver satisfactory resolution \citep[see][]{AylBat2010}. This leads to smoothing lengths at the disc midplane at \rorbit \ of $H/4$ (away from the planet), where $H$ is the disc scaleheight and equals 0.05 at \rorbit. Of course, considerably better resolution is obtained directly surrounding the protoplanet where the densities become much higher. The surface density of the disc goes as $\Sigma \propto r^{-1/2}$, where the value at \rorbit \ is $\Sigma_{\rm p} =$ 75 \sdensity \ to allow comparison with previous works (e.g. \citealt*{LubSeiArt1999}; \citealt{BatLubOgiMil2003}). At the centre of the disc is a fixed potential representing a 1~\solarmass \ star. We implement an inner boundary for the disc that prevents gas flow onto the star, which due to the lack of a self-limiting mechanism would otherwise artificially rapidly drain material from the disc. The outer edge of the disc is bordered by ghost particles which represent the disc beyond 15.6 au, preventing the disc from shear spreading into a vacuum \citep[for more details see][]{AylBat2010}. The initial discs were evolved in the absence of a planet until any transience resulting from settling had dissipated, which required just over 4 orbits of the disc's outer edge. Self-gravity is included in all the calculations described in this paper. The impact of including self-gravity was discussed in \cite{AylBat2010}, where for these relatively low mass discs it was found to make only a marginal difference to migration rates, slightly slowing inward migration. However, self-gravity is essential for building a self-consistent atmosphere within the Hill radius.
\begin{figure}
\centering
\includegraphics[width=1.0 \columnwidth]{figure1.pdf}
\caption{Internal energy ($u$, top panels) and temperature ($T$, bottom panels) profiles along the midplane of unperturbed discs. The dashed lines show the profiles that are given when it is $u$ that is established with the given radial dependence. Conversely, the solid lines show the profiles that are produced when it is $T$ which is compelled to follow the stated profile. For the shallow profile given in the left panels, the temperature profile that results from setting $u \propto r^{-1}$ is similar to that obtained by forcing $T \propto r^{-1}$. However, where a steeper profile is used, as in the right panels, the temperatures reached are sufficient to bring about substantial changes in the specific heat capacity of the gas (i.e. due to dissociation of molecular hydrogen, excitation of rotational and vibrational modes of molecular hydrogen) such that the proportionality of energy and temperature breaks down; we use heat capacities calculated using \protect \cite{BolHarDurMic2007}. As such in this work it is the temperature profile that is set directly ($T \propto r^{-\beta}$), for which the corresponding energy is then found.}
\label{fig:uvstemp}
\end{figure}
We perform calculations with a number of radial temperature profiles described by $T \propto r^{-\beta}$ where $\beta = $ 0, 0.5, 1, 1.5, and 2, and $T_{\rm p} \approx$ 73 K. Due to the use of steeper temperature profiles in this work, which results in higher temperatures at smaller heliocentric radii, the method used to set the disc temperatures is different to that used in \cite{AylBat2010}. In this previous work the initial temperature distribution was set assuming that the heat capacity at constant volume was independent of temperature, such that $T \propto u$, where $u$ is the gas's specific internal energy. This is a good approximation so long as the rotational and vibrational modes of $\rm H_{2}$ are not excited (i.e. $\gamma = 5/3$), which is true for $T \lesssim 100$K. It was then the distribution of $u$ that was actually prescribed to follow an $r^{-1}$ profile, such that $\beta \approx 1$; this case is shown by the dashed lines in the left panels of Fig.~\ref{fig:uvstemp}. The resulting temperature profile is similar to the desired $r^{-1}$ profile as can be seen by comparing the solid line in the lower left panel, for which $T$ is set explicitly, with the dashed line that results from setting $u$. The higher temperatures in some of the disc models considered in this work fall beyond the range in which this assumption can readily be made, as shown by the right hand panels of Fig.~\ref{fig:uvstemp}. As such, for these models we set the initial disc temperatures and boundary temperatures to follow the desired profile (the solid lines in lower panels of Fig.~\ref{fig:uvstemp}) and then the appropriate internal energy for the gas is found (solid lines in top panels of Fig.~\ref{fig:uvstemp}) using tabulated heat capacities based upon \cite{BolHarDurMic2007} which are used in our radiative transfer implementation. We note that at the planet's orbital radius the disc's temperature is such that $\gamma \approx 5/3$. Therefore we adopt this value for $\gamma$ when making comparisons with analytic models.
The viscosity within the disc is introduced in a parameterised form, developed for SPH by \cite{MonGin1983}, which conserves linear and angular momentum, and was modified to deal with high Mach number shocks by \cite{Mon1992}. It is a function of two parameters, the $\alpha_{\textsc{sph}}$-term establishes a viscosity to damp subsonic velocity oscillations that may be produced in the wake of a shock front, whilst the $\beta_{\textsc{sph}}$-term prevents particle interpenetration in supersonic shocks. This viscosity is only applied when the gas is under compression, and should ideally only act near shock fronts. We use the switch developed by \cite{MorMon1997} to try to reduce the action of artificial viscosity where the cause is not a shock. This attributes an $\alpha_{\textsc{sph}}$-value to each particle, which is modified based on the local pressure gradient, such that it increases to a maximum of 1 at a shock, and falls away rapidly to 0.1 as the particle moves away; this significantly reduces unwanted viscous dissipation in the disc.
\section{Calculations}
\label{sec:calc}
\subsection{Planet model and migration measurements}
Into the protoplanetary discs we embed a protoplanetary core, or partially formed protoplanet. In all the calculations discussed here the protoplanet is represented by a point mass which is free to move, and in all but a few cases (discussed below) the point mass is enwrapped by a surface of radius 0.03 times the Hill radius (\rhill) \citep{AylBat2009, AylBat2010}. Gas is able to pile up on this surface to build an atmosphere, and so represent gas accretion in a reasonably realistic manner. As in our previous work, a smooth start to the calculations is achieved by embedding a protoplanet of radius $R_{\rm p} =0.01 r_{\rm p}$ which then shrinks exponentially to the desired radius during the first orbit of the protoplanet.
The point mass is free to move through the disc, enabling us to measure its rate of migration. This measurement is made by using a linear fit to the last 25 orbits of orbital radius evolution, in what is typically a 50 orbit total evolution. The first 25 orbits are given up to ensure that any transient disruption to the disc caused by the sudden introduction of a planet have subsided, and to allow gas in the horseshoe region to reach a quasi-equilibrium state. For a 10~\earthmass \ planet the libration time is $\approx 50$ orbits, for a 333~\earthmass \ planet it is $\approx 8$ orbits, whilst for the a 33~\earthmass \ planet it is 29 orbits. As such, our lowest mass protoplanet mass models have completed a single libration period at the end of the calculations. In \cite{AylBat2010} we presented our results using migration timescales, $\tau$, which are calculated as the time a planet of fixed mass would take to migrate from its starting orbital radius (5.2 AU in all of the models discussed) in to the central star; $\tau = r_{\rm p}/\dot{r}$. This has no real meaning if the protoplanet is migrating outwards, so in this paper we present results using simply the migration rate, $\dot{r}$, in units of $r_{\rm p}$ per orbital period at $r_{\rm p}$ (which for $r_{\rm p} = 5.2$~au is 11.86 years).
We further refine the presentation of our results by quantifying the uncertainties that arise due to our method of measuring the migration rates. To achieve this, in addition to fitting over the entirety of the last 25 orbits, we also divide this interval into 3 smaller overlapping time domains to which individual linear fits are made. For a migration trend that is well represented by a linear fit, these 3 migration rates should all be similar to both one another and to the rate measured over the entire 25 orbits. For less linear trends the measured rates give an indication of the possible variation that could be obtained by measuring at different points in the orbital evolution. We use the maximum and minimum migration rates obtained from these shorter time domain fits to provide error bars for each migration rate that is presented in this work.
In \cite{AylBat2010} we noted a difference between the migration rates measured when modelling a protoplanet with the surface described above and an accreting point mass. To explore this further we conduct a small number of calculations in which the protoplanet is represented as a point mass which accretes gas that comes within $r_{\rm acc}$, where $r_{\rm acc} = 0.03 R_{\rm H}$ in each case. The particles representing this accreted gas are removed from the calculation and their properties are used to modify those of the point mass, ensuring conservation of mass, linear momentum, and angular momentum \citep*[see][]{BatBonPri1995}.
\subsection{Opacity to diffusivity}
\label{sec:diffusivity}
\cite{MasCas2010} developed an analytic expression to calculate the torques acting on protoplanets embedded in discs with temperature profiles described by the relation $T = T_0 (r/r_p)^{-\beta}$, such as the discs we have modelled in this work. They also include terms that describe the impact on the torques of the disc's thermal diffusivity, which in our radiative transfer calculations is effectively varied through changes in the opacity. In order that we might compare our numerical models with \citeauthor{MasCas2010}'s analytic expression, we calculate the thermal diffusivity in our disc at \rorbit \ in the absence of a protoplanet. In the flux-limited diffusion approximation of radiative transfer used in this work the flux in optically thick regions is given by
\begin{equation}
\mbox{\boldmath $F$} = - \frac{c}{3 \chi \rho} \nabla E,
\label{eq:fluxeq}
\end{equation}
\noindent where $c$ is the speed of light, $\chi$ is the Rosseland mean opacity, $\rho$ is the gas density, and $E = a T_r^{4}$ in which $a$ is the radiation density constant, $a = 4 \sigma_{\textsc {sb}}/c$. This allows us to restate equation \ref{eq:fluxeq} as
\begin{equation}
\mbox{\boldmath $F$} = - \frac{16 \sigma_{\textsc {sb}} T_r^3}{3 \chi \rho} \nabla T_r.
\label{eq:fluxtemp}
\end{equation}
\noindent \citeauthor{MasCas2010} give an expression for the heat flux in terms of a form of thermal diffusivity, $k$, as
\begin{equation}
\mbox{\boldmath $F$} = -k \nabla T.
\label{eq:massetflux}
\end{equation}
\noindent Comparing equations \ref{eq:fluxtemp} and \ref{eq:massetflux} we obtain an expression for $k$, which is related to a true thermal diffusivity, $\kappa$, by
\begin{equation}
\kappa = \frac{\gamma -1}{\gamma} \frac{k \mu}{R_g \rho},
\label{eq:diffusivityconvert}
\end{equation}
\noindent which is similar to equation 33 in \cite{MasCas2010}, where $\gamma$ is the adiabatic index. The difference between this equation and that of \citeauthor{MasCas2010} lies in their use of $\Sigma$ in the denominator of equation~\ref{eq:diffusivityconvert} as they are describing a two-dimensional disc, whilst for three-dimensions we use $\rho$. Making use of the approximation that $\rho = \Sigma/H$ gives our final equation
\begin{equation}
\kappa = \frac{\gamma -1}{\gamma} \frac{16 \sigma_{\textsc {sb}} T_r^3 \mu H}{3 \chi \rho R \Sigma},
\label{eq:diffusivity}
\end{equation}
\noindent where $H$ has a value of 0.05 at \rorbit \ regardless of the chosen disc temperature profile. With this equation, taking $\gamma = 5/3$ (which is appropriate at \rorbit), and values from our unperturbed initial disc we are able to calculate the thermal diffusivities in models of differing opacity, $\chi$. For the 100\% interstellar grain opacity disc this gives $\kappa = 6.57 \times 10^{15} {\rm cm^{2}s^{-1}}$, and the 1\% opacity disc has a value of $\kappa = 6.57 \times 10^{17} {\rm cm^{2}s^{-1}}$. These diffusivities correspond with characteristic disc cooling timescales of tens of orbital periods and less than an orbital period, respectively, where this timescale is linearly dependent upon the chosen diffusivity. These timescales indicate that in the higher diffusivity (lower opacity) calculations the imposed boundary will dominate the temperature structure of the disc, but this is not the case in the lower diffusivity (higher opacity) models.
It may also be of note that in these three-dimensional models, the dominant direction for radiative diffusion within the horseshoe region is vertically. For the high opacity calculations, vertical diffusion is found to be more than three times as efficient as radial diffusion.
\section{Results}
\label{sec:results}
\subsection{Is the surface treatment important?}
First we address a comment made in \cite{AylBat2010} in which we suggested that the modelling of migrating protoplanets using a surface treatment rather than an accreting point mass might significantly alter their migration rates, even changing the direction. This was inspired by a single result for a 10~\earthmass \ protoplanet in a disc with a temperature profile exponent of $\beta \approx 1$, in which the migration was seen to change from outward to inward when a crude evacuating sink particle was replaced with a protoplanet with a surface in a high opacity disc (shown in the top panel of Fig.~\ref{fig:nosurface}). To explore this further we have conducted a series of calculations spanning $10 - 100$~\earthmass \ in a disc with a steeper temperature profile, $\beta = 2$. It was envisaged that such a steep profile would lead to outward migration for a low mass protoplanet modelled with either of our treatments by increasing the significance of the coorbital torques, making their effects more apparent. The results of these calculations are shown in the lower panel of Fig.~\ref{fig:nosurface}. It can be seen that using 1\% interstellar grain opacities (plus symbols), even with this steep temperature profile, the protoplanet migration is predominantly inwards, with the 10~\earthmass \ case being the only exception in exhibiting an almost stalled outwards migration. However, at 100\% interstellar grain opacities those protoplanets with masses $\lesssim 33$~\earthmass \ are all migrating outwards, regardless of the chosen protoplanet treatment. Comparing the migration of these low mass protoplanets with those seen in the top panel of Fig.~\ref{fig:nosurface} illustrates the increased importance of the coorbital torques in the disc with a steeper temperature gradient.
\begin{figure}
\centering
\includegraphics[width=0.95 \columnwidth]{figure2.pdf}
\caption{The upper panel shows migration rates, which were presented as migration timescales in \protect \cite{AylBat2010}, for protoplanets modelled using surfaces (symbols encircled, red) embedded in discs with $\beta \approx 1$. These discs were of either 100\% (asterisks) or 1\% (plus symbols) interstellar grain opacity. For a 333~\earthmass \ protoplanet changing the opacity has such a small impact upon migration that the two rates appear indiscernibly overlaid above. The lower panel shows migration rates for protoplanets embedded in a circumstellar disc for which $\beta = 2$, some modelled with evacuating sinks (un-encircled symbols, black) and others using sinks with surfaces (once again, encircled symbols, red). All subsequent figures include only protoplanets modelled using sinks with surfaces.}
\label{fig:nosurface}
\end{figure}
In our new calculations (bottom panel) the 10~\earthmass \ and 20~\earthmass \ protoplanet migration rates obtained with an evacuating sink (un-encircled, black) are more rapidly outwards or less rapidly inwards than their counterparts modelled using sinks with surfaces (encircled, red); the reason for this is unclear, though it is in agreement with the sole case shown in the top panel for our previous calculations. However, at masses $\gtrsim 33$~\earthmass \ this pattern reverses regardless of the disc opacity. \cite{BatLubOgiMil2003} showed that it was for such masses that a protoplanet might begin to form a non-negligible gap in the disc. The evacuation of a gap reduces the mass in the coorbital region and as a result reduces the magnitude of coorbital torques. In \cite{AylBat2010} we found that between the two treatments of the protoplanet used here, it was the evacuating sink particles that developed clearer gaps as a result of drawing in material artificially fast relative to protoplanets modelled with surfaces. This suggests that protoplanets of $\gtrsim 33$~\earthmass \ modelled with evacuating sinks migrate faster inwards (slower outwards) than there equivalents modelled with surfaces because they reduce the coorbital torques, which act outwards, to a greater extent. For a 100~\earthmass \ protoplanet the coorbital region is sufficiently depleted with both treatments to lead to inward migration in a high opacity disc.
As to our findings in \cite{AylBat2010} that changing the protoplanet treatment can reverse the direction of migration for a low mass protoplanet, we must clarify that this appears to be a result of the marginal nature of the migration of a 10~\earthmass \ protoplanet in a high opacity disc with $\beta \approx 1$. The change in the forces acting when comparing a protoplanet modelled using an evacuating sink and one modelled using a surface, happened to be sufficient in this case to change the direction of migration, though more typically it is likely to give a different rate with the same direction (as seen in Fig.~\ref{fig:nosurface}). The essential point is that as a protoplanet's migration rate is found to be at least somewhat dependent upon the method by which its accretion is modelled, it is best that the adopted treatment be as realistic as possible. We therefore remain convinced that using a surface to allow gas-pile-up as a model of accretion is the best method for these purposes.
\begin{figure*}
\centering
\includegraphics[width=1.0 \columnwidth]{figure3a.pdf}
\includegraphics[width=1.0 \columnwidth]{figure3b.pdf}
\includegraphics[width=1.0 \columnwidth]{figure3c.pdf}
\includegraphics[width=1.0 \columnwidth]{figure3d.pdf}
\caption{Migration rates of four different mass protoplanets; across and down the panels 10, 20, 33, and 50 \earthmass \ (as marked). Each protoplanet was modelled in a range of circumstellar discs with different temperature profiles, $T \propto r^{-\beta}$, where the temperature profile exponent $\beta$ is marked on the x-axis. Models in which interstellar grain opacities are used (asterisks) lead to slower or reversed migration when compared with the 1\% interstellar grain opacity models (plus symbols). \protect \cite{MasCas2010} give an expression for the total torque acting upon a protoplanet which includes the dependency upon the disc's temperature profile and its thermal diffusivity, as well as the planet's mass. The resulting migration rates are plotted in each panel, one for a diffusivity akin to our 100\% opacity calculation (solid line), and one for the 1\% opacity case (dot-dash line). Migration rates obtained from \protect \cite{TanTakWar2002} are independent of the temperature profile, and so appear in each panel as a horizontal dashed line, the rate of which is determined by the disc surface density and the protoplanet mass.}
\label{fig:profiles}
\end{figure*}
\subsection{Impacts of the temperature profile}
Our primary goal in this work was to explore the impact of the temperature profile in a circumstellar disc upon the migration of protoplanets therein. In particular we wished to test the predictions of recent analytic models, which by their nature are unable to include details of the gas flow near the protoplanet; Fig.~\ref{fig:profiles} contains the results of our simulations to this end. Shown are the migration rates for protoplanets with masses 10, 22, 33, and 50~\earthmass, embedded in discs with both high and low opacities, and a range of values of $\beta$. A horizontal dashed line in each panel marks the migration rate obtained using \cite{TanTakWar2002}'s analytic form, which depends on the disc surface density and protoplanet mass, but is independent of the value of $\beta$. It can be seen that the low opacity models (1\% interstellar grain opacity, marked with plus symbols) display some scatter, but generally centre around this analytic description. All of the low opacity calculations lead to inward migration, as would be expected for locally-isothermal calculations.
Also marked in Fig.~\ref{fig:profiles} are analytic migration rates based on recent work by \cite{MasCas2010}. In their description there is a dependence on both $\beta$ and on the thermal diffusivity of the disc. As such we plot lines to correspond with both opacities used in calculating our results, where the diffusivities used are those calculated in Section \ref{sec:diffusivity}; note that high opacities correspond to low thermal diffusivities and vice versa. In the case of the high opacity calculations (100\% interstellar grain opacity, marked with asterisks) it can be seen that as the steepness of the temperature profile is increased, the inward migration rates slow towards zero before the direction changes to outwards and the rates begin to increase. In each panel of Fig.~\ref{fig:profiles}, excepting the 10~\earthmass \ case for which the uncertainties are greatest, it is apparent that the rate of this change is in good agreement with the gradient given by the low thermal diffusivity (high opacity) result of \citeauthor{MasCas2010}'s analytic description (solid line). The analytic form also gives reasonable agreement (given the coarse spacing of $\beta$ in our results) as to the temperature profile at which a given mass protoplanet in a high opacity disc will transition from inwards to outwards migration. The formula of \citeauthor{MasCas2010} is also dependent upon the disc's viscosity, which for plotting here we have used a \cite{ShaSun1973} $\alpha$-value of 0.004, which is equivalent to the magnitude of the SPH viscosity in the vicinity of the protoplanet in our models. Note also that this analytic description is valid only up to masses for which the half-width of the horseshoe region, $x_s$, is less than the disc thickness at \rorbit. Adopting the adiabatic approximation of \cite{BarMas2008} for $x_s$ gives a maximum mass of 40~\earthmass, meaning that the 50~\earthmass \ cases shown in Fig.~\ref{fig:profiles} are slightly beyond the range of validity.
Using \citeauthor{MasCas2010}'s analytic form with values that correspond to our low opacity models indicates a reversal of the migration trend with increasing $\beta$, resulting in migration accelerating inwards for steeper profiles. However, the scatter in migration rates measured from our low opacity calculations is large, and prohibits us from concluding anything other that that the magnitudes show reasonable agreement. True locally-isothermal models were also conducted to establish whether they revealed a trend with the disc temperature profile. The results of these models are shown in Fig.~\ref{fig:isothermal}, where once again there appears to be no definite trend in the change of migration rates with increasing $\beta$, and the variation in the rates is small, varying by less than a factor of 2 from the minimum to maximum rate. As such we are unable to say whether or not the migration trend at low opacities given by \cite{MasCas2010} is real.
\begin{figure}
\centering
\includegraphics[width=1.0 \columnwidth]{figure4.pdf}
\caption{Migration rates for a 33~\earthmass \ protoplanet in true locally-isothermal discs with varying radial temperature profiles. There is a small range of migration rates, varying by less than a factor of two from the minimum to the maximum rate. No significant trend is evident.}
\label{fig:isothermal}
\end{figure}
Considering the immutable properties of our disc model, we find that only protoplanets $< 100$~\earthmass \ may migrate outwards for high values of $\beta$. This can be seen in Fig.~\ref{fig:100m} where migration rates for a 100~\earthmass \ protoplanet are plotted, and it is clear that no combination of opacity and $\beta$ is able to cause outward migration. Included in the figure are the same analytic descriptions used in Fig.~\ref{fig:profiles}, but here the \cite{MasCas2010} line is well beyond the mass regime for which it is valid. For such a high mass protoplanet, the evacuation of the coorbital region should effectively end the significant influence of the coorbital torques. As such these effects are overestimated by the analytic model, which is consistent with the fact that our high opacity results all lie well below the solid line. Noticeably the analytic model, even with the overestimation of the coorbital torques which should favour outward migration, does not predict such migration for this high mass protoplanet (for $\beta \lesssim 2.5$) , suggesting it is not possible in discs such as those modelled here.
\begin{figure}
\centering
\includegraphics[width=1.0 \columnwidth]{figure5.pdf}
\caption{Migration rates of a 100~\earthmass \ protoplanet in a range of circumstellar discs with different temperature profiles, $T \propto r^{-\beta}$, where the temperature profile exponent $\beta$ is marked on the x-axis. Models in which interstellar grain opacities are used (asterisks) generally lead to very slightly slower inward migration than their equivalent 1\% interstellar grain opacity (plus symbols) models. However, the trend with $\beta$, which was quite pronounced for the high opacity models of lower mass protoplanets, is more subdued in this case. It also fails to follow the trend predicted using \protect \cite{MasCas2010} for a high opacity disc (solid line), though at 100~\earthmass \ we are applying the analytic model well beyond its range of validity; the analytic trend for a low opacity disc is marked with a dot-dashed line. The change in disc opacity leads to a smaller change in the protoplanet migration rate for this relatively high mass protoplanet when compared with the lower mass protoplanets shown in Fig.~\ref{fig:profiles}. Rates from \protect \cite{TanTakWar2002} are shown as a horizontal dashed line.}
\label{fig:100m}
\end{figure}
\subsection{Impact of absolute temperature}
In this paper, we have studied the impact of the disc's temperature profile on migration rates. In all these calculations, the absolute disc temperature at the planet's orbital radius is the same. As discussed above we find reasonable agreement with the analytic description of \citeauthor{MasCas2010}. However, using \citeauthor{MasCas2010}'s analytic description it is possible to examine the migration rate's dependence upon the absolute temperature at $r_{\rm p}$; this is presented in Fig.~\ref{fig:abstemp}. The migration rates, calculated assuming a disc with a diffusivity of $\kappa = 6.57 \times 10^{15} {\rm cm^{2}s^{-1}}$ (equivalent to our 100\% opacity cases), are shown in the left hand panels. For each of the four protoplanet masses plotted ($m_{\rm p} = $ 10, 20, 33, 50~\earthmass) there is a temperature (scaleheight) at which the outward migration rate achieves a maximum, with this temperature scaling as $m_{\rm p}^{0.85}$ (within the mass range for which the model is valid, $\lesssim 40$~\earthmass). The right hand panels of Fig.~\ref{fig:abstemp} show the same cases for a higher diffusivity of $\kappa = 6.57 \times 10^{17} {\rm cm^{2}s^{-1}}$ (equivalent to our 1\% opacity cases), for which conditions there is no outward migration, regardless of the chosen absolute temperature of the disc, and there is no turnover in the rate of migration, though these rates still vary significantly with temperature.
These analytic results imply that as well as the temperature profile, the chosen initial disc temperature is an important factor in determining the migration behaviour of the various low mass planets. Temperatures in different discs around various types of star may be expected to exhibit temperatures throughout the range shown here of 10-230K at 5.2~au, suggesting further modelling should be carried out for a range of disc conditions to explore the utility of analytic descriptions like \cite{MasCas2010}, which if valid, can be applied more generally than the results of numerical modelling.
\begin{figure}
\centering
\includegraphics[width=1.0 \columnwidth]{figure6.pdf}
\caption{Migration rates of protoplanets in discs of differing absolute temperature, calculated using the analytic description of \protect \cite{MasCas2010}. Rates calculated for conditions equivalent to our 100\% opacity models are shown on the left hand side, and 1\% opacity on the right. The temperature profiles are $T \propto r^{-1}$ and $r^{-2}$ in the top and bottom rows respectively. The four lines in each panel from left to right correspond to protoplanets with masses 10, 20, 33, and 50~\earthmass, respectively. The vertical dashed line marks the temperature at $r_{\rm p}$ used in our numerical models.}
\label{fig:abstemp}
\end{figure}
\section{Discussion}
\label{sec:discussion}
First we address the low opacity calculations that we have performed. The resulting migration rates from these calculations appear scattered, showing no evident trend with changes in the disc temperature profile. This can be seen in Fig.~\ref{fig:lubow} which gathers all the low opacity results shown in Fig.~\ref{fig:profiles} into a single panel, as well as including the analytic rates for an isothermal disc calculated using \cite{TanTakWar2002} (dashed lines) and rates based upon the recent work of \cite{DAnLub2010} (solid lines) for a locally-isothermal disc. At these low opacities we expect the migration results from our models to be similar to those predicted with isothermal analytic expressions as we found in \cite{AylBat2010}. Indeed, despite the scatter, these migration results do reveal a trend with increasing mass that is matched by \cite{TanTakWar2002}'s values. However, \citeauthor{DAnLub2010} have found that in the locally-isothermal (rather than globally isothermal) regime protoplanet migration rates depend upon the parent disc's temperature profile, with steeper negative temperature profiles (higher values of $\beta$) leading to more rapid inward migration. This trend is not seen in our low opacity calculation results within which, as mentioned previously, it is not possible to establish a trend that is distinguishable from the numerical uncertainties; this is also the case in the locally-isothermal calculations shown in Fig.~\ref{fig:isothermal}. However, the variation in migration rates obtained from \cite{DAnLub2010} across the range of $\beta$ considered in this work is relatively small, less than a factor of 2 for each of the masses shown in Fig.~\ref{fig:lubow}. Our rather large uncertainties prevent us from definitively ruling out the model.
\begin{figure}
\centering
\includegraphics[width=1.0 \columnwidth]{figure7.pdf}
\caption{Migration rates of four different mass protoplanets through low opacity discs of differing temperature profiles. The masses are 10~\earthmass \ (plus symbols), 20~\earthmass \ (asterisks), 33~\earthmass \ (diamonds), and 50~\earthmass \ (triangles). Also included are analytic predictions. Those of \protect \cite{TanTakWar2002} without a $\beta$ dependence, for a three-dimensional isothermal disc, appear as horizontal dashed lines. The solid lines show the rates predicted using the $\beta$ dependent expression devised by \protect \cite{DAnLub2010} for a three-dimensional locally-isothermal disc. The masses that these lines correspond to are marked on the plot above the corresponding pairs.}
\label{fig:lubow}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=1.0 \columnwidth]{figure8.pdf}
\caption{Migration rates based on the total torques calculated using analytic expressions for a 33~\earthmass \ protoplanet in a disc with properties equivalent to our models. The dashed line indicates the expected locally-isothermal result for such a protoplanet in a three-dimensional disc, obtained using \protect \cite{DAnLub2010}. The solid lines are for rates obtained using \protect \cite{MasCas2010}, where the colours progressing through the rainbow correspond to differing diffusivities, with the highest shown in red (labelled C). Three labelled thicker lines identify the trend with diffusivity, with the corresponding values given in the key. The locally-isothermal and highest diffusivity cases share the same gradient, with magnitudes differing by $\approx 5/3$. }
\label{fig:lubmas}
\end{figure}
Returning to \cite{DAnLub2010}'s analytic form for an isothermal disc, it is interesting to note that the acceleration of inward migration with increasing $\beta$ is in qualitative agreement with the description of \cite{MasCas2010} discs with high thermal diffusivity. As a brief aside we compare these two analytic models with which we draw comparisons. Fig.~\ref{fig:lubmas} illustrates the migration rates calculated using \citeauthor{DAnLub2010}'s isothermal formula for a 33~\earthmass \ protoplanet (dashed line), with the rates obtained using \citeauthor{MasCas2010}'s formula for a whole range of thermal diffusivities also shown. It can be seen that at the highest diffusivity (thick red line labelled C) the gradient of the change of migration rate with increasing $\beta$ for the two descriptions is extremely similar, and they are displaced in magnitude by a factor of~$\approx 5/3$, equivalent to $\gamma$, as was found in \cite{PaaPap2008}. This similar behaviour is to be expected, with a high thermal diffusivity leading to a disc that resembles a locally-isothermal model, and responds to different temperature profiles in the same manner. As seen earlier when considering our low opacity calculations, the manner in which the migration rate of a protoplanet changes with the discs thermal diffusivity is dependent upon the temperature profile of the disc. For discs with $\beta \gtrsim 0.5$, as the thermal diffusivity is increased from its lowest value (thick purple line labelled A) the migration rate of a protoplanet can be expected to initially increase in the outward direction, or slow down if the migration is inwards. Beyond a certain diffusivity, shown by the thick blue line labelled B in Fig.~\ref{fig:lubmas}, the trend is reversed with increasing diffusivity slowing outward migration, and accelerating inward migration. This behaviour is a result of the edge term of the Horsehoe drag in \citeauthor{MasCas2010}'s model. In future it would be desirable to explore a broader opacity space with greater refinement using our numerical models to determine whether we can capture the effects of the edge terms of the Horseshoe drag.
In our high opacity (low thermal diffusivity) models we find that increasing $\beta$ from 0 leads first to slower inward migration and then accelerating outward migration for protoplanets of $\lesssim 50$~\earthmass. This is in agreement with the analytic work discussed above. Similar outward migration has been found in the numerical models of \cite{PaaPap2008}, \cite{BarMas2008}, \cite{KleBitKla2009}, and \cite{PaaPap2009}, and has been associated with a density enhancement ahead of the planet and a deficit behind it in the horsehoe region. In each of these models the feature has been clearly identified in two-dimensional modelling, but such a feature is likely to be much less apparent in three-dimensional models where gas is not constrained to move in a single plane. The lefthand panels of Fig.~\ref{fig:cylindrical} present plots of $\Sigma r^{1/2}$ averaged over 25 orbits for a disc of $\beta = 2$ containing a 33~\earthmass \ protoplanet. The leftmost panel is a 100\% opacity calculation in which the protoplanet exhibits outward migration, whilst the neighbouring panel is for a 1\% opacity calculation that leads to inward migration. Relative to the low opacity case there is a more substantial deficit of material trailing the protoplanet (negative $\phi$) in the horseshoe region of the high opacity case. The rightmost panel of Fig.~\ref{fig:cylindrical} shows a ratio of the two surface density perturbation plots, and makes it clear that there is a greater density contrast between the leading and trailing positions in the horseshoe region of the high opacity outwardly migrating case. This density structure is reminiscent of that found in the two-dimensional models, though much less clearly defined.
\begin{figure*}
\centering
\includegraphics[width=2.0 \columnwidth]{figure9.pdf}
\caption{The two left panels show the surface density normalised by the density gradient of the initial unperturbed disc. Respectively the panels show the 100\% and 1\% opacity models with $\beta = 2$ and a protoplanet mass of 33~\earthmass. The right hand panel shows the ratio of the two left panels, revealing how the structures differ between the optically thick and thin models. This makes it clear that relative to the low opacity (inward migrating) model the high opacity (outward migrating) model has a greater contrast between leading (positive $\phi$) and trailing (negative $\phi$) densities, thus explaining the differing directions of migration.}
\label{fig:cylindrical}
\end{figure*}
\section{Summary}
\label{sec:summary}
We have conducted a series of calculations to explore the impact of disc radial temperature profiles on the rates of protoplanet migration. The surface density profile remains unchanged in each calculation, thus each temperature profile gives rise to a different entropy profile through the disc. In agreement with analytic predictions, we find that steeper temperature profiles, $\beta >1$, (and so entropy gradients) are much more disposed to bring about outwards migration for low mass ($\lesssim 50$~\earthmass) protoplanets. There remains a dependence on the disc's opacity, which determines how the disc and the embedded protoplanet are able to cool, which can change the discs structure and the resulting torques. In interstellar grain opacity (low thermal diffusivity) discs, low mass protoplanets are found to migrate outwards, whilst in discs with opacities of 1\% this level we find no cases of outward migration. The dependence of migration rates upon the disc's temperature profile in high opacity discs is in agreement with the predictions of recent analytic work by \cite{MasCas2010}.
For the low opacity calculations we find migration rates broadly inline with the rates predicted by \cite{TanTakWar2002}. These conditions can be represented using the formulae of \cite{MasCas2010} with a suitably high thermal diffusivity. Such an approach suggests that a given protoplanet should migrate more rapidly inwards in a disc with a more steeply negative temperature profile. We found this to qualitatively agree with recent work by \cite{DAnLub2010} who find a similar trend for a locally-isothermal disc. However, our efforts to numerically test these predictions were hampered by large numerical uncertainties in the results of our low opacity calculations, leaving us unable to determine the veracity of the analytic predictions.
We also expanded our investigation into the importance of the model used to represent the protoplanet. Using a surface that enables accreted gas to pile up and form an atmosphere results in somewhat different migration rates to those obtained when using a more crude accreting sink particle model. For the lowest-mass protoplanets ($\lesssim 20$~\earthmass) the change in migration rate with protoplanet treatment was found to be small, though for borderline cases between inward and outward migration it could be sufficient to change the direction. Those protoplanets of $\gtrsim 33$~\earthmass \ using the less realistic accreting sink model are found to migrate more rapidly inwards as they more effectively clear the coorbital region of material than equivalent surface models, more rapidly diminishing the positive coorbital torques. In general the effect of changing the protoplanet treatment is small compared to the differences brought about by altering the disc properties.
\section*{Acknowledgments}
We thank the anonymous referee for prompting some further useful investigation. The calculations reported here were performed using the University of Exeter's SGI Altix ICE 8200 supercomputer. Much of the analysis was conducted making use of SPLASH \citep{Pri2007}, a visualisation tool for SPH that is publicly available at http://www.astro.ex.ac.uk/people/dprice/splash. MRB is grateful for the support of a Philip Leverhulme Prize and a EURYI Award which also supported BAA. This work, conducted as part of the award ``The formation of stars and planets: Radiation hydrodynamical and magnetohydrodynamical simulations" made under the European Heads of Research Councils and European Science Foundation EURYI (European Young Investigator) Awards scheme, was supported by funds from the Participating Organisations of EURYI and the EC Sixth Framework Programme.
|
\section{Introduction}
Let $E$ be an $\ell$-dimensional Euclidean space and $\Phi$ be the root system
of the type $\mathit{A}_{\ell}$.
Let $\Phi_{+}$ denote the set of positive roots.
In this paper we explicitly choose $E$ and $\Phi$ as follows:
let
$W = \R^{\ell+1}$
and
$x_{1}, \dots, x_{\ell+1}$
be an orthonormal basis for the dual space $W^{*}$.
Define
\begin{align*}
E &:= \{\sum_{i=1}^{\ell+1} c_{i} x_{i}
\in W^{*} \mid \sum_{i=1}^{\ell+1} c_{i} = 0\},\\
\Phi &:= \{x_{i} - x_{j} \in E
\mid 1\leq i\leq \ell+1, 1\leq j\leq \ell+1,
i\neq j\},
\\
\Phi_{+} &:=
\{
x_{i} - x_{j} \in \Phi \mid i < j
\}.
\end{align*}
Let
$\mathcal{A}(\Phi)=\{ \mathit{H}_{\alpha} \mid \alpha \in \Phi_{+} \}$,
where $\mathit{H}_{\alpha}=\{ v \in \mathit{W} \mid \alpha(v)=0 \}$.
Then
$\A(\Phi)$
is called
a {\bf
braid arrangement},
which is undoubtedly
the most-studied arrangement of hyperplanes
in various contexts.
In the study of the Kazhdan-Lusztig
representation theory of the affine Weyl groups,
J.-Y. Shi introduced
the \textbf{Shi\ arrangements}
in \cite{shi1}
as follows:
Let
\[
\mathit{H}_{\alpha,1} = \{ v \in \mathit{W} \mid \alpha(v)=1 \}.
\]
Then the Shi arrangement is given by
\[
\mathcal{A}(\Phi) \cup \{ \mathit{H}_{\alpha,1} \mid \alpha \in \Phi_{+} \} = \bigcup_{\alpha \in \Phi_{+}} \{ \mathit{H}_{\alpha},
\mathit{H}_{\alpha,1} \}.
\]
Embed the $(\ell+1)$-dimensional space $W$ into $V=\R^{\ell+2}
$ by
adding a new coordinate $z$ such that $W$ is defined by
the equation $z=1$ in $V$.
Then, as in \cite[Definition 1.15]{OT},
we have the {\bf cone} $\Shi_{\ell} $ over the Shi arrangement.
It is a central arrangement in $V$
defined by
$$Q(\Shi_{\ell} ) = z \prod_{1\leq p<q\leq \ell+1} (x_{p} - x_{q})
\prod_{1\leq p<q\leq \ell+1} (x_{p} - x_{q}-z)=0.$$
Let $\mathit{S}$ be the algebra of polynomial functions on $\mathit{V}$ and let $\mathrm{Der}_{\mathit{V}}$ be the module of derivations
of $\mathit{S}$ to itself
\[
\mathrm{Der}_{\mathit{V}}=\{ \theta : \mathit{S} \rightarrow \mathit{S} \mid \theta\ \text{is}\ \mathbb{R} \text{-linear and}\ \theta (fg) =f\theta(g)+ g\theta(f)\ \text{for any}\ f,g \in \mathit{S} \}.
\]
The derivation module $\mathit{D}(\mathcal{S}_{\ell})$ is defined by
\begin{multline*}
\mathit{D}(\mathcal{S}_{\ell})=\{ \theta \in \mathrm{Der}_{\mathit{V}} \mid
\theta(\alpha)\ \text{is divisible by}\ \alpha\\
\text{ and }\
\theta(\alpha-z)\ \text{is divisible by}\ \alpha-z
\text{ for any}\ \alpha \in \Phi_{+} \}.
\end{multline*}
In the present paper, for the first time,
we construct an explicit basis for the derivation module
$\mathit{D}(\mathcal{S}_{\ell})$.
The most important ingredients of our recipe are
the \textbf{Bernoulli polynomials} $\mathit{B}_{k}(x)$
and their relatives $B_{p, q}(x)$ in Definition \ref{Bpqdefinition}.
The explicit costruction of our basis is in Definition \ref{basisdefinition}.
One of the remarkable properties of the Shi arrangement is the fact that its number of chambers is equal to
$(\ell+2)^{\ell}$.
A good number of articles,
including \cite{St96, ER, Hea, Ath, Y04},
study this intriguing property.
Because of Zaslavsky's chamber counting
formula \cite{Z75},
the property follows from the formula
\[\pi(\mathcal{S}_{\ell}, t)= (1+t)\left(1+(\ell +1)t\right)^{\ell} \]
for the Poincar\'e polynomial
\cite{OT} of the cone $\mathcal{S}_{\ell}$.
Ch. Athanasiadis proved that
$\mathit{D}(\mathcal{S}_{\ell})$ is a free $\mathit{S}$-module
with exponents
$(0, 1, \ell+1,\ldots ,\ell+1)$ in \cite{Ath}.
He consequently proved the formula
above
thanks to
the factorization theorem in
\cite{T81}
which
asserts
that
if the derivation module
$\mathit{D}(\mathcal{A})$
is a free $\mathit{S}$-module with a basis
$\theta_{1},\ldots ,\theta_{\ell}$
then the Poincar\'e polynomial
of
$\mathcal{A}$
is equal to
$\prod_{j=1}^{\ell}\left(1 + (\deg \theta_{j})t\right)$.
His proof of the freeness in \cite{Ath} uses the addition-deletion
theorem \cite{T80}.
Later M.
Yoshinaga extended this result in \cite{Y04}
to the generalized
Shi and Catalan arrangements and affirmatively settled the Edelman-Reiner
conjecture
\cite{ER}. However, even in the case of Shi arrangements,
no basis was constructed so far.
The organization of this paper is as follows:
in Section 2, we will define the polynomials
$B_{p,q}(x) $ which includes the Bernoulli polynomials.
In Section 3,
Theorem \ref{main} proves that the derivations constructed in Definition
\ref{basisdefinition} form
a basis for the derivation module $D(\Shi_{\ell} )$.
\section{The Bernoulli polynomials and $B_{p,q}(x)$}
Let $\mathit{B}_{k} (x)$
denote
the $k$-th
{\bf Bernoulli polynomial}.
Let
$\mathit{B}_{k}(0)=\mathit{B}_{k}$
denote the $k$-th
{\bf Bernoulli number}.
The most important property of the Bernoulli polynomial
in this paper
is the following elementary formula (e.g., \cite{APO76}):
\begin{theorem}
\label{bernoullipolynomialdifference}
\[
B_{k} (x+1) - B_{k}(x) = k x^{k-1}.
\]
\end{theorem}
\begin{define}
\label{Bpqdefinition}
For $(p,q) \in (\mathbb{Z}_{\geq 0})^2$, consider a polynomial
$\mathit{B}_{p,q} (x)$ in $x$ satisfying the following two conditions:
\begin{enumerate}
\item $\mathit{B}_{p,q} (x+1)- \mathit{B}_{p,q} (x)=(x+1)^{p} x^{q}$,
\item $\mathit{B}_{p,q} (0)=0$.
\end{enumerate}
It is easy to see that $\mathit{B}_{p,q} (x)$
is uniquely determined by these two conditions.
\end{define}
\begin{example}
\label{examplebpq}
(1)
When $(p,q)=(0,q)$, we have
$$\mathit{B}_{0,q} (x)=\frac{1}{q+1}
\left\{
\mathit{B}_{q+1}(x)-\mathit{B}_{q+1}\right\}$$
because of
Theorem \ref{bernoullipolynomialdifference}.
(2)
When $(p,q)=(p,0)$, we obtain
$$\mathit{B}_{p,0}(x)
=
\frac{(-1)^{p+1}}{p+1}
\left\{
\mathit{B}_{p+1}(-x)
-
\mathit{B}_{p+1}
\right\}
=
(-1)^{p+1} B_{0,p}(-x)
$$
because
\begin{align*}
&~~~~
(-1)^{p+1} B_{0,p}(-x-1)
-
(-1)^{p+1} B_{0,p}(-x)\\
&=
(-1)^{p}
\left\{
B_{0,p}(-x)
-
B_{0,p}(-x-1)
\right\}
= (-1)^{p} (-x-1)^{p}
= (x+1)^{p}.
\end{align*}
(3)
For a general $(p,q) \in (\mathbb{Z}_{\geq 0})^{2}$,
it easily follows from
Theorem \ref{bernoullipolynomialdifference}
that the polynomial has an expression in terms of the
Bernoulli polynomials as
\begin{equation*}
\mathit{B}_{p,q}(x)=
\sum_{i=0}^{p}
\frac{1}{q+i+1}
\binom{p}{i}
\left\{
\mathit{B}_{q+i+1}(x)-\mathit{B}_{q+i+1}
\right\}
=
\sum_{i=0}^{p}
\binom{p}{i}
B_{0,q+i} (x).
\end{equation*}
For example, $B_{1,1}(x) = B_{0,1}(x) + B_{0,2}(x) =
\frac{1}{3} (x^{3} - x) $.
\end{example}
Note that the polynomial $B_{p,q} (x)$
is a polynomial of degree $p+q+1$.
The homogenization $\overline{B}_{p,q} (x, z) $
of $B_{p, q} (x) $ is defined by
\[
\overline{B}_{p,q} (x, z) :=
z^{p+q+1} B_{p,q} \left(\frac{x}{z}\right).
\]
\section{A basis construction}
Let $1\leq j\leq \ell$. Define
$$\mathit{I}_{1}= \{ x_{1},x_{2},\ldots ,x_{j-1} \},
\,\,\,
\mathit{I}_{2}= \{ x_{j+2},x_{j+3},\ldots ,x_{\ell +1}
\}.$$
Let $\sigma_{k}^{(s)}$ denote the elementary
symmetric function
in the variables in
$\mathit{I}_{s}$ of degree $k$
$(s=1,2\ ,\ k \in \mathbb{Z}_{\geq 0})$.
Recall the homogeneous
polynomials
$\overline{B}_{p,q} (x, z)$
of degree $p+q+1$ defined at the end of the previous section.
\begin{define}
\label{basisdefinition}
Let
$\partial_{i} \,\,
(1\leq i\leq \ell+1)
$ and $\partial_{z} $ denote
$\partial/\partial x_{i} $
and
$\partial/\partial z $
respectively.
Define homogeneous derivations
$$
\eta_{1} := \sum_{i=1}^{\ell+1} \partial_{i}
\in D(\Shi_{\ell}),
\,\,\,\,
\eta_{2} := z \partial_{z} +\sum_{i=1}^{\ell+1} x_{i} \partial_{i}
\in D(\Shi_{\ell}),
$$
and
\begin{equation*}
\varphi_{j}
:=
(x_{j}-x_{j+1}-z)
\sum_{i=1}^{\ell +1} \sum_{\substack{
0\leq k_{1} \leq j-1 \\ 0\leq k_{2} \leq \ell-j}}
(-1)^{k_{1} +k_{2} }
\sigma_{j-1-k_{1}}^{(1)}
\sigma_{\ell-j-k_{2}}^{(2)}
\,
\overline{B}_{k_{1}, k_{2}}
(x_{i}, z) \partial_{i}
\end{equation*}
for
$1\leq j \leq \ell$.
\end{define}
We will prove
that the derivations
$
\eta_{1},
\eta_{2},$
and
$
\varphi_{1}, \dots , \varphi_{\ell}$
form a basis for $D(\Shi_{\ell})$.
First we will verify the following Proposition:
\begin{prop}
\label{Proposition6}
The derivations
$\varphi_{j}
\,\,
(1\leq j\leq \ell)
$
belong to the module
$D(\Shi_{\ell} )$.
\end{prop}
\noindent
{\textbf{proof.}}
We first have
\begin{align*}
\varphi_{j}(x_{p}-x_{q}) = &
(x_{j}-x_{j+1}-z) \\
&~\sum_{\substack{0\leq k_{1}\leq j-1 \\ 0\leq k_{2}\leq \ell-j}}
(-1)^{k_{1} +k_{2}}
\sigma_{j-1-k_{1}}^{(1)}
\sigma_{\ell-j-k_{2}}^{(2)}
\left\{
\overline{B}_{k_{1},k_{2}}(x_{p}, z)
-
\overline{B}_{k_{1}, k_{2}}(x_{q}, z)
\right\}.
\end{align*}
Since the right hand side equals zero if we set $x_{p}=x_{q}$,
we may conclude that $\varphi_{j} (x_{p}-x_{q})$
is divisible by
$x_{p} - x_{q} $ for all pairs $(p,q)$ with
$1\leq p < q\leq \ell+1$.
The congruent notation $\equiv$ in the following
calculation is modulo the ideal $(x_{p}-x_{q}-z)$:
\begin{align*}
&~~~\varphi_{j} (x_{p}-x_{q} -z)\\
&\equiv (x_{j}-x_{j+1}-z)\\
&~~~\sum_{\substack{0\leq k_{1}\leq j-1 \\ 0\leq k_{2}\leq \ell-j}}
(-1)^{k_{1} +k_{2}}
\sigma_{j-1-k_{1}}^{(1)}
\sigma_{\ell-j-k_{2}}^{(2)}
\left\{
\overline{B}_{k_{1}, k_{2}}(x_{p}, x_{p} - x_{q})
-
\overline{B}_{k_{1}, k_{2}}(x_{q}, x_{p} - x_{q})
\right\}\\
&=(x_{j}-x_{j+1}-z)
\sum_{\substack{
0\leq k_{1}\leq j-1\\
0\leq k_{2}\leq \ell-j}}
(-1)^{k_{1} +k_{2} }
\sigma_{j-1-k_{1}}^{(1)} \sigma_{\ell-j-k_{2}}^{(2)} \\
& ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
(x_{p}-x_{q})^{k_{1} + k_{2}+1}
\{ \mathit{B}_{k_{1}, k_{2}}
(\frac{x_{p}}{x_{p}-x_{q}}) - \mathit{B}_{k_{1}, k_{2}}
(\frac{x_{q}}{x_{p}-x_{q}}) \} \\
&=(x_{j}-x_{j+1}-z)
\\
&~~~~~~~~~
\sum_{\substack{0\leq k_{1}
\leq j-1 \\ 0\leq k_{2}\leq \ell-j}}
(-1)^{k_{1} +k_{2} }
\sigma_{j-1-k_{1}}^{(1)} \sigma_{\ell-j-k_{2}}^{(2)}
(x_{p}-x_{q})^{k_{1} + k_{2}+1 } (\frac{x_{p}}{x_{p}-x_{q}})^{k_{1}}
(\frac{x_{q}}{x_{p}-x_{q}})^{k_{2}} \\
&= (x_{j}-x_{j+1}-z) (x_{p}-x_{q})
\sum_{\substack{
0\leq k_{1}\leq j-1 \\ 0\leq k_{2}\leq \ell-j}}
(-1)^{k_{1} +k_{2} }
\sigma_{j-1-k_{1}}^{(1)} \sigma_{\ell-j-k_{2}}^{(2)}
x_{p}^{k_{1}} x_{q}^{k_{2}} \\
&= (x_{j}-x_{j+1}-z) (x_{p}-x_{q}) \sum_{k_{1}=0}^{j-1}
\sigma_{j-1-k_{1}}^{(1)}
(-x_{p})^{k_{1}}
\sum_{k_{2}=0}^{\ell -j}
\sigma_{\ell-j-k_{2}}^{(2)} (-x_{q})^{k_{2}}\\
&= (x_{j}-x_{j+1}-z) (x_{p}-x_{q})
\prod_{s=1}^{j-1}(x_{s}-x_{p}) \prod_{s=j+2}^{\ell +1}(x_{s}-x_{q})
\equiv
0
\end{align*}
for all pairs $(p, q)$ with $1\leq p < q \leq \ell +1$.
$\square$
\begin{lemma}
\label{detSigma}
Suppose $\ell\geq 1$.
Let $N$ be the $\ell\times\ell$-matrix whose
$(i, j)$-entry is equal to the elementary symmetric function of degree
$\ell-i$
in the variables $x_{1}, \dots, x_{j-1}, x_{j+2}, \dots, x_{\ell+1 } $.
Then
\[
\det N
=
(-1)^{\ell(\ell-1)/2}
\prod_{\substack{1\leq p<q\leq\ell\\ q-p>1}} (x_{p} - x_{q} ).
\]
\end{lemma}
\noindent
{\textbf{proof.}}
Note that we have the equality
\begin{align*}
&~~~
\left[
1 \,\,\, -x_{p} \,\,\,\, (-x_{p})^{2} \dots (-x_{p})^{\ell-2}
\,\,\,\, (-x_{p})^{\ell-1}
\right]
N\\
&=
\left[
\prod_{\substack{1\leq s\leq \ell+1\\
s\not\in \{1,2\}}}
(x_{s} - x_{p})
\,
\prod_{\substack{1\leq s\leq \ell+1\\
s\not\in \{2,3\}}}
(x_{s} - x_{p})
\,
\dots
\,
\prod_{\substack{1\leq s\leq \ell+1\\
s\not\in \{\ell-1,\ell\}}}
(x_{s} - x_{p})
\,
\prod_{\substack{1\leq s\leq \ell+1\\
s\not\in \{\ell,\ell+1\}}}
(x_{s} - x_{p})
\right]
\end{align*}
for any $1\leq p\leq \ell$.
Suppose that
$$1\leq p < q \leq \ell+1,
\,\,\,
q-p>1.$$
Set $x_{p} = x_{q} $ in $N$, and we get
$N_{pq} $.
Then we may conclude that
\begin{align*}
&~~~
\left[
1 \,\,\, -x_{p} \,\,\,\, (-x_{p})^{2} \dots (-x_{p})^{\ell-2}
\,\,\,\, (-x_{p})^{\ell-1}
\right]
N_{pq}
={\mathbf 0}.
\end{align*}
This implies that
$\det N_{pq} =0$
and that
$
\det N
$
is divisible by
$x_{p} - x_{q}.$
Since
\[
\deg
(\det
N)
=
\ell(\ell-1)/2
=
\deg
\prod_{\substack{1\leq p<q\leq\ell+1\\ q-p>1}}
(x_{p} - x_{q}),
\]
there exists a constant $C$ such that
\[
\det N
=
C\,
(-1)^{\ell(\ell-1)/2}
\prod_{\substack{1\leq p<q\leq\ell+1\\ q-p>1}}
(x_{p} - x_{q})
=
C
\prod_{\substack{1\leq p<q\leq\ell+1\\ q-p>1}}
(x_{q} - x_{p}).
\]
By comparing the coefficients of
$
x_{3} x_{4}^{2} \dots x_{\ell}^{\ell-2} x_{\ell+1}^{\ell-1}
$
on both sides, we obtain $C=1$.
$\square$
\begin{prop}
\label{Proposition8}
The derivations
$\eta_{1}, $
$\eta_{2}, $
$\varphi_{1} , \dots , \varphi_{\ell}$
are linearly independent over $S$.
\end{prop}
\noindent
{\textbf{proof.}}
Set $z=0$ in $\varphi_{j} $ and we get
$\phi_{j} $ as follows:
\begin{align*}
\phi_{j} &:= \varphi_{j} |_{z=0}
=
(x_{j}-x_{j+1})
\sum_{i=1}^{\ell +1} \sum_{\substack{
0\leq k_{1} \leq j-1 \\ 0\leq k_{2} \leq \ell-j}}
\frac{(-1)^{k_{1} +k_{2} }}{k_{1} +k_{2} +1}
\sigma_{j-1-k_{1}}^{(1)}
\sigma_{\ell-j-k_{2}}^{(2)}
\,
x_{i}^{k_{1} +k_{2} +1} \partial_{i}\\
&=
(x_{j}-x_{j+1})
\sum_{k=1}^{\ell}
\frac{(-1)^{k-1}}{k}
\left(
\sum_{\substack{k_{1} +k_{2} +1 =k\\
0\leq k_{1} \leq j-1 \\ 0\leq k_{2} \leq \ell-j}}
\sigma_{j-1-k_{1}}^{(1)}
\sigma_{\ell-j-k_{2}}^{(2)}
\right)
\sum_{i=1}^{\ell +1}
x_{i}^{k}
\partial_{i}\\
&=
(x_{j}-x_{j+1})
\sum_{k=1}^{\ell}
\frac{(-1)^{k-1}}{k}\,
\sigma_{\ell-k}
(x_{1}, \dots, x_{j-1}, x_{j+2}, \dots, x_{\ell+1})
\,
\sum_{i=1}^{\ell +1}
x_{i}^{k}
\partial_{i}.
\end{align*}
Here
$
\sigma_{\ell-i}
(x_{1}, \dots, x_{j-1}, x_{j+2}, \dots, x_{\ell+1})
$
stands for
the elementary symmetric function
of degree $\ell-i$ in the variables
$x_{1}, \dots, x_{j-1}, x_{j+2}, \dots, x_{\ell+1}.
$
This is equal to
the $(i, j)$-entry
$N_{ij} $ of the matrix
$
N
$
in Lemma \ref{detSigma}.
Thus we have
\begin{equation}
\phi_{j} (x_{i}) =
(x_{j}-x_{j+1})
\sum_{k=1}^{\ell}
\frac{(-1)^{k-1}}{k}\,
x_{i}^{k}
N_{kj}.
\label{phijxi}
\end{equation}
Define
two $(\ell+1)\times(\ell+1)$-diagonal
matrices
$D_{1} $ and
$D_{2} $
by
\begin{align*}
D_{1}
&:=
\left[1\right]\oplus\left[1\right]\oplus\left[(-1)^{1}/2
\right]\oplus \left[(-1)^{2}/3\right]
\oplus\dots\oplus [(-1)^{\ell-1}/\ell],
\\
D_{2}
&:=
[1]\oplus[x_{1} -x_{2}]\oplus[x_{2} - x_{3}]
\oplus
\dots
\oplus
[x_{\ell}-x_{\ell+1}],
\end{align*}
where $\oplus$ stands for the direct sum of matrices.
Also define
two $(\ell+1)\times(\ell+1)$-matrices
$\tilde{N} $ and
$M$
by
$$
\tilde{N} := [1]\oplus N,
\,\,\,\,\,
M:=
\left[
x_{i}^{j-1}
\right]_{
1\leq i\leq\ell+1,
1\leq j\leq\ell+1
} .
$$
From (\ref{phijxi})
we obtain
\[
P
:=
\begin{bmatrix}
1&\phi_{1}(x_{1}) &\dots&\phi_{\ell}(x_{1})\\
1&\phi_{1} (x_{2}) &\dots&\phi_{\ell}(x_{2})\\
1&\phi_{1} (x_{3}) &\dots&\phi_{\ell}(x_{3})\\
.& . & \dots& .\\
.& . & \dots& .\\
1&\phi_{1} (x_{\ell+1}) &\dots&\phi_{\ell} (x_{\ell+1})
\end{bmatrix}
=
M
D_{1}
\tilde{N}
D_{2}.
\]
Thus, by applying the Vandermonde determinant
formula
and Lemma \ref{detSigma}, we deduce
\begin{align*}
\det P
&=
(\det M)
(\det D_{1})
(\det \tilde{N})
(\det D_{2})
\\
&=
\left(
\prod_{1\leq p < q\leq \ell+1} (x_{q}-x_{p})
\right)
\left(
\frac{(-1)^{\ell(\ell-1)/2}}{\ell !}
\right)
\left(
\det {N}
\right)
\prod_{1\leq p \leq \ell} (x_{p}-x_{p+1})
\\
&=
\left(
\frac{(-1)^{\ell(\ell+1)/2}}{\ell !}
\right)
\prod_{1\leq p < q \leq \ell+1} (x_{p}-x_{q})^{2}
\neq 0.
\end{align*}
Thus
$\eta_{1}, \phi_{1}, \dots, \phi_{\ell}$
are linearly independent.
This implies that
$\eta_{1}, \varphi_{1}, \dots, \varphi_{\ell}$
are linearly independent.
Since $\eta_{2}(z) = z $
and
$\eta_{1}(z)=\varphi_{1}(z)= \dots =\varphi_{\ell}(z)=0$,
we finally conclude that
$\eta_{1}, \eta_{2}, \varphi_{1}, \dots, \varphi_{\ell}$
are linearly independent.
$\square$
\bigskip
\noindent
{\em Remark.}
The derivations
$\phi_{1}, \dots, \phi_{\ell} $ are a basis
for the derivation module of the double
Coxeter arrangement of the
type $A_{\ell}$ studied in \cite{ST98}
(cf. \cite{T02}).
\begin{theorem}
\label{main}
The derivations $\eta_{1} , \eta_{2}, \varphi_{1}, \dots, \varphi_{\ell} $
form a basis for $D(\Shi_{\ell})$.
\label{main}
\end{theorem}
\noindent
{\textbf{proof.}}
We may apply
Saito's criterion
\cite{S80} \cite[Theorem 4.23]{OT}
thanks to Propositions \ref{Proposition6}
and
\ref{Proposition8}
because
\[
\deg \eta_{1} + \deg \eta_{2} +\sum_{j=1}^{\ell} \deg \varphi_{j}
=
1+\ell(\ell+1)=|\Shi_{\ell} |.
\,\,\,
\square
\]
\noindent
{\textit{Remark.}}
The Bernoulli polynomials explicitly appear
in the first derivation $\varphi_{1} $ and the last one
$\varphi_{\ell} $ because of Example \ref{examplebpq} (1) and (2):
\begin{align*}
\varphi_{1}
&=
(x_{1}-x_{2}-z)
\sum_{i=1}^{\ell +1} \sum_{k_{2} =0}^{\ell-1}
(-1)^{k_{2} }
\sigma_{\ell-1-k_{2} }^{(2)}
\,
\overline{B}_{0,k_{2} }
(x_{i}, z) \partial_{i}\\
&=
(x_{1}-x_{2}-z)
\sum_{i=1}^{\ell +1}
\sum_{k=1}^{\ell}
\frac{(-1)^{k-1}}{k}
\sigma_{\ell-k}^{(2)}
z^{k}
\left(
B_{k}(x_{i}/z)
-
B_{k}
\right)
\partial_{i},
\end{align*}
and
\begin{align*}
\varphi_{\ell}
&=
(x_{\ell}-x_{\ell+1}-z)
\sum_{i=1}^{\ell +1} \sum_{k_{1} =0}^{\ell-1}
(-1)^{k_{1} }
\sigma_{\ell-1-k_{1} }^{(1)}
\,
\overline{B}_{k_{1} ,0}
(x_{i}, z) \partial_{i}\\
&=
(x_{\ell}-x_{\ell+1}-z)
\sum_{i=1}^{\ell +1} \sum_{k=1}^{\ell}
\frac{(-1)^{k-1}}{k}
\sigma_{\ell-k}^{(1)}
\,
(-z)^{k}
\left(
B_{k}(-x_{i}/z)
-
B_{k}
\right)
\partial_{i}.
\end{align*}
Here
$\sigma^{(1)}_{d}$
and
$\sigma^{(2)}_{d}$
are the elementary
symmetric functions
of degree $d$ in the variables
$x_{1}, \dots, x_{\ell-1} $
and
$x_{3}, \dots, x_{\ell+1} $
respectively.
\begin{example}
For $A_{3}$, we have
\begin{align*}
\eta_{1} &= \partial_{1} +\partial_{2} +\partial_{3} +\partial_{4}, \,\,\,\,\,\,\,\,
\eta_{2} = x_{1} \partial_{1} +x_{2} \partial_{2} +x_{3} \partial_{3} +x_{4} \partial_{4} +z \partial_{z}, \\
\varphi_{1}&=
x_{1}(x_{1}-x_{2}-z)
\left\{
x_{3}x_{4}
-
\frac{1}{2}
(x_{3}+x_{4})
(x_{1}
-z)
+
\frac{1}{3}
\left(x_{1}^{2}
-
\frac{3}{2}x_{1}z
+
\frac{1}{2}
z^{2}\right)
\right\}
\partial_{1} \\
&\, +
x_{2}
(x_{1}-x_{2}-z)
\left\{
x_{3}x_{4}
-
\frac{1}{2}
(x_{3}+x_{4})
(x_{2}
-z)
+
\frac{1}{3}
\left(x_{2}^{2}
-
\frac{3}{2}x_{2}z
+
\frac{1}{2}
z^{2}\right)
\right\}
\partial_{2} \\
&\ \ -\frac{1}{6}x_{3}(x_{1}-x_{2}-z)(x_{3}+z)(x_{3}-3x_{4}-z)\partial_{3} \\
&\ \ -\frac{1}{6}x_{4}(x_{1}-x_{2}-z)(x_{4}+z)(x_{4}-3x_{3}-z)\partial_{4},
\\
\varphi_{2}&=-\frac{1}{6}x_{1}(x_{2}-x_{3}-z)(x_{1}-z)(x_{1}-3x_{4}-2z)
\partial_{1} \\
&\ \ +
x_{2}(x_{2}-x_{3}-z)
\left\{
x_{1}x_{4}
-
\frac{1}{2}x_{1}
(x_{2}-z)
-
\frac{1}{2}x_{4}
(x_{2}+z)
+
\frac{1}{3}
(x_{2}^{2}
-z^{2}
)\right\}
\partial_{2} \\
&\ \ +
x_{3}(x_{2}-x_{3}-z)
\left\{
x_{1}x_{4}
-
\frac{1}{2}x_{1}
(x_{3}-z)
-
\frac{1}{2}x_{4}
(x_{3}+z)
+
\frac{1}{3}
(x_{3}^{2}
-z^{2}
)\right\}
\partial_{3} \\
&\ \ +\frac{1}{6}x_{4}(x_{2}-x_{3}-z)(x_{4}+z)(3x_{1}-x_{4}-2z)\partial_{4},
\\
\varphi_{3}&
=-\frac{1}{6}x_{1}(x_{3}-x_{4}-z)(x_{1}-z)(x_{1}-3x_{2}+z)\partial_{1} \\
&\ \ \, -\frac{1}{6}x_{2}(x_{3}-x_{4}-z)(x_{2}-z)(x_{2}-3x_{1}+z)\partial_{2} \\
&\, +
x_{3}(x_{3}-x_{4}-z)
\left\{
x_{1}x_{2}
-
\frac{1}{2}
(x_{1}+x_{2})
(x_{3}
+z)
+
\frac{1}{3}
\left(x_{3}^{2}
+
\frac{3}{2}x_{3}z
+\frac{1}{2}
z^{2}\right)
\right\}
\partial_{3} \\
&\,
+
x_{4}(x_{3}-x_{4}-z)
\left\{
x_{1}x_{2}
-
\frac{1}{2}
(x_{1}+x_{2})
(x_{4}+z)
+
\frac{1}{3}
\left(x_{4}^{2}
+
\frac{3}{2}x_{4}z+
\frac{1}{2}
z^{2}\right)
\right\}
\partial_{4}.
\end{align*}
\end{example}
\bigskip
\noindent
{\textit{Problem.}}
Construct a basis for the derivation module
$D(\A)$ when $\A$ is the cone over the generalized
Shi arrangement of one
of the other types: $B_{\ell}, C_{\ell},
D_{\ell}, E_{6}, E_{7}, E_{8},
F_{4}$ and $G_{2}.$
It has been known that the derivation modules
are free modules for these cases by M. Yoshinaga
\cite{Y04}. It seems interesting to learn what
kind of polynomials plays the role of
the Bernoulli polynomials in the case of
the type $A_{\ell}$.
|
\section{Introduction}
The Higgs boson decaying to four hard and isolated leptons provides the cleanest channel (the so-called ``gold-plated'' channel) for its discovery. The only significant standard model (SM) background is di-boson ($ZZ$) production. Other, usually noxious SM backgrounds, such as $t\bar{t}$, $W/Z+\text{ jets}$, are negligible. However, the gold-plated channel gets activated only for a relatively heavy SM Higgs boson ($m_h \gtrsim 130~\text{GeV}$~\cite{Aad:2009wy}), when the Higgs can decay to four isolated leptons via $ZZ^*$ or $ZZ$. Unfortunately, in weak-scale supersymmetry, which remains the most well-motivated ultraviolet (UV) extension of the standard model (SM), the mass of the lightest CP-even Higgs boson is predicted to be $\lesssim 125~\text{GeV}$ for stop masses and mixings that do not exceed $1~\text{TeV}$~\cite{Carena:2000dp} . Higgses in this mass range decay dominantly to $\bar{b} b$ and are notoriously hard to find. Only by using sophisticated techniques such as jet-substructure can we hope to find the Higgs in this mass range at $5~\sigma$ confidence from $10~\text{fb}^{-1}$ of data~\cite{Kribs:2009yh,Kribs:2010hp,Kribs:2010ii}.
The goal of this paper is to open up the $4\,\ell$ mode for light, supersymmetric Higgses. This requires extending the Minimal Supersymmetric Standard Model (MSSM). However, as we show, when the $4\,\ell$ channel is resurrected for light Higgses, the implications are spectacular. A tiny branching fraction of $h \rightarrow 4\,\ell$ can render the Higgs easily discoverable, even with $1~\text{fb}^{-1}$ of data at the LHC operating at $\sqrt{s} = 7~\text{TeV}$ center of mass energy.
In many extensions of the MSSM, a pseudo-scalar (namely, $a$) can be lighter than the Higgs boson, and will mediate unconventional and exciting decay modes~\cite{Chang:2005ht, Carena:2007jk, Chang:2008cw}. The simplest unconventional mode is a light Higgs decaying to four final state particles, though decays to even higher multiplicity states are also possible~\cite{Chang:2005ht}. However, $a$ has Higgs-like couplings, so it decays dominantly to the heaviest particles that are kinematically allowed rather than to light leptons. Typically, $a \rightarrow \bar b b$, implying $h \rightarrow 2b\, 2\bar b$, but there are corners of parameter space where $2 m_\tau < m_a < 2 m_b $ and $a$ decays to $2\,b$ are forbidden. In these corners, $h \rightarrow 2 a \rightarrow 2 \tau\, 2 \mu$ has been shown to be a viable discovery mode~\cite{Lisanti:2009uy}. However, such ultra-light $a$ give rise to soft and, often, collinear muons, and are tricky to analyze.
In numerous recent models of dark matter~\cite{ArkaniHamed:2008qn,Cheung:2009qd, Morrissey:2009ur}, as in hidden valley models~\cite{Strassler:2006im,Strassler:2006qa, Han:2007ae}, a new abelian gauge boson $Z'$ is introduced. If the Higgs couples to these $Z'$, $h \rightarrow Z'Z' \rightarrow 4\, \ell$ is feasible. Unfortunately, $Z'$ introduced for dark matter purposes tend to be extremely light $M_{Z'} \sim~\text{GeV}$, making their decay products predominantly collinear and/or soft. The final states from $h \rightarrow Z'Z' \rightarrow 4\, \ell$ do contain $4$ leptons, but rarely $4$ {\it hard} and {\it isolated} leptons. Though studies have shown that it is possible to extract the Higgs by looking for jets made entirely of collinear leptons (so-called lepton jets~\cite{Falkowski:2010gv}), these methods are highly sensitive to detector effects. It remains to be seen how well these objects can be identified and how well the Higgs mass can be reconstructed in a realistic detector simulation.
The introduction of a heavier Higgsphilic $Z'$ resurrects the gold-plated channel for discovery of a light Higgs. Specifically, this $Z'$ must be sufficiently heavy so that it remains un-boosted after the decay $h \rightarrow Z'Z'$. Provided $M_{Z'} \gtrsim 0.1\,m_h$, the boost is sufficiently small and, consequently, massless $Z'$ decay products are separated enough to be individually distinguished in a detector. An upper limit on the $Z'$ mass comes simply from kinematics, $M_{Z'} \lesssim m_h/2$. In short, a $Z'$ mass on the order of the electroweak scale does the job quite well. A hidden valley with such a $Z'$ was recently proposed for the sake of involving the SM light Higgs to four fermions decay~\cite{Gopalakrishna:2008dv}. However, the structure of supersymmetry severely restricts couplings and arbitrary mixing of a generic hidden abelian sector with Higgs is not allowed.
The fusion of supersymmetry with Higgsphilic $Z'$ can do more than just facilitate Higgs discovery. In fact, combining multiple leptons signals with large missing energy, the result is an extremely clean channel for discovering supersymmetry itself. In the MSSM, multi-lepton plus $\slashchar E_T$ signals arise via cascades containing di-bosons or combinations of light sleptons and $W/Z$s. These rates are usually small, suppressed by the small fraction of cascades which contain $W/Z$. However, if a weak-scale $Z'$ is added to the model, we open up the possibility of large rates to $4\,\ell + \slashchar E_T$. In this scenario, heavier neutralinos prefer to decay to lighter neutralinos by emitting a $Z'$. As all supersymmetric events must terminate in a pair of the lightest neutralinos (the lightest supersymmetric particle in this setup), a large fraction of all supersymmetric events contain $Z'Z' + \slashchar E_T$, paving the way for many events with $4\,\ell + \slashchar E_T$. As we will show, the success of this channel is controlled essentially by the Higgsino masses. If the Higgsinos are light enough, the $4\, \ell + \slashchar E_T$ channel can bring discovery regardless of how heavy the squarks, sleptons, and gluinos are.
Remarkably, a $Z'$ with exactly the desired properties is predicted in the ``viable'' gravity-mediated supersymmetry breaking model~\cite{Kribs:2010md}. In this model, the flavor problem~\cite{Gabbiani:1988rb,Gabbiani:1996hi,Bagger:1997gg,Ciuchini:1998ix} associated with gravity mediation~\cite{Chamseddine:1982jx,Barbieri:1982eh,Ibanez:1982ee,Hall:1983iz,Ohta:1982wn,Ellis:1982wr,AlvarezGaume:1983gj,Nilles:1983ge,Nath:1983fp} is solved -- not by imposing continuous flavor symmetries~\cite{Barbieri:1995uv,Barbieri:1996ww} or
gauged discrete family symmetries~\cite{Kaplan:1993ej,ArkaniHamed:1996xm} to align the mass matrices~\cite{Hall:1990ac,Nir:1993mx} --
but by the emergence of an approximate and an accidental $\text{U(1)}_\text{R}$ symmetry. The complete model contains an additional gauged abelian symmetry (say, $\text{U(1)}_X$), which is broken along with the electroweak symmetry right at the electroweak scale. Unlike the hidden sector models, the $\text{U(1)}_X$ is completely {\it visible} to us since there exists electroweak doublets that carry $X-$charges. The coexistence of supersymmetry along with an electroweak scale Higgs-phylllic $Z'$, ensures significant discovery potential for Higgs/higgsionos via the discovery of $Z'$.
Instead of analyzing the full model in Ref.~\cite{Kribs:2010md}, we construct a ``bare-bones'' model where we integrate out all particles not necessary for a discussion of $Z'$ phenomenology. The bare-bones model contains only the SM fermions, the CP-even lightest Higgs boson, Higgsinos, and an additional singlino. This simplified approach to models has recently been advocated in Ref.~\cite{Alwall:2008ag}.
Within current experimental constraints as calculated from the Tevatron results, we find that $Z'$ in the bare-bones model can be discovered easily in a sample of Higgs events (events with $4\,\ell$) as well as in a sample of supersymmetric events ($4\,\ell + \slashchar{E}_T$ events) easily and, thus, can pave the way for the discovery of supersymmetry and a supersymmetric Higgs.
We begin with a description of the bare-bones model in Sec.~\ref{sec:model}. In Sec.~\ref{sec:pheno} we
show that the Tevatron data puts severe constraints on $h \rightarrow Z'Z' \rightarrow 4\, \ell$ branching fraction, and on Higgsino masses. In the same section we also discuss the discovery potential of $Z'$, Higgs, and Higgsinos at the LHC within these constraints. For readers interested in how the bare-bones model arises from a more fundamental theory, we show the mapping between the viable gravity mediation scenario and the bare-bones model in Sec.~\ref{sec:viable-grav-medi}. Our concluding reflections are in Sec.~\ref{sec:conclusion}.
\section{The bare-bones model of a supersymmetric Higgsphilic $Z'$}
\label{sec:model}
We seek a bare-bones model with the signal
\begin{itemize}
\item $4\,\ell$: signature of Higgs decaying to $2 Z' \rightarrow 4\,\ell$.
\item Inclusive $4\,\ell + \slashchar{E}_T$: signature of supersymmetric events with $2Z' + \text{LSPs} \rightarrow 4\,\ell + \slashchar{E}_T$.
\end{itemize}
The bare-bones model is not manifestly supersymmetric, but it should be understood as an effective
theory of a fully supersymmetric and UV-complete model. More precisely, it is the effective theory of the viable gravity mediation. The particle content of the bare-bones model is listed below:
\begin{table}[h]
\label{bare-bones-model-content}
\begin{tabular}{ c | c }
\hline
$h$ & The lightest CP-even Higgs boson \\
$Z'$ & \parbox[c][0.4 in]{2.2in}{
New neutral gauge boson of a broken symmetry $\text{U(1)}_X$, with mass $M_{Z'}$ }\\
$\tilde S$ & \parbox[c][0.5 in]{2.2in}{ A Dirac singlino and the candidate for the lightest supersymmeric particles (LSP)} \\
$\tilde H \equiv
\begin{pmatrix}
\tilde C \\ \tilde N
\end{pmatrix}
$ & \parbox[c][0.5 in]{2.2in}{ An electroweak doublet of Dirac inos. $\tilde N $ and $\tilde C $ are the neutral and charged components respectively.} \\
\hline
\end{tabular}
\caption{List of particles in the bare-bones model}
\end{table}
Next, we compile the list of interactions involving $Z'$, that are necessary to generate the signals itemized in the beginning of this section.
\begin{gather}
-\frac{\epsilon}{2} \ Z'_{\mu \nu} B^{\mu \nu}
\label{kin-mix}\\
g_0 \: M_{Z'} \ h\, Z'_{\mu} Z'^{\mu}
\label{int-hz'z'}\\
g_X\: \eta \ \bar{\tilde N} \gamma_\mu P_L \tilde S \ Z'^{\mu}
\ + \ g_X \: \zeta \ \bar{\tilde N} \gamma_\mu P_R \tilde S \ Z'^{\mu} + H.c.
\label{int-nhz'}
\end{gather}
In the above, $B_\mu$ is the $\text{U(1)}_Y$ gauge boson, $\epsilon, g_0, \eta, \zeta$ are small coupling constants, and $g_X$ is the strength of broken abelian symmetry $\text{U(1)}_X$. In the bare-bones model $g_X$ is weaker than the weak scale gauge couplings $g_Z$ or $g_W$, the strength of interactions between $Z$ or $W$ and their respective current.
The coupling $\epsilon$ in Eq.~\eqref{kin-mix} dictates the amount of mixing between $Z'$ and $B$. Its effects can be unearthed after redefining $B_\mu \rightarrow B_\mu - \epsilon Z'_\mu$. Such a redefinition, undoes the mixing but generates coupling between the hypercharge current $J_Y$ and $Z'$. Consequently, non zero Higgs vacuum expectation value (vev) gives rise to an additional $Z-Z'$ mixing. Undoing the $Z-Z'$ mixing, we find an additional coupling between $J_Z$ (the $Z-$current) and $Z'$~\cite{Gopalakrishna:2008dv}. The effect of the operator in Eq.~\eqref{kin-mix} can be summarized in a following interaction that shows the couplings of $Z'$ with the SM fermions in a straight-forward way:
\begin{equation}
\label{eq:kin-mix-final}
\epsilon \: g _Y Z'_\mu \left( J_Y^\mu \ + \
\frac{M_Z^2}{M_Z^2 - M_{Z'}^2} J_Z^\mu \right) \ \; ,
\end{equation}
where $g_Y$ is the hypercharge gauge coupling. Since the parameter $\epsilon$ is constrained to be less than or in the order of $10^{-2}$~\cite{Hook:2010tw}, as long as $Z'$ appears at the electroweak scale, the couplings in Eq.~\eqref{eq:kin-mix-final} have little phenomenological impact. However, the $\mathcal O(\epsilon)$ couplings do become important if the $Z'$ is kinematically restricted and can only decay via Eq.~\eqref{eq:kin-mix-final}. In this case, while the total width of $Z'$ depends on $\epsilon$, its branching fractions to various SM fermions do not. In fact, in the limit $M_Z^2 \gg M_{Z'}^2$, the branching fractions are completely determined:
\begin{equation}
\label{eq:5}
\begin{split}
& \text{Br}\left( Z' \rightarrow \tau^+ \tau^- \right) \sim 13\% \quad
\text{Br}\left( Z' \rightarrow \ell^+ \ell^- \right) \sim 28\% \\
& \quad \text{Br}\left( Z' \rightarrow q \bar{q} \right) \sim 60\% \quad \quad
\text{Br}\left( Z' \rightarrow b \bar{b} \right) \sim 7\% \\
& \qquad \qquad \qquad \quad
\text{Br}\left( Z' \rightarrow \nu \nu \right) \lesssim 1\% \; ,
\end{split}
\end{equation}
where $q$ refers to all quarks and $\nu$ refers to neutrinos of all flavor.
Since the bare-bones model has been pointed out to be an effective description of a supersymmetric UV complete model, the Higgs boson $h$ should be considered to be the part of a fully supersymmetric Higgs sector. We will show later in Sec.~\ref{sec:viable-grav-medi} that when the bare-bones model is derived from the viable gravity mediation, the couplings of the Higgs to the SM fermions and the electroweak gauge bosons remain SM-like to the leading order of small model parameters. Higgs in the bare-bones inherits all SM interactions with approximately SM-like strengths.
Its production is thus identical to a SM Higgs. The only difference in properties of the Higgs boson is that if $m_h > 2 M_{Z'}$, it can decay to a pair of $Z'$ because of the operator in Eq.~\eqref{int-hz'z'}. The branching width of $h \rightarrow 2 Z' \rightarrow 4\,\ell$ is quadratically sensitive to the coupling constant $g_0$. However, because of the tiny total decay width of the lightest CP-even MSSM Higgs, even a small $g_0$ results in a large branching fraction.
Operators in Eq.~\eqref{int-nhz'} are also responsible for the decay of heavier neutralinos to lighter neutralinos with the emission of a $Z'$. Note that, since $\tilde N$ carries electroweak charge, it can be produced by Drell-Yan processes at the LHC, while $\tilde S$ cannot. Hence, realistically speaking, we must also assume $M_{\tilde N} > M_{\tilde S}$ in order to generate any $Z'$ events.
To flesh out what the bare-bones model predicts for supersymmetric events, we need to list additional interactions. These interactions do not involve $Z'$ directly, but have important implications for $Z'$ phenomenology.
\begin{gather}
- \frac{g_Z}{2}\: \eta \ \bar{\tilde N} \gamma_\mu P_L \tilde S \ Z^{\mu}
\ - \ \frac{g_Z}{2} \: \zeta \ \bar{\tilde N} \gamma_\mu P_R \tilde S \ Z^{\mu} + H.c.
\label{int-nhz} \\
- \frac{g_Z}{\sqrt{2}}\: \eta \ \bar{\tilde C} \gamma_\mu P_L \tilde S \ W^{\mu}
\ - \ \frac{g_Z}{\sqrt{2}} \: \zeta \ \bar{\tilde C} \gamma_\mu P_R \tilde S \ W^{\mu}+H.c.
\label{int-ncw}
\end{gather}
The operators in Eq.~\eqref{int-nhz} allow $\tilde N \rightarrow \tilde S + Z$ decay, however as long as $ M_{Z} > ( M_{\tilde N} - M_{\tilde S} ) > M_{Z'}$, the decay mode $\tilde N \rightarrow \tilde S + Z' $ dominates. For other mass hierarchies, both the decay modes compete. Decays to $Z$ are slightly suppressed because of a larger $Z$ mass, but are enhanced in the small $g_X/g_Z$ limit.
The last set of operators (Eq.~\eqref{int-ncw}) provide an alternative path for charged Higgsinos ({\it i.e.} $\tilde C$) to decay to neutralinos, $\tilde C \rightarrow \tilde S + W$. As $\tilde C$ and $\tilde N$ are part of the same electroweak doublet, they are much closer in mass than $\tilde C$ and $\tilde S$. Consequently, kinematics suppresses $\tilde C\rightarrow \tilde N + W$ compared to $\tilde C\rightarrow \tilde S + W$, with the $W$ in the former case often off-shell. Thus, neither chargino pair production nor associated chargino plus neutralino production will generate events contains $2\,Z'$; chargino pairs lead to $W^+W^- + \slashchar E_T$, while chargino plus neutralino leads to $W^{\pm}Z' + \slashchar E_T$. As a result, the only supersymmetric source of $4\,\ell + \slashchar E_T$ events is neutralino pair production. \\
Summarizing,
\begin{itemize}
\item The $Z'$ branching fractions to various SM fermions is completely determined given its mass.
It has a large width to leptons and, hence, is easier to find in $Z' \rightarrow 2 \ell$. Unlike $Z$ bosons, the $Z'$ width to neutrinos is minuscule, suppressed by $\left( M_{Z'}/ M_Z \right)^2$ compared to the leptonic width.
\item The channel $h \rightarrow Z'Z' \rightarrow 4\,\ell$ is a clean channel for the simultaneous discovery of $Z'$ and Higgs as long as $m_h > 2 M_{Z'}$. For a given $m_h$ and $M_{Z'}$, the rates of $4\,\ell$ events depend only on the branching fraction of $h \rightarrow Z'Z'$, which is governed by the parameter $g_0$.
\item Inclusive $4\,\ell + \slashchar{E}_T$ is another clean signature to find $Z'$ along with supersymmetry in the bare-bones model. Provided $ M_{Z} > ( M_{\tilde N} - M_{\tilde S} ) > M_{Z'}$ almost all neutralino-pair events contain two $Z'$, and the rate for $\tilde N \tilde N \rightarrow 2 Z' + 2 \tilde S \rightarrow 4 \ell + \slashchar{E}_T$ depends only on the mass of $\tilde N$.
\end{itemize}
\section{New Physics via Signals of $Z'$ at the Collider}
\label{sec:pheno}
In this section, we investigate the discovery potential of three distinct signals of the bare-bones model: $(i)$ signals of Higgs in $4\,\ell$, $(ii)$ signals of supersymmetry in $4\,\ell + \slashchar{E}_T$, and finally
$(iii)$ signals of long lived $Z'$ in $ h \rightarrow Z'Z'$ events.
\subsection{ $4 \ell$ without Missing Energy}
\label{sec:higgs_prompt}
Unlike most collider studies of the lightest CP-even Higgs boson in the MSSM~\cite{Djouadi:2005gj}, we seek Higgses produced via gluon fusion.
\begin{equation}
\label{eq:h-topo}
p p \text{ or } p \bar{p} \rightarrow h \rightarrow Z' Z' \rightarrow 4\,\ell \; .
\end{equation}
In the narrow-width approximation, this rate depends on the product of the production cross-section and the partial decay width of Higgs to $Z'$. As mentioned before, Higgses in the bare-bones model carries all the usual SM interactions with approximately SM-like strength couplings. Hence, the production cross-section of the bare-bones Higgs is essentially the same as the SM Higgs production cross section,
\begin{equation}
\label{eq:h-prod}
\begin{split}
\sigma \left( gg \rightarrow h \right) \big|_\text{bare-bones}
\
\simeq \ &
\sigma \left( gg \rightarrow h \right) \big|_\text{SM} \; .
\end{split}
\end{equation}
The partial decay width of Higgs to $Z'$ depends on the coupling $g_0$ in Eq.~\eqref{int-hz'z'}. The strongest constraint on the decay width, and hence on $g_{0}$, comes from the Tevatron. Even though CDF and D0 are not currently looking for $h \rightarrow 4\, \ell$ for a light Higgs, both experiments do have inclusive and isolated $4\,\ell$ searches, primarily to measure the $ZZ$ cross-section at the Tevatron. The CDF search at $5\,~\text{fb}^{-1}$ is public~\cite{CDF:10238}. They see $6$ candidate events with $4$ isolated leptons, among which $4$ events are selected to be good candidates for the decay product of $ZZ$. One of the remaining events fits the profile of a $ZZ\rightarrow 2\, \ell + 2\, \tau$ event, where both the $\tau$s have decayed leptonically. The last remaining event is {\it curious}, since at first glance it does appear to have stemmed from a couple of promptly-decaying resonances with a mass of $\sim\,55~\text{GeV}$. Although a single event can by explained away by a possible mis-recombination of leptons, it remains to be seen whether the early runs of LHC and even CDF/D0 data at larger luminosity finds similar anomalous events in the same mass-window.
In this work, we impose bounds on $\sigma_h \times \text{Br}\left( h \rightarrow Z'Z' \right)$ based on null observation of candidate $4\, \ell$ events. The maximum number of allowed $4\, \ell$ events at $95\%$ confidence level ($3.0$) is translated into a cross section bound by equating the $ZZ$ cross section measured by CDF ($1.7^{+1.2}_{-0.7}\, \pm 0.2~\text{pb}$) with four $4\, \ell$ events~\cite{CDF:10238}. At $95\%$ confidence level
\begin{equation}
\label{eq:cdf-lim-br}
\begin{split}
\sigma_h \times \text{Br}\left( h \rightarrow Z'Z' \right) \leq \frac{3}{4} \:
\sigma_{ZZ} \: \frac{\text{Br}(Z \rightarrow 2\ell)^2 }{\text{Br}(Z' \rightarrow 2\ell)^2 }\,
\mathcal A_\text{rel},
\end{split}
\end{equation}
where we have divided by the square of the $Z' \rightarrow \ell^+\ell^-$ branching fraction of Eq.~\eqref{eq:5}%
\footnote{This procedure assumes that the Higgs width is small. However, for the range of Higgs masses and parameters we are considering this is a perfectly valid approximation.}. The factor $\mathcal A_{rel}$ contains the ratio of acceptances for $ZZ$ events compared to $Z'Z'$ events. We divide this ratio into three factors:
\begin{equation}
\mathcal A_\text{rel} = a_\text{fiducial} \times a_{4\,\ell-\text{cut}} \times a_\text{id}.
\label{eq:effs}
\end{equation}
The $a_\text{fiducial}$ factor designates how often all four leptons from $ZZ \rightarrow 4\,\ell$ events pass CDF's tight kinematic criteria ($p_T > 10~\text{GeV}, \ |\eta| < 1.1, \text{ and } \Delta R_{\ell\ell} > 0.4$) compared to leptonic $Z'Z'$ events. As the $Z$ and $Z'$ have different masses and different production mechanisms, there is slight difference in kinematic acceptance. The second factor $ a_{4\,\ell-\text{cut}} $ is the ratio of acceptance for the total lepton invariant mass cut, $m_{4\,\ell} < 300~\text{GeV}$%
\footnote{ $m^2_{4\,\ell} = (\sum_{i=1}^4 p_{\ell_i})^2\; .$}
, while the final factor $a_\text{id}$ accounts for any difference in identifying/reconstructing leptons ({\em after} passing all kinematic cuts) related to the origin of the lepton. We take $a_{id}$ to be $1$ in this work.
The total SM $ZZ \rightarrow 4\,\ell$ acceptance can be found in Ref.~\cite{CDF:10238}. However, the quoted fraction ($\sim 0.11$) contains the id efficiency, which we have just argued should cancel in a ratio such as Eq.~\eqref{eq:effs}. Therefore, to estimate $a_\text{fiducial}$ and $a_{4\,\ell-\text{cut}}$, we resort to Monte Carlo. We generate samples of $p\bar p \rightarrow ZZ \rightarrow 4\,\ell$ and $p \bar p \rightarrow h \rightarrow Z'Z' \rightarrow 4 \ell$ with MadGraphV4~\cite{Alwall:2007st} + PYTHIA6.4~\cite{Sjostrand:2006za}. These events are passed through the fast-detector simulator PGS~\cite{PGS}, then analyzed. Following this procedure, we find a kinematic acceptance of $\sim 0.2$ for $ZZ$ and $\sim 0.15$ for $h\rightarrow Z'Z'$. In the surviving events we next form the total lepton invariant mass and impose the cut $m_{4\,\ell} < 300~\text{GeV}$. Roughly $\sim 86\% $ of SM events pass this cut, while -- given that the four leptons sum to $m_h$ in the $Z'$ case, all $Z'$ events pass. The Higgs mass used in these samples was $120~\text{GeV}$, though we expect identical numbers for all $m_h \lesssim 125~\text{GeV}$\@%
\footnote{ Numbers produced above assume relatively heavy $Z'$. For $M_{Z'} \lesssim 15~\text{GeV}$, a significant numbers of decayed leptons are not isolated. As a consequence, the acceptance of $Z'Z'$ events decrease drastically with decreasing $M_{Z'}$.}.
\begin{figure*}[!ht]
\includegraphics[height=2.75in]{zprime_discovery}
\hspace{0.5cm}
\includegraphics[height=2.75in]{hzz4ell}
\caption{In the left pane we display LHC $h \rightarrow Z'Z' \rightarrow 4\,\ell$ events in a two-dimensional plane $M_{\ell\ell,1}, M_{\ell\ell,2}$ where $M_{\ell\ell,1}$ ($M_{\ell\ell,2}$) is the lepton pair with highest $p_T$ (subleading $p_T$), and we stack signal and background events on top of each other. The cluster of events at $\sim 55~\text{GeV}$ shows the prominence of the $Z'$ signal (color online). For the same events, we plot the total $4$ lepton invariant mass in the right-hand pane. Both signal (red) and background (green) events have been generated with MadGraph. This figure is analogous to Figure.~12 in Ref~\cite{CDF:10238} created for $4\,\ell$ events at the Tevatron. }
\label{fig:fourlep}
\end{figure*}
Combining these factors, our derived value of $\mathcal A_\text{rel}$ is $1.15$. Plugging in to Eq.~\eqref{eq:cdf-lim-br}, we find that
\begin{equation}
\label{eq:limit-h-cs}
\sigma_h \times \text{Br}\left( h \rightarrow Z'Z' \right) \ \lesssim \ 84~\text{fb} \quad @95\% \text{CL}
\end{equation}
Dividing by the Higgs production cross section at $\sqrt s = 1.96~\text{TeV}$ (at NNLO~\cite{Anastasiou:2008tj}), we translate the cross section bound into a (Higgs mass dependent) limit on $\text{BR}(h \rightarrow Z'Z')$:
\begin{equation}
\text{BR}(h \rightarrow Z'Z') \leq \ 7.7\times 10^{-2} \quad \text{for } m_h = 120~\text{GeV}\; .
\label{eq:brlimits}
\end{equation}
We can immediately turn these branching fraction limits into predictions for the LHC. We simply reverse the process used to obtain $BR(h \rightarrow Z'Z')$, replacing the Tevatron Higgs production cross section with the most recent calculation of $\sigma(pp \rightarrow h)$ at $\sqrt s = 7~\text{TeV}$~\cite{Dittmaier:2011ti}. While this gives us the net number of four-lepton events, we still need to account for acceptance. We estimate the acceptance using Monte-Carlo events, generated as before. Events are kept if they contain four isolated leptons with $p_T > 10~\text{GeV}, |\eta| < 2.5$, $\Delta R_{\ell\ell} > 0.4$. We check further that, out of these $4\,\ell$ events, there are two pairs of same-flavor, opposite sign leptons. After this procedure, we find that roughly $\epsilon_{\ell} \sim 28\%$ of signal four-lepton events survive. Combining these factors, we find:
\begin{equation}
\label{eq:lhc-h}
\begin{split}
\sigma(pp \rightarrow Z'Z' \rightarrow 4\,\ell) _{\text{LHC}@7} & \simeq \\
\sigma(pp \rightarrow h)_{\text{LHC}@7} \times \text{BR} &
\left( h \rightarrow Z'Z' \rightarrow 4\,\ell\right) \times \epsilon_{\ell} \\
& \lesssim \ 28~\text{events } ~\text{fb}^{-1} \; .
\end{split}
\end{equation}
The $Z'Z' \rightarrow 4\,\ell$ events have essentially no standard model background. Di-boson production, $pp \rightarrow ZZ $ followed by $ZZ \rightarrow 4\,\ell$, is the only SM process that results in $4$ high-$p_T$ leptons produced in the primary interaction. Of the di-boson backgrounds, $Z\rightarrow e^+ e^-$ and $Z\rightarrow \mu^+\mu^-$ can be easily recognized by looking at the invariant mass of the lepton pair, $M_{\ell\ell}$. Provided $M_{Z'} \ne M_Z$, these background events will be well-separated from the signal. Leptonically decaying $\tau^{\pm}$ pairs will have a continuous spectrum in $M_{\ell\ell}$ and could be problematic. However, the rate for this process is additionally suppressed by leptonic branching fractions of the $\tau$ and by cut acceptance. Each of the $\tau$ daughters has only a fraction of the $\tau$ momentum, so daughter leptons are less likely to pass the basic kinematic cuts than $e/\mu$ coming directly from the $Z$.
Other, reducible backgrounds in this channel come from the mis-identification of jets as leptons. Using the detector simulation PGS to model lepton-misidentification, we have checked the contributions in the $4\,\ell$ final state from $Z(\ell\ell) + \text{jets}$, $WW/WZ$, and leptonic $t\bar t + \text{jets}$. We find that none of these leave a trace in the signal region. Given that the probability for a jet to fake a lepton is, very conservatively, $0.1\%$, the small size of the reducible background is not completely surprising.
Combining the backgrounds with the signal, we show all $4\,\ell$ events in Fig.~\ref{fig:fourlep}. To generate the signal, we use a sample point, $m_h = 120~\text{GeV}, M_{Z'} = 55~\text{GeV}$, and $BR(h\rightarrow Z'Z') = 7.7\%$. We first plot events in the $M_{\ell\ell,1}-M_{\ell\ell,2}$ plane, where $M_{\ell\ell,1} \text{ and }M_{\ell\ell,2}$ correspond to the mass of the lepton pair with leading and subleading $p_T$ respectively.
Fig.~\ref{fig:fourlep} clearly shows the cleanliness of the $Z'$ signal. For our choice of $Z'$ mass the leptonic $Z$ background events are well separated from the signal and the total $4\,\ell$ invariant mass for each event finds the Higgs peak cleanly.
Both the Higgs and $Z'$ signals are extremely clean and clearly spectacular. By optimizing the lepton acceptance (by, for example, admitting lower $p_T$), it may be possible to increase the rate even further. While the signal above shows a $h\rightarrow Z'Z'$ branching fraction right on the edge of the Tevatron bound, slightly lower branching fractions would still lead to shockingly early light Higgs discovery. Even with a $h\rightarrow Z'Z'$ branching fraction as low as $1.4\%$, we expect $\sim 5$ signal events within $1~\text{fb}^{-1}$ of $7~\text{TeV}$ LHC running -- discover far sooner than with traditional light-Higgs modes like $\gamma\gamma$ or $\tau^+\tau^-$. Note that the signals and rates shown in this section apply to promptly decaying $Z'$ only. For Higgses decaying to long-lived $Z'$s, a completely different set of signals, equally spectacular to those in the prompt case, are possible. We explore this case later in this section.
\subsection{$4\,\ell$ with Missing Energy }
\label{sec:higgsino-pheno}
\begin{figure*}
\includegraphics[height=2.75in]{susy_zprime}
\hspace{0.5cm}
\includegraphics[height=2.75in]{susy_mt2}
\caption{$Z'$ signal (left panel) and the $M_{T_2}$ distribution (right panel) for the benchmark supersymmetry point. We have assumed a $7~\text{TeV}$ collider and an integrated luminosity of $5~\text{fb}^{-1}$ (color online). The final state is $4\,\ell + \slashchar E_T$, which has no sizable standard model backgrounds. We have imposed a missing energy cut $\slashchar E_T > 20~\text{GeV}$, which almost completely removes the $Z$ events from the plots. The number of events is smaller than in the Higgs case because we have to produce two neutralinos, rather than just one Higgs, to get the final state we want. The $M_{T_2}$ distribution was created using the true value of $M_{\tilde S}$. The distribution begins at $\sim M_{\tilde S} + M_{Z'}$ and has a drop-off at $\sim 160~\text{GeV} \sim M_{\tilde N}$. The shape of $m_{T_2}$ -- something resembling a peak rather than a turn-on plus edge -- is an artifact of the small mass difference between $\sim M_{\tilde N}$ and $M_{\tilde S} + M_{Z'}$. The tail of the distribution comes from events where a $Z'$ was attributed to the incorrect parent neutralino. The small clump of events around $180~\text{GeV}$ comes from $ZZ$ background which survived the $\slashchar{E}_T$ cut. }
\label{fig:susy_signs}
\end{figure*}
Requiring large missing energy, in addition to $4\,\ell$, gives us the ultimate clean signal. The SM $ZZ \rightarrow 4\,\ell$ events which constitute the only background in $4\,\ell$ channel do not survive a sufficiently large $\slashchar{E}_T$ cut. Such a clean environment is the ideal laboratory to study complicated signals, such as from supersymmetry. In the MSSM, it is hard to utilize this clean $4\,\ell + \slashchar{E}_T$ environment as only a small fraction of signal events have the required ingredients (there, however, exists a class of models with suitably designed mass hierarchies that result in multi-lepton and missing energy signals~\cite{DeSimone:2008gm,DeSimone:2009ws,Katz:2009qx,Katz:2010xg}).
However, once the MSSM is extended by a Higgsphilic $Z'$, many avenues to $4\,\ell + \slashchar{E}_T$ open up instantly. In this section, we investigate the discovery potential of the bare-bones model in the $4\,\ell + \slashchar{E}_T$ channel, concentrating on the following topology
\begin{equation}
\label{eq:susy-topo}
p p \text{ or } p \bar{p} \rightarrow \tilde N \tilde N \rightarrow Z' Z' + \tilde S \tilde S \rightarrow 4\,\ell + \slashchar{E}_T
\end{equation}
As in previous subsections, we first use the Tevatron data ($p \bar{p}$ initial state) to constrain the production cross-section of $4\,\ell + \slashchar{E}_T$, then discuss discovery potential at the LHC in the allowed region.
There is no dedicated $4\,\ell\,+\,\slashchar E_T$ search at the Tevatron. The closest search is Ref.~\cite{CDF:10238}, the same $4\,\ell$ analysis we used to bound $h\rightarrow Z'Z'$ from Higges. While this search focused on exclusive $4\,\ell$ states, events with missing energy were not vetoed, so this search also limits $Z'$ produced from the decay of supersymmetric particles.
Using the same logic as in the $h\rightarrow Z'Z'$ section, we can bound the production cross section of certain supersymmetric particles. Assuming that no events have been found at the Tevatron with $5~\text{fb}^{-1}$ of data, we find that (as usual, at $95\%$CL)
\begin{equation}
\label{eq:susy-bound}
\begin{split}
\sigma \left(p\bar p \rightarrow \tilde N \tilde N \rightarrow 4\,\ell + \slashchar{E}_T \right) \ & \leq
\ 5.7~\text{fb} \left( \frac{ \mathcal{A}_{\text{sm}} } { \mathcal{A}_{\text{susy}} }\right) \\
& \simeq 4.3~\text{fb} \; .
\end{split}
\end{equation}
where the value $5.7~\text{fb}$ comes directly from Eq.~\eqref{eq:cdf-lim-br} once the $BR(Z'\rightarrow \ell\ell)$ factors are removed. The $\mathcal A_{\text{susy}}$ in Eq.~\eqref{eq:susy-bound} accounts for any difference in acceptance between SM $Z \rightarrow \ell\ell$ events and supersymmetric $Z'$ events. We have evaluated $\mathcal{A}_{\text{sm}} / \mathcal{A}_{\text{susy}} $ using a set of events generated with a MadGraph implementation of the bare-bones model. We find the basic kinematic acceptance of supersymmetric $Z'$ events to be slightly higher ($23\%$) than for Higgs-induced $Z'$ while the $m_{4\,\ell}$ cut remains approximately equally efficient.
Let us assume that $ M_{Z} > ( M_{\tilde N} - M_{\tilde S} ) > M_{Z'}$. This choice of parameters ensures that $\tilde N \rightarrow \tilde S + Z'$ is almost $100\%$. It is particularly useful because, in this case, bounds on $4\,\ell + \slashchar{E}_T$ can be directly converted to bounds on $\tilde N$ masses. If the branching fraction of $\tilde N$ to $Z'$ is less than $100\%$, all our constraints are on production cross-section times the branching fraction.
\begin{equation}
\begin{split}
\label{eq:N-bound}
\sigma \left( p\bar p \rightarrow \tilde N \tilde N \right) &\times \Big( \text{Br}\left( Z' \rightarrow 2\ell\right)\Big)^2 \: \leq 4.3~\text{fb} \\
\Rightarrow \quad & \quad M_{\tilde N} \gtrsim 151~\text{GeV}.
\end{split}
\end{equation}
If the splitting among $\tilde N$ and $\tilde C$ is small, as is often the case, Eq.~\eqref{eq:N-bound} also constrains $M_{\tilde C}$.
Another Tevatron measurement that is often used to constrain the color-neutral sector of supersymmetric models is the bound on tri-lepton/di-lepton + track production~\cite{CDF:9817}. This limit is most constraining when the sleptons are light and charginos/(heavier) neutralinos decay to $\tilde{\ell} + \nu/\tilde{\ell} + \ell$, essentially $100\%$ of the time. In the bare-bones model, $\tilde N-\tilde C$ decay to gauge bosons rather than to sleptons. Subsequent decays of gauge bosons do yield leptons sometime, but the branching fractions of $W/Z/Z'$ to leptons suppresses the $3\,\ell$ rate to well below the current limit as established in Eq.~\eqref{eq:N-bound}.
Having established the Tevatron constraints on sparticle masses, we now show an example of bare-bones supersymmetry production at the LHC. The benchmark spectra we choose is
\begin{equation}
\label{eq:bench}
M_{Z'} = 55~\text{GeV} \quad M_{\tilde S} = 90~\text{GeV} \quad M_{\tilde N} = 160~\text{GeV} \; .
\end{equation}
The results are shown in Fig.~\ref{fig:susy_signs}. This point was generated using the selection criteria we used for the Higgs-induced $Z'$ signal at the LHC in Sec.~\ref{sec:higgs_prompt} plus an additional missing energy requirement $\slashchar E_T > 20~\text{GeV}$. While unnecessary for $Z'$ discovery, the $\slashchar E_T$ cut is useful for extracting information about the superpartners. As in the Higgs case (Fig.~\ref{fig:fourlep}), we show the $Z'$-pair candidates in the $M_{\ell\ell,1}$ vs. $M_{\ell\ell,2}$ plane. However we show signal and background events for $5~\text{fb}^{-1}$ of data at $7~\text{TeV}$ rather than $1~\text{fb}^{-1}$. More luminosity is necessary since the supersymmetric signal is smaller -- we need to produce two neutralinos to get two $Z'$ instead of producing just one Higgs. As each event contains two invisible particles, we cannot simply combine all leptons to reconstruct the parent Higgsinos. However, we can get some idea on the mass scale of $M_{\tilde N}$ by using transverse variables. The second plot of Fig.~\ref{fig:susy_signs} shows the generalized transverse mass variable $M_{T_2}$~\cite{Lester:1999tx, Barr:2003rg, Cheng:2008hk} for these events. The generalized transverse mass depends on the mass of the invisible particle ($M_{\tilde S}$ for the benchmark point), however for simplicity we assume $m_{\tilde S}$ is already known in our calculations. With this additional assumption, the endpoint of $M_{T_2}$ gives the mass of $\tilde N$. If $M_{\tilde S}$ is not known, we could still get some insight into $M_{\tilde N}$ by scanning over a range of $M_{\tilde S}$ values and looking for kink features in the $\text{max}(M_{T_2})-M_{\tilde{S}}$ plane~\cite{Barr:2003rg}. More involved techniques will certainly require larger datasets. However, because we are using $M_{T_2}$ in a purely leptonic environment, this endpoint and/or kink features are less susceptible to initial state radiation or experimental smearing than in hadronic applications.
\subsection{Signatures of long lived particle}
\label{sec:long-lived}
Having surveyed the signals for prompt $Z'$ from Higgs and supersymmetric decays, in this section we address how these signals are modified when the $Z'$ is long-lived on collider scales. Before we begin, two comments are in order. First, the displaced bursts of energy left by a long-lived $Z'$ are highly exotic and free of any background; discoverability is limited solely by how often and how reliably such bursts can be triggered. Therefore, for this section we are not restricted to the leptonic decays of $Z'$. Second, exploring prospects of discovery of Higgs via long-lived $Z'$ is {\em highly} detector dependent and a full simulation is necessary to extract a completely reliable number. Such elaborate work is beyond the scope of this work. However, an approximate estimation of the number of triggerable events in the ATLAS is feasible based on the report Ref.~\cite{Aad:1175196}. \\
\begin{figure}[!ht]
\includegraphics[width=0.45\textwidth]{lifetime.pdf}
\caption{ Contours showing the $Z'$ lifetime (in meters) as a function of the mixing parameter $\epsilon$ and the mass of $Z'$. }
\label{fig:life-time}
\end{figure}
The total decay width of $Z'$, and, therefore, the lifetime $\tau_{Z'}$ is sensitive on the value of $\epsilon$. As long as $\epsilon \gtrsim 10^{-7}$, its decay is mostly prompt and the conventional collider studies in previous sections apply. For a smaller $\epsilon$, $Z'$ can be considered as long-lived and will decay at some macroscopic, displaced distance from the primary interaction. Since we are interested in decays within the detector volume, we are restricted to $\epsilon \gtrsim 10^{-9}$, corresponding to a decay length of $c \tau_{Z'} \lesssim 20~\text{m}$. This number comes from Ref.~\cite{Aad:1175196} and was based on the requirement that a sizable fraction of $Z'$ decays are subject to various triggers designed to find hidden valley particles~\footnote{Specifically, at $c \tau_{Z'} \lesssim 20~\text{m}$, at least $20\%$ of $Z'$ decays occur before the first Muon Spectrometer trigger plane in the ATLAS detector.}. For even smaller mixing (or longer decay lengths), $Z'$s mostly decay outside the detector. Occasionally, a $Z'$ with $c\tau_{Z'} \gg 20\,m$ will decay inside the detector, though any realistic chance of seeing these events involves analyzing a much bigger sample of data than the first few inverse femtobarn samples.
In Ref.~\cite{Aad:1175196}, a Higgs which decays mostly to $4\,b$ via a long-lived hidden sector particle was studied, and the article summarizes the performance and efficiency of three triggers: $(i)$ the Muon RoI cluster trigger $(ii)$ the $\log_{10}(E_\text{had}/E_\text{em})$ trigger, and $(iii)$ the ID-Trackless-jet+Muon trigger. The trigger are each designed to find sudden bursts of activity in different areas of the detector: The Muon RoI trigger looks for activity near the end of the HCal or in the Muon Spectrometer, the $\log_{10}(E_\text{had}/E_\text{em})$ trigger looks at jets from decays inside the calorimeters, and the ID-Trackless-jet+Muon trigger looks for jets stemming from the Inner Detector beyond the first pixel layer and contains muons in the cone ($b$ or $c$ jets).
\begin{figure}[!ht]
\centering
\includegraphics[width=3.0 in]{long-lived_to_jets.pdf}
\caption{ Number of $h \rightarrow Z'Z' \rightarrow \text{hadrons}$ events to be captured by the existing long-lived triggers in the ATLAS detector. The red(solid), black(dot-dashed), and blue(dashed) lines refer to events captured by the $\log_{10}(E_\text{had}/E_\text{em})$ trigger, the ``muon RoI cluster trigger'' and the ``ID-trackless-jet + muon trigger'' respectively. We have assumed the Higgs mass to be $120~\text{GeV}$ and BR$(h \rightarrow Z'Z')$ to be $100\%$. }
\label{fig:longlive}
\end{figure}
To estimate the number of triggerable events from the study done in Ref.~\cite{Aad:1175196} we concentrate only on the hadronic decays of $Z'$ as well as on the $Z'$ decays to hadronic $\tau$s. We also assume that $c-$hadrons give rise to muons at the same rate as the $b-$hadrons. In Fig.~\ref{fig:longlive}, we have plotted the number of triggerable events by various trigger objects at the ATLAS detector for $m_h = 120~\text{GeV}$ and BR$(h \rightarrow Z'Z')$ to be $100\%$ at the LHC running at $\sqrt{s} = 7~\text{TeV}$. The ID-trackless-jet + muon trigger is the least efficient trigger since $c$ or $b$ hadrons decay to muons only $15-20\%$, even though $Z'$ decays to $c$ or $b$ $\sim 27\%$.
We do not give a similar estimation $Z'$ decaying to leptons because of the lack of a detailed study as performed in Ref.~\cite{Aad:1175196}. Assuming a flat trigger rate of $20\%$, we find that around $1300$ events per inverse$~\text{fb}$ will be triggered. However, a dedicated simulation is definitely warranted in order to estimate any realistic number.
\section{From Viable Gravity Mediation to the Bare-bones Model}
\label{sec:viable-grav-medi}
Having surveyed how supersymmetric and Higgs phenomenology are altered by a Higgsphilic $Z'$, in this section we connect the bare-bones model to a particular limit of the viable gravity-mediated model of Ref.~\cite{Kribs:2010md}.
\subsection{Viable Gravity Mediation in a Nutshell }
\label{sec:spectr-viable-grav}
Firstly, let us point out the key features in the spectrum of the viable mediation that distinguishes it from a typical MSSM study point.
\begin{itemize}
\item Gauginos are Dirac: they acquire masses along with fermionic components of superfields in the adjoint representation of the corresponding gauge group
~\cite{Polchinski:1982an,Hall:1990hq,Randall:1992cq,Fox:2002bu,Nelson:2002ca,Chacko:2004mi,Carone:2005iq,Nomura:2005rj,Nomura:2005qg,Carpenter:2005tz,Antoniadis:2006uj,Nakayama:2007cf,Amigo:2008rc}.
\item Soft squared-masses of the scalar superpartners of the SM fermions are {\emph not} flavor diagonal. There are neither $a-$type terms nor $\mu-$type Higgsino mass terms and, as a consequence, there are no left-right mixings among scalar soft-masses~\cite{Kribs:2007ac,Kribs:2009zy,Fok:2010vk}.
\item Gauginos are typically heavier than all other scalars.
\item The extended Higgs sector is charged under the $\text{SU(2)}_L \times \text{U(1)}_Y \times \text{U(1)}_X$ gauge groups which are spontaneously broken down to the $\text{U(1)}_\text{em}$ at the electroweak scale. The resultant spectrum contains the usual $W_\pm, Z$ and an additional massive charge neutral vector boson $Z'$.
\end{itemize}
In the remaining part of this subsection we discuss the Higgs sector in detail as we show later that the effective description of it gradually becomes the bare-bones model constructed in Sec.~\ref{sec:model}. Below we have listed all the particles in the Higgs sector along with their charges:
\begin{equation}
\begin{tabular}{ c|c|c|c|c }
& $\text{SU(3)}_C$ & $\text{SU(2)}_L$ & $\text{U(1)}_Y$ & $\text{U(1)}_X$ \\
\hline
$H_{u,d}$ & ${\bf 1}$ & ${\bf 2}$ & $\pm 1/2$ & $0$ \\
$R_{u,d}$ & ${\bf 1}$ & ${\bf 2}$ & $\mp 1/2$ & $\pm 1$ \\
$S_{u,d}$ & ${\bf 1}$ & ${\bf 1}$ & $0$ & $\mp 1$ \\
$T_{u,d}$ & ${\bf 1}$ & ${\bf 1}$ & $0$ & $\pm 2$ \\
\hline
\end{tabular} \; .
\label{H-charges}
\end{equation}
We use capitalized letters to designate the complete chiral supermultiplets, small letters for the scalar component, and capital letters with tildes for the fermionic components of the same supermultiplet.
The superpotential in the model consistent with the charges listed in eq.~\eqref{H-charges}
\begin{equation}
\begin{split}
W \ \supset \
\alpha_u \: S_u R_u H_u \
+ \ \alpha_d \: S_d R_d H_d \ + \\
\qquad \qquad \qquad \frac{1}{2} \beta_u \: T_u S_u^2
\ + \ \frac{1}{2} \beta_d \: T_d S_d^2 \; ,
\end{split}
\label{eq:marginal-op}
\end{equation}
where we have neglected the usual Yukawa terms. The scalar potential contains the usual squared-soft masses for all multiplets as well as new $b-$type soft masses because of additional vector-like chiral supermultiplets.
\begin{equation}
\label{eq:int-higgs}
\begin{split}
V_{\text{soft} } \ \subset \ \sum_{\phi} \left( m_{\phi_u}^2 \: \phi_u^* \phi_u \ + \
m_{\phi_d}^2 \: \phi_d^* \phi_d \right) \\
- \ \sum_{\phi} \left( b_{\phi} \: \phi_u \phi_d \ + \ \text{c.c.} \right) \; ,
\end{split}
\end{equation}
where field $\phi_{u,d}$ runs over the set of fields $ \{ h_{u,d}, r_{u,d}, s_{u,d}, s_{u,d}\}$: the scalar components of the corresponding chiral supermultiplets. The Higgs potential is minimized around $\avg{h_{u,d}} \equiv v_{u,d} \neq 0$ and $\avg{s_{u,d}} \equiv v_{s_{u,d}} \neq 0$ and $\avg{r_{u,d}} = \avg{t_{u,d}} = 0$. As a result
the electroweak symmetry is broken down to electromagnetism ($\text{SU(2)}_L \times \text{U(1)}_Y \rightarrow \text{U(1)}_\text{em}$). The additional symmetry $\text{U(1)}_X$ is broken completely which results in a massive $Z'$.
\subsection{To the Bare-bones Model}
\label{sec:eff2bare-boness}
In order to extract the bare-bones model from the viable gravity mediation spectrum, let us consider the following set of limits:
\begin{equation}
\label{eq:hscalar-limits}
\begin{split}
g_X \ \ll \ g\, \text{ or } g' \; , & \quad M_{Z'} \ \lesssim M_{Z/W} \\
\alpha_{u,d} \ \simeq \ \beta_{u,d} \ & \simeq \ \left(0.\text{few } - 1 \right) \\
\tan \beta \equiv \frac{v_u}{v_d} \gtrsim 3 \; , & \quad \tan \beta_s \equiv \frac{v_{s_u}}{v_{s_d}} \sim 1 \; .
\end{split}
\end{equation}
The requirement of large $\tan \beta$ limit is more general than just to derive the bare-bones model. It is necessary in order to push the CP-even lightest Higgs boson above the LEP limit.
\subsubsection*{Gauge bosons, gauginos and matter superpartners}
\label{eff-gauge}
The $Z'$-mass is given by:
\begin{equation}
\label{eq:MZ'}
M_{Z'} \ = \ \sqrt 2\, g_X v_{s} \; , \quad \text{where} \quad
v_s^2 \ = \ \left( v_{s_u}^2 + v_{s_d}^2 \right) \; .
\end{equation}
The implication of a small $g_X$ in Eq.~\eqref{eq:hscalar-limits} is that $M_{Z'} \ll v_s$. In fact, since we seek $Z'$ at the electroweak scale, $v_s \gg M_Z, M_W$ and all degrees of freedom at or above $v_s$ can be easily integrated out.
Also, note that at one-loop level, kinetic mixing of $\text{U(1)}_Y$ and $\text{U(1)}_X$ is inevitably induced from the loops of $R_{u,d}$ which carry both charges. This generates the operator shown in Eq.~\eqref{kin-mix}. The mixing parameter $\epsilon$ in the viable gravity mediation is given by
\begin{equation}
\label{eq:epsilon}
\epsilon \simeq 9 \times 10^{-4} \left(\frac{g_X}{0.1} \right) \log \frac{\Lambda}{1~\text{TeV}} \;
\end{equation}
where $\Lambda$ denotes the scale above which there is no mixing (for example, the scale at which a non-Abelian group is broken to $\text{U(1)}_X$).
The Dirac nature of gauginos has a profound implication in the squared-masses of sleptons and squarks: the scalar masses do not get log-enhanced contributions due to the gaugino masses at one loop. As a result, the pattern of superpartner masses is quite different from the typical MSSM spectra. Gluinos are the heaviest particles and are significantly heavier than the rest of the superpartners. Other gauginos have mass comparable to the squark and slepton masses, which, in turn, are of the order of scalar vev $v_s$. We therefore integrate out all gauginos, squarks and sleptons from the electroweak effective theory.
\subsubsection*{Scalar Higgs Multiplets}
Next, we turn our attention to the scalars in the Higgs sector. The scalars in the superfields $R_{u,d}$ or $T_{u,d}$ do not participate in the breaking of any symmetries and get physical masses simply from positive mass-squareds $m^2_{r_u, r_d}$ and $m^2_{t_u, t_d}$ respectively. We can easily decouple them along with squarks and sleptons.
Analysis of the neutral degrees of freedom in $H_{u,d}$ or $S_{u,d}$ is more involved, since all four states get non-zero vevs and mix among themselves. The gauge-eigenstate fields can be expressed in terms of the mass eigenstate fields as:
\begin{equation}
\label{eq:h-expn}
\begin{split}
\begin{pmatrix}
h_u^0 \\ h_d^0 \\ s_u \\ s_d
\end{pmatrix}
\ = \ \begin{pmatrix}
v_u \\ v_d \\ v_{s_u} \\ v_{s_d}
\end{pmatrix} &
\ + \ \frac{1}{\sqrt 2}\ \mathcal{O}_1
\begin{pmatrix}
h \\ H \\ s \\ S
\end{pmatrix}
\\ + \ \frac{i}{\sqrt 2} \ & \mathcal{O}_2
\begin{pmatrix}
G_0 \\ A_h \\ 0 \\ 0
\end{pmatrix}
+ \ \frac{i}{\sqrt 2}\ \mathcal{O}_3
\begin{pmatrix}
0 \\ 0 \\ G'_0 \\ A_s
\end{pmatrix} \; .
\end{split}
\end{equation}
In the above, $ \mathcal{O}_{1,2,3}$ are orthogonal rotation matrices, $G_0$ and $G'_0$ are the would-be Nambu- Goldstone bosons, which become the longitudinal modes of the $Z$ and $Z'$ massive vector bosons respectively, $A_h, A_s$ are the CP-odd states and finally $h,H,s,S$ are the CP-even states.
In the limits of Eq.~\eqref{eq:hscalar-limits}, one also finds that the masses of the pseudo-scalars ($A_h, A_s$) are greater than or in the order of $v_s$, and there is minimal mixing between $h-$type states and $s-$type states. The mixing matrix can very well be approximated to be:
\begin{equation}
\label{eq:defn-o}
\mathcal{O}_1 \sim
\begin{pmatrix}
\cos \alpha & \sin \alpha & X & X \\
-\sin \alpha & \cos \alpha & X & X \\
X & X & \cos \alpha_s & \sin \alpha_s \\
X & X & -\sin \alpha_s & \cos \alpha_s
\end{pmatrix} \; ,
\end{equation}
where by $X$ we represent small parameters, the squares of which can be disregarded. As a result of Eq.~\eqref{eq:hscalar-limits} the spectrum of CP-even scalars contains only one field ($h$) with mass lighter than $M_Z$ at the tree level. In the limit that $v_s$ is large while the $\alpha$ and $\beta$ parameters in Eq.~\eqref{eq:hscalar-limits} are $\mathcal{O}(0.\text{few})$, both the scalars $s \text{ and }S$ are heavy. Similarly, a large mass of the pseudo-scalar $A_h$ implies a large mass for $H$.
One implication of the large $A_h$ mass and the specific form of the rotation matrix $\mathcal{O}_1 $ in Eq.~\eqref{eq:defn-o} is that the lightest CP-even Higgs scalar is similar to the MSSM Higgs in the decoupling limit - all its coupling to the SM fermions and the electroweak gauge bosons are SM like to the leading order. We additionally obtain the operator in Eq.~\eqref{int-hz'z'}, because of the non-vanishing $3\!\!-\!\!1$ and $4\!\!-\!\!1$ elements of $\mathcal{O}_1$ (call them $\delta_1$ and $\delta_2$ respectively) with the coefficient:
\begin{equation}
\label{eq:g0}
g_0 \ =\ g_X ( \delta_1 \sin \left( \beta_s \right) + \delta_2 \cos \left( \beta_s\right))
\end{equation}
Turning to the charged Higgses, in the limit of Eq.~\eqref{eq:hscalar-limits}, they generically have masses similar to that of the pseudo scalar $A_h$ and, hence, decouple. In case of large mixing one of the charged scalars might become light in the order of the EWSB scale. However, for the kind of phenomenology we discuss, this is mostly irrelevant.
\subsubsection*{Charged and Neutral Higgsinos}
Lastly, we come to the charginos and neutralinos. Expanding the superpotential in Eq.~\eqref{eq:marginal-op} around the Higgs vevs, we find the chargino masses to be:
\begin{equation}
\label{eq:chr-mass}
\alpha_u \: v_s \sin \beta_X \: R_u^{-} \tilde H_u^{+} \
+ \ \alpha_d \: v_s \cos \beta_X \: H_d^{-} R_d^{+} \; .
\end{equation}
Note that if either of the product $\alpha_u \sin \beta_X$ or $\alpha_d \cos \beta_X$ is bigger than the other, one of the chargino decouples. Without loss of generality we assume that
$\alpha_u \: \sin \beta_X \ll \alpha_d \cos \beta_X$ and we integrate out the charginos
$H_d^{-}-R_d^{+}$. The Dirac chargino $\tilde C$ in our bare-bones model is, therefore, represented by
\begin{equation}
\label{eq:defn-C}
\tilde C \ = \
\begin{pmatrix}
\tilde H_u^+ \\ \tilde R_{u}^{- *}
\end{pmatrix} \; .
\end{equation}
The neutral Higgsino sector is slightly more complicated, since we need to deal with eight (two component) neutralinos. However, because of the pure Dirac nature of the neutralinos, the mass matrix breaks up into $4\times 4$ blocks. Each $4\times 4$ neutralino mass matrix can be further reduced to $2\times 2$ blocks since the $U-$type and $D-$type spinors are totally disconnected%
~\footnote{ The $ U\leftrightarrow D$ symmetry in the superpotential shown in Eq.~\eqref{eq:marginal-op} results in $U$ and $D$ sector being completely disjoint to the leading order. This is however not true at higher orders:
because of Higgs vevs $v_{u,d}$, a tiny mixing between the $U$ and $D$ sector of order $g_i^2 v_{u}v_{d}/M_i^2$ gets generated, where $M_i$ refers to the gaugino mass of the $i-$th gauge group.}.
Before proceedings let us specify the full neutralino mass matrix:
\begin{equation}
\label{eq:neut-mass}
\begin{bmatrix}
\tilde H_u^0 & \tilde S_u^0
\end{bmatrix} \!\!
\begin{bmatrix}
\alpha_u \: v_s \sin \beta_X & 0 \\
\alpha_u \: v \sin \beta & \beta_u \: v_s \sin \beta_X
\end{bmatrix} \!\!
\begin{bmatrix}
\tilde R_u^0 \\ \tilde T_u^0
\end{bmatrix} + u\rightarrow d
\end{equation}
In the limit $\alpha_u \: \sin \beta_X \ll \alpha_d \cos \beta_X$ and $\beta_u \: \sin \beta_X \ll \beta_d \cos \beta_X$, entire $d-$sector decouples and we need to worry about only four neutralinos $\tilde H_u, \tilde R_u, \tilde S_u, \tilde T_u$. Also note that in the limit $v_s \gg v$, (implied by Eq.~\eqref{eq:hscalar-limits}) there is minimal mixing between the $\tilde H- \tilde R$ states and $\tilde S - \tilde T$ states, and the mixing can be treated as perturbation. The Dirac neutralinos $\tilde N$
and $\tilde S$ in the bare-bones model are then simply given as
\begin{equation}
\label{eq:defn-NS}
\tilde N \ = \
\begin{pmatrix}
\tilde H_u \\ \tilde R_{u}^{*}
\end{pmatrix} \text{ and }\quad
\tilde S \ = \
\begin{pmatrix}
\tilde S_u \\ \tilde T_{u}^{*}
\end{pmatrix}\; .
\end{equation}
With the mapping established in Eqs.~\eqref{eq:defn-C} and \eqref{eq:defn-NS}, we can now derive the operators in Eq.~\eqref{int-nhz'} with the identification
\begin{equation}
\label{eq:matching-NS}
\begin{split}
\eta \ = \ \left( \frac{\alpha_u^2}{\beta_u^2 - \alpha_u^2} \right) \left( \frac{v \sin \beta}{v_s \sin \beta_X} \right) \\
\zeta \ = \ \left( \frac{\alpha_u \beta_u}{\beta_u^2 - \alpha_u^2} \right) \left( \frac{v \sin \beta}{v_s \sin \beta_X} \right)
\end{split}
\end{equation}
\subsection{A Numerical Example}
\label{sec:num}
In the previous subsections we have outlined the limits in which the bare-bones emerges as an effective theory of viable gravity mediation. Following Ref.~\cite{Kribs:2010md}, it is easy to check that the gauginos and the matter superpartners are heavier than the electroweak scale, and needs no additional supporting argument. It is slightly nontrivial to analyze neutral scalar Higgses (CP-even scalars $h,H,s,S$, CP-odd scalars $A_h, A_s$), neutral and charged Higgsinos, and then to recover the bare-bones model from it. To demonstrate the mapping discussed in Subsec.~\ref{sec:eff2bare-boness}, we find a point in the parameter space of the viable gravity mediation and show numerically that it reduces to the benchmark point in Eq.~\eqref{eq:bench}.
First of all, note that in order to fully specify the Higgs sector, we need four fewer parameters because of the four equalities that result from the minimization conditions $\partial V/ \partial v_{u,d}^0 = \partial V/ \partial v_{s_{u,d}} = 0$. We use them to eliminate four input parameters, $m_{H_{u,d}}^2$, $m_{S_{u,d}}^2$ in terms of other soft and supersymmetry breaking parameters.
Below we list the initial set of parameters that we use to calculate the spectrum:
\begin{equation}
\label{eq:param-in}
\begin{split}
& \qquad v_s \: = \: 700~\text{GeV} \>,\qquad g_X \: = \: 0.06 \, , \\
& \qquad \tan\beta = 1 \>,\qquad \tan\beta_X = 10 \, , \\
& B_{h} = (300~\text{GeV})^2 \>,\quad B_s = (300~\text{GeV})^2 \,, \\
& \alpha_u = 0.3 \>,\quad \alpha_d = 0.6 \>,\quad
\beta_u = 0.2\>, \quad \beta_d = 0.7 \, .
\end{split}
\end{equation}
Let us first start with the neutral scalar spectrum%
\footnote{Because of the Dirac nature of gauginos, the D-term contributions to the scalar quartic are usually suppressed. In this work, however, we are assuming large squared soft masses for the scalar adjoints so that the suppression is negligible and we recover MSSM-like quartic.}:
\begin{equation}
\begin{split}
\label{eq:bench-scalar-mass}
& \qquad \quad m_h = 61~\text{GeV} \>,\qquad m_H = 954~\text{GeV} \, , \\
& \qquad \quad m_s = 231~\text{GeV} \>, \qquad m_S = 512~\text{GeV} \, , \\
& \quad \quad m_{A_h} = 953~\text{GeV} \>, \qquad m_{A_s} = 424~\text{GeV} \, .
\end{split}
\end{equation}
The lightest CP-even Higgs $h$ obtain a tree level mass of only $61~\text{GeV}$. Large contributions to its mass arise due to radiative corrections from top-stop loop and neutral Higgs/Higgsino loop, which help it to avoid the LEP limit~\cite{Kribs:2010md}.
Next, consider the Higgsino spectrum:
\begin{equation}
\begin{split}
\label{eq:bench-ino-mass}
& \qquad \quad M_{\tilde S_u - \tilde T_u} = 91~\text{GeV} \>,\qquad M_{\tilde H_u - \tilde R_u} = 162~\text{GeV} \, , \\
& \qquad \quad M_{\tilde S_d - \tilde T_d} = 347~\text{GeV} \>,\qquad M_{\tilde H_d - \tilde R_d} = 296~\text{GeV} \, , \\
& \quad \quad M_{\tilde H_u^+ - \tilde R_u^-} = 148~\text{GeV} \>,\qquad
M_{\tilde H_d^- - \tilde R_u^+} = 297~\text{GeV} \, , \, .
\end{split}
\end{equation}
Considering only the particles with mass less than $200~\text{GeV}$, we recover the benchmark point in Eq.~\eqref{eq:bench} with the identifications as listed in Eqs.~\eqref{eq:defn-C} and \eqref{eq:defn-NS}.
\section{Conclusion}
\label{sec:conclusion}
The narrow width of a light Higgs boson ( $m_h \lesssim 130~\text{GeV}$ ), makes its decays particularly sensitive to new physics. Beyond-the-SM particles which interact with the Higgs, can open up important and even dominant Higgs decay modes while still coupling quite weakly. This susceptibility to new decay modes is particularly prevalent in weak-scale supersymmetry, where the underlying supersymmetric structure forces it to be light.
The particular scenario we explore here is an extension of the minimal supersymmetric setup by a new and massive gauge boson $Z'$. This $Z'$ has infinitesimal couplings to all matter, with the exception of Higgses and Higgsinos. Higgses in this scenario can decay to two $Z'$ gauge bosons, each of which subsequently decays to two leptons. Thus, with this small modification, we resurrect the gold-plated $h \rightarrow 4\,\ell$ decay mode for light Higgses. For maximum effect, $Z'$ should neither be extremely light nor extremely heavy, a case which has been under-represented in $Z'$ phenomenology.
A $Z'$ with exactly the right properties emerges automatically from the recent, viable gravity-mediated scenario of Ref.~\cite{Kribs:2010md}. Inspired by this model, we study Higgsphilic $Z'$ coming from the decays of supersymmetric particles, as well as $Z'$ coming from Higgs decay. For simplicity, and in the spirit of Ref.~\cite{Alwall:2008ag}, we perform these studies using a bare-bones version of the scenario in Ref~\cite{Kribs:2010md}, which contains only the phenomenologically important part of the Lagrangian.
For prompt $Z'$ decays we impose bounds from recent CDF $4\,\ell$ searches~\cite{CDF:10238}. Percent level $h \rightarrow Z'Z'$ fractions are completely consistent with all bounds and can lead to impressive and unmistakable signals, including $5~\sigma$ Higgs and $Z'$ discovery within the first $~\text{fb}^{-1}$ of LHC. We also discuss the possibilities for long-lived $Z'$ detection with $1~\text{fb}^{-1}$ of data.
Within this setup, we also find $4\ell + \slashchar{E}_T$ to be a superb channel for supersymmetry discovery. As a conservative estimate, we study electroweak pair-production of neutral Higgsinos that emit $Z'$ as they decay to the LSPs, leading to final states containing $Z'Z' + \slashchar{E}_T$. As the Higgsinos need to be pair produced, Higgsino discovery in $4\ell + \slashchar{E}_T$ requires more luminosity than the Higgs discovery: $\gtrsim 5~\text{fb}^{-1}$ of integrated luminosity. However, given that the majority of all supersymmetric events will eventually cascade-decay into neutralinos, the total rate of supersymmetric events yielding $4\ell + \slashchar{E}_T$ is potentially much higher than the rate we considered here.
\section*{Acknowledgments}
We thank Graham Kribs for comments on the manuscript. TSR acknowledges private communications to Henry Lubatti. AM is supported by Fermilab operated by Fermi Research Alliance, LLC under contract number DE-AC02-07CH11359 with the US Department of Energy. TSR is supported in part by the US Department of Energy under contract numbers DE-FGO2-96ER40956.
|
\section{Introduction}
It is commonly known that, according to the First Law of Dynamics, in the absence of an external force, every object remains in rest or moves along a straight line with uniform velocity. The aforementioned law is valid under the assumption that the mass of the object in question is constant. In many textbooks (see e.g.~\cite{1}) and dedicated papers (see e.g.~\cite{2}) the problem of variable mass bodies or systems was thoroughly analyzed. However, in all the examples it is assumed that an influx or efflux of mass is a uniformly changing function of time. In order to fill the gap we focus here on the dynamics of an oscillating mass object. It turns out that the study of behavior of such a body brings forward interesting physical results and consequences.
\section{Equations of Motion of a Variable Mass Object}
Let us consider an object with an initial mass of $m_{0}$ and its mass is some function of time $m=m(t)$. Additionally, we assume that there is no external force acting on the body ($\mathbf{F}=0$). The equation of motion in this case can be written down in the form
\begin{equation}\label{p}
\mathbf{F}=\frac{d\mathbf{p}(t)}{dt}=0,
\end{equation}
from which it clearly follows that $\mathbf{p}(t)=\mathbf{p}_{0}=const$.
Now, since mass varies with time, we find that
\begin{equation}\label{mfv}
\mathbf{p}_{0}=m(t)\mathbf{v}(t).
\end{equation}
Substituting Eq.~(\ref{mfv}) into (\ref{p}) we get that
\begin{equation}\label{Fgen0}
\frac{dm(t)}{dt}\mathbf{v}(t)+m(t)\frac{d\mathbf{v}(t)}{dt}=0.
\end{equation}
The product in the second term of LHS of Eq.~(\ref{Fgen0}) may be equal to zero only if the first term vanishes. Under the initial assumption that the mass changes in time, it happens only if $\mathbf{v}(t)=0$.
Rearranging the terms in Eq.~(\ref{Fgen0}) we immediately arrive at the equation
\begin{equation}\label{a}
\frac{d\mathbf{v}(t)}{dt}\equiv\mathbf{a}(t)=-\frac{dm(t)}{dt}\frac{\mathbf{v}(t)}{m(t)}=-\dot{m}(t)\frac{\mathbf{v}(t)}{m(t)}
\end{equation}
or if we utilize Eq.~(\ref{mfv}) we get
\begin{equation}\label{af}
\mathbf{a}(t)=-\frac{\dot{m}(t)}{(m(t))^{2}}\mathbf{p}_{0}.
\end{equation}
As we can see, in case $m=const$ we observe no acceleration and therefore we obtain the classical equation of motion of a body upon which no external force is acting ($\mathbf{a}=0$). However, in all other cases the acceleration turns up that is proportional to the mass change ratio. It also should be stressed here that, in general, the acceleration $\mathbf{a}(t)$ is not constant in time.
\section{Equations of Motion of an Oscillating Mass Object}
In this section we deal with a massive object whose mass oscillates in time. The choice of such a mass function is due to a sound assumption that energy should be conserved in time, i.e. an influx and efflux of energy are then compensated. It also means that the mass of the body remains constant if averaged over an integer number of oscillations.
Here we assume the mass function to be of the form
\begin{equation}\label{mf}
m(t)=m_{0}+\Delta m\sin\omega t,
\end{equation}
where $\Delta m$ is an amplitude of oscillations and $m_{0}\gg\Delta m$, and hence
\begin{equation}\label{mfder}
\dot{m}(t)=\omega\Delta m\cos\omega t.
\end{equation}
Substituting Eqs.~(\ref{mf}) and (\ref{mfder}) into (\ref{af}) we get that
\begin{equation}\label{acc}
\mathbf{a}(t)=-\frac{\omega\Delta m\cos\omega t}{(m_{0}+\Delta m\sin\omega t)^{2}}\mathbf{p}_{0}.
\end{equation}\newline
\begin{figure}[h]
\centering
\includegraphics[width=0.5\textwidth]{a1.eps}
\caption{Acceleration 10$\bm{a}(t)$ given by Eq.~(\ref{acc}) (solid) vs.~$-\cos(\omega t)$ function (dashed). $\Delta m=0.1m_{0}$, $m_{0}=1$, $|\bm{p}_{0}|=1$ and $\omega=1$ were assumed here.}
\label{fig1}
\end{figure}
As we can see in Fig.~\ref{fig1}, the acceleration function given by Eq.~(\ref{acc}) is significantly similar to a cosine function, and therefore resembles the oscillatory motion in which acceleration of a body is described by such a function.\newline
By integrating Eq.~(\ref{mfv}) in which the mass function $m(t)$ is given by Eq.~(\ref{mf}) we get
the solution of the equation of motion in the form \cite{2}
\begin{eqnarray}\label{xft}
\mathbf{x}(t) & = & \mathbf{p}_{0}\int\frac{dt}{m_{0}+\Delta m\sin\omega t} \nonumber \\
& = & \mathbf{p}_{0}\frac{2}{\omega\sqrt{m_{0}^{2}-(\Delta m)^{2}}}\arctan\frac{m_{0}\tan\frac{1}{2}\omega t+\Delta m}{\sqrt{m_{0}^{2}-(\Delta m)^{2}}}.\nonumber \\
\end{eqnarray}
The function described by Eq.~(\ref{xft}) is plotted in Fig.~\ref{fig2} for the same initial conditions and values as in Fig.~\ref{fig1}.
It is clearly noticeable that this type of motion is periodic, however it bears no resemblance to the regular oscillatory motion. It can also be easily checked that for $\Delta m=0$ Eq.~(\ref{xft}) turns into well-known formula $\mathbf{x}(t)=\mathbf{v}t$.
\section{Conclusions}
As we have shown, in the absence of an external force applied to a body it is possible that the body will accelerate provided its mass changes with time. Except for the rest reference frame, an observer notices non-zero acceleration of the body. The example directly indicates that the acceleration of an oscillating mass body is similar to the case of a classical oscillator, however the solution of the equation of motion reveals a significant difference from the oscillatory motion. These results may be important for many-body systems in which mass is exchanged between particles or bodies. In view of the aforementioned facts it may also be a contribution to a comprehension of quantum null oscillations on a classical ground by applying the Heisenberg uncertainty principle $\Delta E\Delta t\geqslant\frac{\hbar}{2}$ which states that the energy (and mass) of the system is not \emph{strictly} conserved and therefore serve as an attempt to establish a common ground for classical and quantum mechanics.
\begin{figure}[ht]
\centering
\includegraphics[width=0.5\textwidth]{a2.eps}
\caption{Solution of the equation of motion given by Eq.~(\ref{xft}). $\Delta m=0.1m_{0}$, $m_{0}=1$, $|\bm{p}_{0}|=1$ and $\omega=1$ were assumed here.}
\label{fig2}
\end{figure}
|
\section{Introduction}
Tidal streams are a widespread phenomenon in astrophysics, emerging from star clusters \citep{Grillmair1995}, dark matter subhalos \citep{Diemand2008} or satellites of galaxies \citep{Ibata1994}. It is present on all scales from galaxy clusters \citep{Calcaneo-Roldan2000} down to the the very smallest dark matter substructures \citep{Schneider2010}. The gravitationally unbound material forms spectacular long streams that trace the past and future orbit of the host system.
A wide class of tidal streams can be treated as collisionless systems, since they are dominated by stars or dark matter particles - the local relaxation time within the stream is much longer than the age of the Universe. Some streams contain gaseous material and are much more complicated to understand. For example, the oldest example of a 'stream' is the spectacular Magellanic HI stream, trailing well over 100 degrees behind the Magellanic Clouds. Initially modelled as a tidal mass loss feature from the Large Magellanic Clouds (\citet{Lin1977}), an alternative explanation is that it resulted from a more complex gravitational interaction between the Large and the Small Magellanic Cloud prior to their infall in the Milky Way potential \citep{Besla2010}. However, it may also be the case that this feature is purely hydrodynamical in origin since it contains no stars \citep{Moore1994}.
Galaxy mergers often create spectacular tidal tails that are somewhat different from the streams we consider in this paper. These streams are rapidly and violently created and they can contain dwarf galaxies aligned along the tails. \citet{Barnes1992} carried out simulations of galaxy mergers and found collapsed objects populating the stellar tails. Therefore they proposed collisionless collapse as the creation mechanism of tidal dwarf galaxies. However, \citet{Wetzstein2007} identified these collapsed objects as numerical artefacts due to insufficient resolution. They rather found that it's the gaseous part of the streams that triggers the collapse which leads to tidal dwarf galaxies.
The dynamics of tidal streams from star clusters and dwarf galaxies in our own halo have been extensively studied to constrain the mass and shape of the Galactic potential \citep{Johnston1999,Law2009}, alternative gravity models \citep{Read2005} as well as the orbital history of the satellites \citep{Kallivayalil2006,Lux2010}. However, the detailed evolution of the internal phase space structure of streams has received less attention \citep{Helmi1999,Eyre2010}.
Simulations, as well as observational data, show variations in the width and the internal structure of tidal streams. Likewise the density along a stream can vary considerably, the most prominent example being the symmetric streams originating from the globular cluster Palomar 5 with its equally spaced density clumps \citep{Odenkirchen2001, Odenkirchen2002}. There are different explanations for these clumps such as disc shocking \citep{Dehnen2004}, effects due to the dark matter substructures \citep{Mayer2002,Yoon2010} or epicyclic motions in the stellar orbits \citep{Kuepper2008,Just2009}.
Another interpretation was given recently by Quillen and Comparetta \citep{Quillen2010,Comparetta2010}, who argued that clumps in streams are the result of longitudinal Jeans instabilities. In their model they describe a tidal stream as an extended static cylinder of stars and they use the results of \citet{Fridman1984}, that infinitely extended cylinders are gravitationally unstable. With an estimated relation for the velocity dispersion and the linear density in the stream, Quillen \& Comparetta find a longitudinal Jeans length of several times the stream width. Comparing their results to the observations of Palomar 5, they find agreement between the distance between clumps in the streams and their fastest growing mode of the gravitational instability.
However, their model of a static cylinder does not take into account the expansion that happens due to the diffluence of the stars in the stream. Once in the stream the stars are no longer bound to the cluster, their intrinsic dispersion causes the stream to grow along the orbital direction. Escaping stars also have an intrinsic dispersion, related to the dispersion in the outer cluster region.
Another way of understanding the expansion is by considering the velocity difference between the substructure and the outflowing stars, which depends on the tidal radius. Since the tidal radius is shrinking with time, stars that leave the cluster at later times are slower than stars that left before and this leads to the expansion of the stream. In reality the situation is even more complicated. The stream length is actually oscillating during one orbit, being stretched at pericenter and compressed at apocenter. The linear expansion only acts on average over several orbital periods. Therefore for short timescales, the periodic oscillating effect must be taken into account.
This paper is structured as follows: In section 2 we construct a simplified model for a tidal stream and we find a relation between the stream density and its velocity dispersion. Sections 3 and 4 are dedicated to the study of the stream stability, where we first derive the linearised equation of perturbations and then look at the one dimensional collapse along the stream direction. In section 5 we take a critical look at our model by comparing with the detailed dynamics of streams using N-body simulations. The orbital oscillation of the stream length and its influence on collapse are considered. Finally we give our conclusions in section 6.
\section{Modelling a tidal stream}
The general case of a streaming cluster is a problem of many particle dynamics that can be solved self-consistently with simulations. Analytical statements can be made by considering a model with simplifying assumptions. Thereby one has to be careful to avoid over-simplification. We model a tidal stream as a self gravitating cylinder of collisionless matter with an expansion in the direction of the cylinder axis. For the cluster as well as for the host we choose isothermal spheres so that we can use the simplifying relations
\begin{equation}\label{isorelations}
\frac{r_t}{R}\sim c\left(\frac{m}{M}\right)^{1/3}\sim c^{3/2}\left(\frac{\sigma_{cl}}{\sigma_{gal}}\right),\hspace{0.5cm}GM=2\sigma_{gal}^2 R,
\end{equation}
where $m$, $M$ and $\sigma_{cl}$, $\sigma_{gal}$ are the masses respectively the velocity dispersions of the cluster and the host. The distances $r_t$ and $R$ are the tidal radius of the cluster and the orbital radius to the host. For an isothermal sphere the correction factor $c\sim0.8$ \citep{Binney2008}.
Whilst many systems can be reasonably well described by an isothermal potential over the scales of interest, our results would not apply to systems orbiting within very different potentials. For example, a star cluster within a constant density potential would not even produce streams. However, for Palomar 5 and for many of the streams in our Galactic halo, or within galaxy clusters, an isothermal potential is a good approximation over the range $0.01-0.5R_{virial}$ \citep{Klypin2002}.
In order to test the basic assumptions of our model we perform simulations of a star cluster orbiting with different eccentricities within an isothermal host potential. The velocity dispersion is chosen to be $\sigma_{cl}=4$ km/s for the cluster and $\sigma_{gal}=200$ km/s for the host. Every simulation starts with the star cluster at a radius of 20 kpc and we choose different perpendicular initial velocities from 283 km/s for a circular orbit to 50 km/s for the most eccentric orbit. The global potential is a fixed analytic potential whilst the star cluster is modelled using $2\times10^5$ stars, set up in an equilibrium configuration at the starting position \citep{Zemp2008}. The evolution is followed using the N-body code PKDGRAV \citep{Stadel2001}, adopting high precision parameters for the force accuracy. The softening length of the star particles is $\epsilon=0.005$ kpc. All simulated orbits are illustrated in Fig. \ref{orbits}.
\begin{figure}
\centering
\includegraphics[scale=0.85]{plots/orbits.pdf}
\caption{Orbits of the star clusters in our simulations. The eccentricity is given in terms of the parameter $b$ defined as $\dot{R}=b V$ (with $b=0$ for a circular orbit and $b=1$ for a radial infall). In increasing eccentricity: continuous ($b=0$), narrow-dotted ($b=0.14$) broad-dotted ($b=0.34$), dashed-dotted ($b=0.54$) dashed ($b=0.74$) and continuous ($b=0.88$).}
\label{orbits}
\end{figure}
There are two mechanism responsible for the stream growth, on the one hand, the outflow of matter leaving the cluster with a certain velocity difference $\Delta V$ and on the other hand the stream expansion due to diffluence of the initial dispersion. The expansion velocity $w$ is given by $w\sim 2\sigma_{cl}$, which corresponds to the diffluence velocity of a bunch of particles leaving the cluster at the same time. Using conservation of angular momentum $L$ leads to the velocity difference $\Delta V$:
\begin{equation}\label{deltav}
L=RV\sin\theta=(R+r_t)(V-\Delta V)\sin\theta\hspace{0.2cm}\Rightarrow\hspace{0.2cm}\frac{\Delta V}{V}\sim \frac{r_t}{R}
\end{equation}
Here we have assumed $R\gg r_t$ and $V\gg\Delta V$. In an isothermal potential the value of $V$ must be somewhere between $V_R=(4/\pi)^{1/2}\sigma_{gal}$ and $V_c=2^{1/2}\sigma_{gal}$, the radial and circular velocities. Using (\ref{isorelations}) we therefore obtain $\Delta V\sim \sigma_{cl}$ as well as
\begin{equation}\label{velocityrelation}
w\sim 2\Delta V.
\end{equation}
Physically this means that all particles belonging to the stream at $t_0$ will be distributed over the entire stream length at all time $t>t_0$. Or in other words, even if there is no more outflow from the cluster, the stream always stays attached to the cluster.
The amount of diffluence can be estimated in the simulation by marking particles at a certain time $t_0$ and looking at where they are found in the stream at $t\gg t_0$. The first image of Fig. \ref{marked_particles} shows a cluster at apocenter after one orbit (195 Myr) with the particles of one stream marked in red. In the second image we see the cluster at apocenter after nine orbits (1756 Myr) along with the distribution of the particles marked before. The particles marked at the early time are located throughout the stream at later times, confirming our above statement.
\begin{figure}
\centering
\begin{minipage}{8cm}
\includegraphics[scale=0.58]{plots/plot00195.pdf}
\end{minipage}
\begin{minipage}{8cm}
\includegraphics[scale=0.63]{plots/plot01756.pdf}
\end{minipage}
\caption{\textit{Simulation of an isothermal cluster with an eccentricity of $b = 0.74$. The image on the top shows the cluster after one orbit where the particles of one stream are marked in red. The image on the bottom shows the cluster after nine orbits with the distribution of the particles marked before.}}
\label{marked_particles}
\end{figure}
The width of the stream depends on the velocity dispersion $\sigma$. A particle with an energy excess during the outflow will be on an orbit with a slightly different eccentricity and will therefore complete a full oscillation within the stream during one orbital time $T$. The radius corresponding to half of the stream width is then approximately given by
\begin{equation}\label{streamwidth}
r_{\bot}\sim\frac{1}{2}\sigma T.
\end{equation}
On the other hand the length of the stream after one orbit is simply
\begin{equation}\label{length}
l_0\sim wT\sim 2\sigma T,
\end{equation}
assuming the approximate relation $\sigma _{cl}\sim\sigma$. After one orbit a single stream should therefore be about twice as long as it is wide. This is the case in all our simulations and can be checked in the first image of Fig \ref{marked_particles}.
The linear density of a stream is given by the relation
\begin{equation}\label{mu}
\mu=\frac{dm}{dz}=\frac{\dot{m}}{\dot{z}}\sim\frac{\dot{m}}{2\Delta V},
\end{equation}
Here we have used $\dot{z}\sim w\sim 2\Delta V$, what results in an additional factor of two compared to a static stream because of the stretching effect of the expansion. The rate of outstreaming matter is estimated to be
\begin{equation}\label{dotm}
\dot{m}=\frac{(m_a-m_p)}{T}\sim\frac{2c^{\frac{3}{2}}}{T}\frac{\sigma_{cl}^3}{G\sigma_{gal}}(R_a-R_p)\sim\frac{c^{\frac{3}{2}}}{G}\frac{\sigma_{cl}^3}{\sigma_{gal}}\dot{R},
\end{equation}
where we have used the relations (\ref{isorelations}). The outflow of the matter is averaged over one orbital period. With the relation $\dot{R}=bV$, where the parameter $b$ depends on the cluster orbit (with $b=0$ for a circular orbit and $b=1$ for a radial infall), the linear density becomes
\begin{equation}
\mu\sim\frac{c^{3/2}}{2G}\frac{\sigma_{cl}^3}{\sigma_{gal}}\left(\frac{V}{\Delta V}\right)b\sim\frac{\sigma_{cl}^2b}{2G},
\end{equation}
and the Toomre parameter is then given by
\begin{equation}\label{Toomreparameter}
q\equiv\frac{\sigma^2}{2G\mu}\sim\frac{1}{b}.
\end{equation}
The smallest value for the Toomre parameter is therefore $q\sim1$ which corresponds to a radial orbit. The Toomre parameter of relation (\ref{Toomreparameter}) is four times larger than the one obtained by \citet{Quillen2010}, the reason being a factor of two which comes in at equation (\ref{mu}) as well as the averaging of the mass outflow in equation (\ref{dotm}). Both effects are directly related to the expansion of the stream, not considered by Quillen and Comparetta.
An independent way to calculate the Toomre parameter is by using the virial theorem for an isothermal sphere, truncated at the tidal radius $r_t$:
\begin{equation}
\sigma_{cl}^2=\frac{|W|}{m_{cl}}=\frac{4\pi G}{m_{cl}}\int_{0}^{r_t}dr r \rho(r)M(r)=\frac{4\sigma_{cl}^4 r_t}{Gm_{cl}},
\end{equation}
\begin{equation}
\sigma_{cl}^2=\frac{Gm_{cl}}{4 r_t}.
\end{equation}
Using the approximation $\sigma_{cl}\sim\sigma$ then leads to
\begin{equation}
q=\frac{\sigma^2l_0}{2Gm_{st}}\sim\frac{1}{8}\frac{m_{cl}}{m_{st}}\frac{l_0}{r_{t}}.
\end{equation}
For the extreme case of a radial orbit $m_{cl}\sim 2 m_{st}$ and we obtain $q\sim 1$. This means that for all orbits $q$ must be larger than one, a result that confirms the relation (\ref{Toomreparameter}) above.
The Toomre parameter can also be determined in the simulations by measuring the velocity dispersion and the linear density. However, it turns out that the dispersion is very difficult to quantify accurately because over one orbital period it strongly fluctuates at any Lagrangian point (for example, around any star). This is due to the oscillation of stream-length and stream width, which happens because the particles in the stream are on nearly free orbits around the host. The velocity dispersion therefore is affected by the number of stars used in its measurement since that changes the region of the stream over which the dispersion is calculated.
In Fig. \ref{q(b)} we plot the Toomre parameter, where the dispersion is measured in the middle of the stream, at apocenter after one orbital period and assuming an isotropic distribution (looking at the ration of tangential to radial velocity dispersions after one orbit we can see that this assumption is approximately valid). We notice that the simulations roughly follow the theoretical prediction which is given by the solid gray line.
\begin{figure}
\centering
\includegraphics[scale=0.8]{plots/qb.pdf}
\caption{The Toomre parameter as a function of the eccentricity parameter $b$. The black dots are the measurements from the different simulations. The solid gray line corresponds to equation (\ref{Toomreparameter}), while the dotted line is the prediction from \citet{Quillen2010}.}
\label{q(b)}
\end{figure}
Already at that stage of our analysis it becomes clear that the expansion has a strong stabilising effect because it leads to a significant boost of the Toomre parameter. In the next section we will see that the stability of a stream is additionally enforced, since the expanding environment leads to a damping in the the evolution of perturbations.
\section{Perturbations in an expanding cylinder}
In order to find a criterion for the stability, we are now modeling a tidal stream as a non-rotating elongated cylinder of collisionless matter that is linearly expanding in the direction of its long axis. For the expansion we introduce the comoving coordinate $s=az$ with $a(t)=\alpha t$ and set $z=l_0$, where $l_0$ is the stream length after one orbital period $T$. The expansion factor then becomes
\begin{equation}\label{alpha}
\alpha=\frac{1}{T}.
\end{equation}
The orbital period is a natural time measure since the outstreaming from the cluster into the tails is mainly happening during the cluster orbit from apo- to pericenter when the tidal radius is shrinking. During the other half of the orbit the tidal radius is growing again and there is nearly no streaming mass loss.
An analytical treatment of the stability of an expanding cylinder is possible either on scales much smaller or much larger than the cylindrical radius. In the former case we can treat the fluid as homogeneous and we therefore get the usual Jeans length
\begin{equation}\label{homJeanslength}
\lambda^h_J=\sqrt{\frac{\pi \sigma^2}{G\rho}}.
\end{equation}
With the relation (\ref{Toomreparameter}) as well as the linear density $\mu=\pi r_{\bot}^2\rho$ we then obtain
\begin{equation}
\frac{\lambda^h_J}{r_{\bot}}=\sqrt{2\pi^2 q}\sim\sqrt{\frac{2\pi^2}{b}}.
\end{equation}
Since the eccentricity parameter $b$ is always larger than one, the Jeans length exceeds the radius of the cylinder and we can exclude collapse on scales smaller than $r_{\bot}$.
However there is still the possibility of collapse in the longitudinal direction of the cylinder on scales larger than $r_{\bot}$. This is the second analytically treatable case which leads to a very different stability criterion. In order to determine the behaviour of longitudinal perturbations we are now going to derive the equations for the evolution of density perturbations. This is usually done by integrating and linearising the collisionless Boltzmann equation \citep{Peebles1980}. Since we are looking at a thin cylinder, we can assume a phase-space density of the form
\begin{equation}\label{phasespacedensity}
f(z,p,t) = \left\lbrace \begin{array}{cc} a[\rho_b+\rho_1(z,t)]f(p), & r<r_{\bot} \\ 0, & r>r_{\bot} \end{array} \right.
\end{equation}
Here we have introduced a homogeneous background density $\rho_b$ as well as a first order perturbation $\rho_1$. An integration of the phase-space density immediately leads to the stream density
\begin{equation}
\rho= \frac{1}{a}\int dpf(z,p,t)=\frac{(\mu_b+\mu_1)}{\pi r_{\bot}^2}=\frac{1}{\pi r_{\bot}^2}\frac{\mu_0}{a}(1+D),
\end{equation}
where $D=\mu_1/\mu_b$ is the dimensionless overdensity.
The evolution of the phase-space density is described by the one dimensional collisionless Boltzmann equation with expanding coordinate
\begin{equation}\label{Boltzmann}
\frac{\partial}{\partial t}f(z,p,t) + \frac{p}{a^2}\frac{\partial}{\partial z} f(z,p,t) - \frac{\partial\Phi}{\partial z}\frac{\partial}{\partial p}f(z,p,t)=0.
\end{equation}
It is now straightforward to derive the continuity and the momentum equation of the stars in the cylinder. They are given by
\begin{equation}
\partial_t(1+D) + \frac{1}{a}\partial_z\left[\langle v\rangle(1+D)\right] = 0,
\end{equation}
\begin{equation}
\partial_t\left[a\left<v\right>(1+D)\right] + \partial_{z}\Phi(1+D) + \partial_z\left[\langle v^2\rangle(1+D)\right] = 0,
\end{equation}
where
\begin{equation}\label{avvelavdisp}
\langle v\rangle=\frac{\int p f dp}{a\int fdp},\hspace{1cm}\langle v^2\rangle=\frac{\int p^2fdp}{a^2\int fdp}.
\end{equation}
By substituting the derivative of the second equation into the first we finally find the equation of perturbation:
\begin{equation}\label{deltaequation}
\partial_t^2 D + 2\frac{\dot{a}}{a}\partial_t D = \frac{1}{a^2}\partial_z\left[(1+D)\partial_z\Phi\right] + \frac{1}{a^2}\partial_z^2\left[(1+D)\langle v^2\rangle\right].
\end{equation}
In order to solve this differential equation we still need to know the potential of a cylinder. The simplest assumption is to take
\begin{equation}
\Phi(r,z,t)=\Phi^{(0)}(r)+\Phi^{(1)}(r,z,t),
\end{equation}
\[
\Phi^{(1)}(r,z,t)=\phi^{(1)}(r,t)e^{ik_0z}
\]
\citep{Fridman1984}, where $k_0$ is the comoving wave number in z-direction. The Poisson equation for the zero-order term is simply
\begin{equation}
\frac{1}{r}\frac{d}{dr}\left(r\frac{d\Phi^{(0)}}{dr}\right)=4\pi G \left\lbrace\begin{array}{cc}\rho_0, & r<r_{\bot} \\ 0, & r>r_{\bot}\end{array}\right.
\end{equation}
with the solution $\Phi^{(0)}(r)=\pi G \rho_0 r^2 + const$. In contrast to a self gravitating cylinder, a stream is embedded in the dominating potential of the host and the zero order term looks different. However, a dependence of the potential in the z-direction only comes in as a first order effect due to the internal structure of the stream. Therefore we obtain the following Poisson equation at first order
\begin{equation}
\partial_r^2\Phi^{(1)} + \frac{1}{r}\partial_r\Phi^{(1)} - \frac{k_0^2}{a^2}\Phi^{(1)} = \left\lbrace\begin{array}{cc}4\pi G \rho_1, & r<r_{\bot} \\ 0, & r>r_{\bot}\end{array}\right.
\end{equation}
where $\rho_1$ may vary along the axis of the cylinder. Two independent solutions of this homogeneous differential equation are the modified Bessel equations of first and second kind $I_0[x]$ and $K_0[x]$. The general inner and outer solution are given by
\begin{equation}
\Phi_{<}^{(1)}(r) = AI_0\left[\frac{k_0 r}{a}\right]+ BK_0\left[\frac{k_0 r}{a}\right] - \frac{4\pi G \rho_1 a^2}{k_0^2},
\end{equation}
\begin{equation}
\Phi_>^{(1)}(r) = A'I_0\left[\frac{k_0 r}{a}\right]+ B'K_0\left[\frac{k_0 r}{a}\right],
\end{equation}
where the boundary conditions require $B=A'=0$. With the matching conditions $\Phi_<^{(1)}(r_{\bot})=\Phi_>^{(1)}(r_{\bot})$ and $\partial_r\Phi_<^{(1)}(r_{\bot})=\partial_r\Phi_>^{(1)}(r_{\bot})$ we find
\begin{equation}
A=\frac{4\pi G \rho_1 a^2 K_0'\left[k_0r_{\bot}/a\right]}{k_0^2W\left[k_0 r_{\bot}/a\right]},\hspace{0.2cm}W = -\frac{k_0}{a}(I_0K_1 + I_1K_0).
\end{equation}
We now look at the case of large perturbations in a thin stream ($k_0r_{\bot}/a<<1$). In the asymptotic limit we get
\begin{equation}
A \simeq \frac{2G\mu_0}{a}\left[\frac{2a^2}{(k_0r_{\bot})^2} + \gamma + \log\left(\frac{k_0r_{\bot}}{2a}\right)\right]D.
\end{equation}
The first order potential inside the stream is then given by
\begin{equation}\label{potential}
\Phi_{<}^{(1)} \simeq \frac{\sigma_0^2}{q_0 a}\left[\log\left(\frac{k_0r_{\bot}}{2a}\right)+\gamma\right]D,
\end{equation}
where we have used $\mu=\pi r_{\bot}^2\rho$ together with relation (\ref{Toomreparameter}). The Euler constant is $\gamma=0.577$.
Using (\ref{deltaequation}) and (\ref{potential}) we obtain a closed set of equations for the perturbations $D$ that can now be linearised. We therefore set $D<<1$, as well as $\langle v^2\rangle(z,t) = \sigma^2(t) + O(v_1^2)$ which gives
\begin{equation}
\ddot{D} + 2\frac{\dot{a}}{a}\dot{D} = \frac{1}{a^2}\partial_z^2\Phi_<^{(1)} -\frac{k_0^2\sigma^2}{a^2}D,
\end{equation}
\begin{equation}\label{lindeltaequation0}
\ddot{D} + 2\frac{\dot{a}}{a}\dot{D} = -\frac{k_0^2}{a^2}\left\lbrace \frac{2G\mu_0}{a}\left[\log\left(\frac{k_0r_{\bot}}{2a}\right)+\gamma\right] + \sigma^2\right\rbrace D.
\end{equation}
In a collisionless cylinder the longitudinal velocity dispersion decreases as $\sigma=\sigma_0a^{-1}$ (see eq. \ref{avvelavdisp}), while the perpendicular velocity dispersion stays constant. Equation (\ref{lindeltaequation0}) can therefore be written as
\begin{equation}\label{lindeltaequation}
\ddot{D} + 2\frac{\dot{a}}{a}\dot{D} = -\frac{\sigma_0^2 k_0^2}{q_0 a^3}\left\lbrace\log\left(\frac{k_0r_{\bot}}{2a}\right) + \gamma + \frac{q_0}{a}\right\rbrace D.
\end{equation}
The perturbation, $D$, is damped if the right hand side of equation (\ref{lindeltaequation}) is negative. Therefore we can define a Jeans length
\begin{equation}\label{Jeanslength}
\lambda_J=\pi r_{\bot} \exp\left(\frac{q_0}{a}+\gamma\right),
\end{equation}
which is very different from the stability criterion in a homogeneous surrounding (\ref{homJeanslength}). The geometry of a thin cylinder leads to a Jeans length with an exponential form that guarantees stability up to much larger scales.
Equation (\ref{lindeltaequation}) can now be simplified using (\ref{streamwidth}) and taking $a$ as variable:
\begin{equation}\label{lindeltaequation2}
D''(a) + \frac{2}{a}D'(a) =
\end{equation}
\[
-\frac{4 (k_0r_{\bot})^2}{q_0 a^3}\left\lbrace\log\left(\frac{k_0r_{\bot}}{2a}\right) + \gamma + \frac{q_0}{a}\right\rbrace D(a),
\]
There are two remaining free parameters, namely $q_0$ and $k_0r_{\bot}$, which describe the eccentricity of the orbit and the scale of the perturbation compared to the width of the stream. In Fig. \ref{perturbations} we plotted the numerical solutions for different sets of parameters. For a small Toomre parameter the perturbations will become nonlinear and we expect gravitational collapse to occur. However, for larger $q$ the perturbations either undergo a damped oscillation or they freeze out after an unsubstantial phase of growth. Comparing these results with the relation (\ref{Toomreparameter}) leads to the conclusion that the Toomre parameter of a tidal stream is always large enough to assure stability in all cases of interest.
\begin{figure}
\centering
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/D05.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/D075.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/D1.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/D2.pdf}
\end{minipage}
\caption{\textit{Dynamics of the perturbation $D$ with respect to the scale factor $a$ for the comoving factors $k_0r_{\bot}=$ 0.9 (solid), 0.7 (wide-dashed), 0.5 (narrow-dashed) 0.3 (dotted) and 0.1 (dashed-dotted). From top left to bottom right: $q=$ 0.5, 0.75, 1, 2}}
\label{perturbations}
\end{figure}
\begin{figure}
\centering
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/sD05.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/sD1.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/sD075.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/sD2.pdf}
\end{minipage}
\caption{\textit{Perturbations $D$ in the case of a static cylinder, where $o$ is the number of orbits. The different lines represent the factors $kr_{\bot}=$ 0.9 (solid), 0.7 (wide-dashed), 0.5 (narrow-dashed) 0.3 (dotted) and 0.1 (dashed-dotted) in static coordinates. From top left to bottom right: $q=$ 0.5, 0.75, 1, 2}}
\label{staticperturbations}
\end{figure}
In order to see the effect due to the linear expansion, we also look at the case of a static cylinder. The evolution of perturbations is then given by equation (\ref{lindeltaequation}) with $a=1$ and $\dot{a}=0$ and its behaviour is plotted in Fig. \ref{staticperturbations}.
Even for a large Toomre parameter $q$, there are always collapsing modes supposing an infinitely extended cylinder. This is fundamentally different in the expanding case, where all modes are damped for a high enough $q$, leading to stability on all scales.
Until now we analysed the stability of a stream with linear perturbation theory. In the next section we take a different look at the stream stability by exploring the longitudinal collapse of cylindrical slices. This somehow more heuristic approach is not restricted to the linear regime and gives an independent analysis of the problem.
\section{Shell collapse in a cylinder}
In a one dimensional case of an extended cylinder the spherical collapse reduces to the longitudinal collapse of thin slices. We therefore consider a homogeneous and infinitely long expanding cylinder with a top hat perturbation at the time $t_i$. The stream can then be cut into slices, which evolve at constant energy. The energy at a certain distance $s$ is given by
\begin{equation}
E_i= \frac{1}{2}v_i^2 + \Phi(s)=\frac{1}{2}\left(\frac{\dot{a}_i}{a_i}\right)^2s^2 -\frac{GM_i(s)}{s}.
\end{equation}
Since the mass $M_i$ evolves as
\begin{equation}
M_i(s) = \int (1 + \delta)\mu_b(t_i) ds = \frac{2\mu_0}{a_i}(1 + \delta)s
\end{equation}
we obtain the energy
\begin{equation}\label{energy}
E_i = \frac{1}{2}\alpha^2\left(\frac{s}{a_i}\right)^2 - 2G\frac{\mu_0}{a_i}(1+\delta).
\end{equation}
Slices with a positive total energy will never collapse and therefore $E_i\geq0$ is our stability condition. Equation (\ref{energy}) then leads to
\begin{equation}\label{stabcon0}
s \geq r_{\bot}\sqrt{\frac{8(1+\delta)a_i}{q_0}},
\end{equation}
where we have used the definition of the Toomre parameter (\ref{Toomreparameter}). Slices further away are stable while nearby ones will collapse. The critical distance below which the stream becomes unstable is growing with the square root of time.
In the picture of shell collapse the velocity dispersion is completely ignored, since the diffusion of particles into other slices makes the problem much more complicated. We will however account for the dispersion by an \textit{ad-hoc} introduction of the Jeans length $\lambda_J$, which guaranties the stream stability on small scales. With (\ref{stabcon0}) and (\ref{Jeanslength}) we can then construct the stability criterion
\begin{equation}
q_0 \geq \frac{8(1+\delta)}{\pi^2}a_i e^{-2(q_0a_i^{-1}+\gamma)},
\end{equation}
which is fulfilled at the beginning ($a=1$) and may be violated at some later times ($a>a_c$).
This means that for $t_i=t_0$ all instable slices are below $\lambda_J$ and therefore all the stream is stable. Later on however and depending on $q_0$ unstable modes may appear just above $\lambda_J$.
Since the Jeans length gives a minimum size for the final structure, the initial collapse must start at a scale well above this. A calculation of the collapse-time $t_{coll}$ shows however that $t_{coll}$ dramatically grows with the distance of the slice. Slices only a few times further away than the Jeans length already have a $t_{coll}$ that largely exceeds one Hubble time, at least for $q_0\geq1$. This means that even though there are unstable modes in an expanding stream, they will never have enough time to grow substantially. Collapse only occures for very small values of $q_0$ well below the limit given by (\ref{Toomreparameter}).
This qualitative picture is in agreement with the results plotted in Fig. \ref{perturbations}, where a phase of damped oscillation is followed by a phase of growth, freezing out at a very low level still in the linear regime.
\section{Towards a realistic stream}
A realistic treatment of a tidal stream orbiting its host galaxy can become very complex, which leads us to consider the possibility that our model of an expanding cylinder is an over-simplification and therefore we are missing some important dynamics. In the following we treat possible deviations to our model and discuss their influence on the stability:
\begin{list}{\labelitemi}{\leftmargin=0em}
\item In general, the host galaxy is not simply isothermal, but can have a triaxial shape that varies with time, and it contains substructures. The orbit of a cluster is then no longer within a plane and it may lose its regularity. The analysis of stability effects in such a complex situation is best tackled with full numerical simulations. Nevertheless, there is no a-prior reason to believe that one of these effects could fundamentally alter the stability criterion.
\item A stream approximately traces the orbit of its cluster and is therefore more and more curved the longer it gets. This does not correspond to the straight cylinder used in the model. However the effect of the bending is rather stabilising the stream against longitudinal Jeans instabilities and can therefore confidently be ignored.
\item A much more severe limitation of our model is the assumption of a cylindrical form. In reality streams are often more sheet-like and their thickness strongly varies during the orbital period. The closer a stream approaches the centre of the host, the thinner it gets. The reason for this behaviour is the form of the isothermal host potential which leads to orbits that occupy a narrower real-space volume closer to its centre.
In Fig. \ref{halo_v100} the image of a stream on an eccentric orbit is illustrated. The difference in the stream-width is very pronounced and the sheet like structure at apocenter is also visible. Even though the variation in the thickness has a major influence on the local stream density, it does not affect the longitudinal collapse condition, which only depends on the linear density. Incorporating the effect of the flattening of the stream is somewhat more difficult because it affects the potential (\ref{potential}). However, it is again unlikely that the sheet-like structure would have an enhancing effect on the collapse since it is stretching the stream which reduces its density.
\begin{figure}
\centering
\includegraphics[scale=0.45]{plots/hrhalo_v100_denmap.pdf}
\caption{Density map of a star cluster with a leading and tailing stream after 2 Gyr in an isothermal host potential. The orbit lies in the (y,z)-plane and has an eccentricity factor of $b=0.74$. The high eccentricity leads to strong variations in the stream width. While the streams are narrower and denseer at pericentre, they become flattend at apocentre with the typical umbrella-like form.}
\label{halo_v100}
\end{figure}
\item As the stream orbits between apocenter and pericenter, its length is oscillating, a fact that is not included in our model assumptions and may affect the stream stability. In fact, the stream only expands linearly on average, its length oscillates during one orbit, being stretched at pericenter and compressed at apocenter. In Fig. \ref{avdis} the average distance of random points in streams on different orbits are illustrated and the orbital oscillation as well as the overall linear expansion are clearly visible.
\begin{figure}
\centering
\includegraphics[scale=0.45]{plots/avdis.pdf}
\caption{The evolution of the distance between chosen particles in the stream for different simulations with $b=0.14$ (full), $b=0.34$ (dashed), $b=0.54$ (dashed-dotted), $b=0.74$ (narrow-dotted) and $b=0.88$ (broad-dotted). Whilst there is linear growth averaged over the orbital motion, the length is oscillating with the orbit, and the amplitude of the oscillation is larger for higher eccentricity. }
\label{avdis}
\end{figure}
These oscillations have an effect on the longitudinal perturbations. From peri- to apocenter, when the stream-length is shrinking, we are no longer in a stable regime and we expect growth. However, this growth happens on a timescale longer than the orbital period so that perturbations do not have time to collapse.
This can be shown by approximating the shrinking of the stream with a linearly decreasing scale factor of the form
\begin{equation}
a(t)=d-\left(d-1\right)\frac{2}{T}t,
\end{equation}
where $d=l_{max}/l_{0}$. The stream length $r(t)=a(t)l_{0}$ now runs from $l_{max}$ to $l_{0}$ in half of an orbital period. We then use the equation of perturbation (\ref{lindeltaequation}) and replace the time variable with the scale factor. The result is
\begin{equation}\label{lindeltaequation3}
D''(a) + \frac{2}{a}D'(a) =
\end{equation}
\[
-\frac{(k_0r_{\bot})^2}{(d-1)^2q_0 a^3}\left\lbrace\log\left(\frac{k_0r_{\bot}}{2a}\right) + \gamma + \frac{q_0}{a}\right\rbrace D(a),
\]
as well as the initial conditions $D(d)=0.1$ and $D'(d)=0$. We find growing solutions if the right hand side of the above equation is positive, where the actual value determines the growth rate. A large value of $d$ (high eccentricity) gives a small growth factor for a long interval of integration, whilst a small value (low eccentricity) gives a large growth factor for a short interval, the reason being the $d^2$-term in the denominator of (\ref{lindeltaequation3}). Hence, the actual growth of perturbations stays negligibly small in all cases even for a $q$ as low as 0.5 and the overall stability of our streams is therefore ensured. In Fig. \ref{growth} we plotted the evolution of the perturbations between peri- and apocenter for the case of $q=0.5$ and $q=1$ and with $d=2$ and $d=10$.
\begin{figure}
\centering
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/shrinkingq05c2.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/shrinkingq05c10.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/shrinkingq1c2.pdf}
\end{minipage}
\begin{minipage}{4cm}
\includegraphics[scale=0.4]{plots/shrinkingq1c10.pdf}
\end{minipage}
\caption{\textit{Growing perturbations $D$ for a shrinking scale factor $a$ with $k_0r_{\bot}=$ 0.9 (solid), 0.7 (wide-dashed), 0.5 (narrow-dashed), 0.3 (dotted), 0.1 (dashed-dotted). The plots should be read from right to left. Top: $q=0.5$ with $d=2$ (left) and $d=10$ (right). Bottom: $q=1$ with $d=2$ (left) and $d=10$ (right).}}
\label{growth}
\end{figure}
\item Because of the longitudinal contraction at apocenter and the transversal contraction at pericenter the density and the total velocity dispersion are oscillating twice as fast as the stream length. This can be observed in Fig. \ref{sigma}, where we plotted the longitudinal and transversal velocity dispersion of a stream with high orbital eccentricity. The doubling of the frequency comes from the fact that the stream is longitudinally compressed at apocenter and transversely compressed at pericenter.
\begin{figure}
\centering
\includegraphics[scale=0.8]{plots/sigma_v100.pdf}
\caption{\textit{Evolution of the velocity dispersion for a high eccentricity orbit ($b=0.75$). The dispersion parallel to the stream is plotted at the top, the one perpendicular to the stream at the bottom. The grey dahsed curves show the time evolution predicted by the model. The vertical lines correspond to the apocenter passage of the cluster.}}
\label{sigma}
\end{figure}
Fig. \ref{sigma} can be understood qualitatively by assuming that the particles in the stream are on nearly free epicyclic orbits around the host, which means that the host potential is dominating and that the stream particles are not feeling each other. Slightly displaced orbits are then crossing at apo- and again at pericenter which leads to large peaks in the velocity dispersion.
Our model predicts a longitudinal dispersion that decreases on average, an effect that is not clearly visible in the plot on the top of Fig. \ref{sigma}. Whilst the minima in the longitudinal dispersion seem to decrease as predicted, the maxima are growing with time. This growth comes from the fact that the particle orbits separate more and more to end up at distinct free orbits with the same eccentricity but with a shift in the azimuthal angle. The particles are then all crossing at the same place leading to a sharp peak in the dispersion. The orbital oscillation is also visible in the plot of the transversal dispersion at the bottom of Fig. \ref{sigma}. On average however the transversal dispersion seems to stay constant as predicted by the model. A more detailed study of the stream dispersion was done by \citet{Helmi1999}, who found a similar evolution of the dispersion over many more orbital periods.
\end{list}
\section{Conclusions}
We have studied the gravitational stability of tidal streams by modelling them as thin linearly expanding cylinders of collisionless matter. Such a model leads to a stability criterion that has an exponential dependence on the one dimensional Toomre parameter.
We derive a perturbation analysis and also use energetic arguments, to show that a cylinder with the dispersion, the density and the growth rate of a tidal stream is stable for all times.
We used numerical simulations to test our main approximations and to study the detailed phase space evolution of tidal streams. As a final consistency check, we note that none of our simulations show any evidence for gravitational instability.
In reality, a stream is only linearly expanding on average, its length is oscillating during one orbit. This leads to a time interval between apo- and pericenter, where the scale factor shrinks again and the stream is in an unstable regime. Nevertheless, this time interval is too short for the perturbations to grow substantially and the oscillation of the stream length has therefore no influence on the stability.
Collisionless stellar or dark matter streams should therefore evolve smoothly in time, simply stretching further away from the parent system. The structure observed in tidal streams, such as Palomar 5 must have an external origin, perhaps disk shocking or encounters with molecular clouds or dark matter substructures.
Our stability analysis could in principle also be extended to other systems producing streams. However, systems with non spherical shapes and net angular momentum are extremely difficult to analyse with analytical methods, since the alignment of the interacting objects is important. Merging disk galaxies for example produce streams with internal structures strongly depending on the initial alignment of the disks and on their angular momenta. In such cases, high resolution numerical simulations are the indispensable tool for a consistent stability analysis.
\section*{Acknowledgements}
We thank Sebastian Elser and George Lake for helpful discussions. This research is supported by the Swiss National Foundation.
|
\subsubsection*{Acknowledgements}
We thank Francesco D'Eramo, Andreas Hohenegger, Chris Jillings,
Alexander Kartavtsev, Rafael Lang, Maxim Pospelov, and Jesse Thaler
for helpful discussions and correspondence. Research at the Perimeter
Institute is supported in part by the Government of Canada through
NSERC and by the Province of Ontario through MEDT. Research at
Max-Planck-Institut f${\ddot{\it u}}$r Kernphysik is supported by
DFG-Sonderforschungsbereich Transregio~27.
|
\section{Introduction}
\IEEEPARstart{P}{seudocodewords} represent the intrinsic mechanism of
failure of binary linear codes under linear-programming (LP) or
message-passing (MP) decoding (see, e.g., \cite{KV-characterization,
KV-long-paper}). The concept of \emph{pseudoweight} of a
pseudocodeword was introduced in~\cite{Wiberg} and~\cite{FKKR} (see
also~\cite{KV-long-paper}) as an analog to the pertinent parameter in
the maximum likelihood (ML) decoding scenario, i.e., the signal
Euclidean distance in the case of the additive white Gaussian noise
channel (AWGNC), or the Hamming distance in the case of the binary
symmetric channel (BSC). Accordingly, for a binary linear code $\mathcal{C}$
and a parity-check matrix $\boldsymbol{H}$ of $\mathcal{C}$, the (AWGNC or BSC) minimum
pseudoweight $\mathsf{w}_{\min}(\boldsymbol{H})$ may be considered as a
first-order measure of decoder error-correcting performance for LP or
MP decoding. Another closely related measure is the max-fractional
weight, which we sometimes also call pseudoweight in order to simplify
statements; it serves as a lower bound on both AWGNC and BSC
pseudoweights.
In order to minimize the decoding error probability under LP (or MP)
decoding, one might want to select a matrix~$\bldH$ which maximizes the
minimum pseudoweight of the code for the given channel. Adding
redundant rows to the parity-check matrix introduces additional
constraints on the so-called \emph{fundamental cone}, and thus may
improve the performance of LP decoding and increase the minimum
pseudoweight.\footnote{We note that for message-passing iterative
decoding, apart from the case of decoding over the binary erasure
channel there is no general statement that additional parity-checks
are beneficial.} However, such additions increase the decoding
complexity under MP decoding, especially since linear combinations of
low-density rows may not yield a low-density result. On the other
hand, there exist classes of codes for which sparse parity-check
matrices exist with many redundant rows, e.g.,
\cite{Kou_Lin_Fossorier}.
For the AWGNC, BEC (binary erasure channel), BSC, and max-fractional
pseudoweights, define $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})$, $\rho_{\mbox{\tiny BEC}}(\mathcal{C})$,
$\rho_{\mbox{\tiny BSC}}(\mathcal{C})$, and $\rho_{\mbox{\tiny max-frac}}(\mathcal{C})$, respectively, to be
the minimum number of rows in any parity-check matrix $\bldH$ such that
the minimum pseudoweight of $\mathcal{C}$ with respect to this matrix is equal
to the code's minimum Hamming distance $d$. For the sake of
simplicity, we sometimes use the notation $\rho(\mathcal{C})$ when the type of
channel is clear from the context. The value $\rho(\mathcal{C})$ is called
the (AWGNC, BEC, BSC, or max-fractional) \emph{pseudocodeword
redundancy} (or pseudoredundancy) of $\mathcal{C}$. If for the code $\mathcal{C}$
there exists no such matrix $\bldH$, we say that the pseudoredundancy is
infinite.
The BEC pseudocodeword redundancy, which is equivalent to the
\emph{stopping redundancy}, is studied in~\cite{Schwartz_Vardy}, where
it is shown that for any linear code the BEC pseudoredundancy is
finite; the paper also contains bounds on $\rho_{\mbox{\tiny BEC}}(\mathcal{C})$ for
general binary linear codes, and for some specific families of codes.
These bounds were subsequently improved, for instance
in~\cite{Han_Siegel}. The study of BSC pseudoredundancy was initiated
in~\cite{Kelley_Sridhara}, where the authors presented bounds on
$\rho_{\mbox{\tiny BSC}}(\mathcal{C})$ for various families of codes.
In this work, we further investigate pseudoredundancy for the AWGNC,
BSC, and max-fractional pseudoweight. We show that for most codes
there exists no $\bldH$ such that the minimum pseudoweight (with respect
to $\bldH$) is equal to $d$, and therefore the AWGNC, BSC, and
max-fractional pseudocodeword redundancy (as defined above) is
infinite for most codes. For some code families for which the
pseudoredundancy is finite, we provide upper bounds on its value. We
consider in particular constructions of new codes from old and codes
based on designs. Furthermore, we compute the pseudocodeword
redundancies for all codes of small length (at most $9$), and we
investigate cyclic codes for which the eigenvalue bound of Vontobel
and Koetter~\cite{KV-lower-bounds} is sharp.
The outline of the paper is as follows. In Section
\ref{sec:general_settings} we provide detailed definitions and some
background information on LP decoding, pseudocodewords, the minimum
pseudoweight, and the pseudocodeword redundancy; we also discuss
related notions appearing in the literature. Subsequently, we show in
Section~\ref{sec:random_codes} that the pseudocodeword redundancy for
random codes is infinite with high probability. The next four
sections are concerned with upper bounds on the pseudoredundancy for
some particular classes of codes; we investigate punctured codes and
codes of dimension 2 in Section~\ref{sec:puncture}, constructions of
codes from other codes in Section~\ref{sec:inheritance}, parity-check
matrices of row-weight 2 in Section~\ref{sec:row_weight_2}, and codes
based on designs in Section~\ref{sec:designs}. The final two sections
are devoted to experimental results; Section~\ref{sec:short_codes}
examines the pseudocodeword redundancy for all codes of small length,
and Section~\ref{sec:kv_bound} deals with cyclic codes that meet the
eigenvalue bound on the minimum AWGNC pseudoweight by Vontobel and
Koetter.
\section{General Settings}\label{sec:general_settings}
Let $\mathbb{F}_2$ be the binary field and let $\mathbb{R}$ be the field of real
numbers. Addition and multiplication (including matrix-vector and
matrix-matrix multiplication) are carried out in $\mathbb{F}_2$ when the
operands are defined over $\mathbb{F}_2$, and in $\mathbb{R}$ when the operands are
defined over the reals. Occasionally, we will explicitly convert
elements in $\mathbb{F}_2$ into real numbers; in this case we identify
$0\in\mathbb{F}_2$ with $0\in\mathbb{R}$ and $1\in\mathbb{F}_2$ with $1\in\mathbb{R}$.
Let $\mathcal{C}$ be a code of length $n \in \mathbb{N}$ over the binary field
$\mathbb{F}_2$, defined by \[ \mathcal{C} = \ker \bldH = \{ \boldsymbol{c} \in \mathbb{F}_2^n \mid \bldH
\boldsymbol{c}^T = \boldsymbol{0}^T \} \] where $\bldH$ is an $m \times n$
\emph{parity-check matrix} over $\mathbb{F}_2$ of the code~$\mathcal{C}$. Obviously,
the code $\mathcal{C}$ may admit more than one parity-check matrix, and all
the codewords form a linear vector space of dimension $k \ge n-m$. We
say that $k$ is the \emph{dimension} of the code $\mathcal{C}$. We denote by
$d(\mathcal{C})$ (or just $d$) the minimum Hamming distance (also called the
minimum distance) of $\mathcal{C}$. The code $\mathcal{C}$ may then be referred to as
an $[n,k,d]$ linear code over $\mathbb{F}_2$.
Denote the set of column indices and the set of row indices of the
parity-check matrix $\bldH$ by $\mathcal{I} = \{1,\dots,n\}$ and $\mathcal{J} =
\{1,\dots,m\}$, respectively. For any row index $j\in\mathcal{J}$ we let
$\mathcal{I}_j \stackrel{\mbox{\tiny $\triangle$}}{=} \{ i \in \mathcal{I} \mid H_{j,i} \neq 0 \}$ denote the set of
the column indices where the parity-check matrix is nonzero; similarly
for any column index $i\in\mathcal{I}$ we let $\mathcal{J}_i \stackrel{\mbox{\tiny $\triangle$}}{=} \{ j \in \mathcal{J} \mid
H_{j,i} \neq 0 \}$ denote the corresponding set of row indices.
The matrix $\bldH$ is said to be \emph{$(w_c,w_r)$-regular} if
$|\mathcal{J}_i|=w_c$ for all $i\in\mathcal{I}$ and $|\mathcal{I}_j|=w_r$ for all $j\in\mathcal{J}$; a
$(w,w)$-regular matrix is also called simply \emph{$w$-regular}.
\subsection{LP decoding}
We give a brief review of LP decoding. Consider data transmission
over a memoryless binary-input output-symmetric channel with channel
law $p_{Y|X}(y|x)$. Based on the received vector $\boldsymbol{y} =
(y_1,\dots,y_n)$ we can define the log-likelihood-ratio vector
$\boldsymbol{\gamma} = (\gamma_1,\dots,\gamma_n)\in\mathbb{R}^n$ by $\gamma_i \stackrel{\mbox{\tiny $\triangle$}}{=}
\log(p_{Y|X}(y_i|0))-\log(p_{Y|X}(y_i|1))$ for $i\in\mathcal{I}$. Viewing the
code $\mathcal{C}$ canonically as a subset of $\{0,1\}^n\subset \mathbb{R}^n$, one
can then express ML decoding as the minimization problem
\[ \hat{\boldsymbol{x}} \stackrel{\mbox{\tiny $\triangle$}}{=} \argmin_{\boldsymbol{x}\in\mathcal{C}}\, \langle \boldsymbol{x}, \boldsymbol{\gamma}
\rangle \:. \]
This is equivalent to the linear programming problem
\[ \hat{\boldsymbol{x}} \stackrel{\mbox{\tiny $\triangle$}}{=} \argmin_{\boldsymbol{x}\in\operatorname{conv}(\mathcal{C})} \langle \boldsymbol{x},
\boldsymbol{\gamma} \rangle \:, \] where $\operatorname{conv}(\mathcal{C})$ denotes the convex hull
of $\mathcal{C}$ in $\mathbb{R}^n$. However, since the number of defining
hyperplanes of $\operatorname{conv}(\mathcal{C})$ usually grows exponentially with the block
length, this minimization problem becomes impractical.
Instead one might consider a relaxation of the above minimization
problem (see \cite{Feldman, Feldman-et-al, KV-long-paper}), where the
convex hull $\operatorname{conv}(\mathcal{C})$ is replaced by the so-called fundamental
polytope $\mathcal{P}(\bldH)$ to be defined next. For $j\in \mathcal{J}$, let $\boldsymbol{h}_j$
denote the $j$-th row of the parity-check matrix $\bldH$, and consider
the local code
\[ \mathcal{C}_j = \{ \boldsymbol{c} \in \mathbb{F}_2^n \mid \boldsymbol{h}_j \boldsymbol{c}^T = 0 \} \]
consisting of all binary vectors satisfying the $j$-th parity-check,
so that $\mathcal{C} = \bigcap_{j\in\mathcal{J}}\mathcal{C}_j$.
Then the \emph{fundamental polytope} $\mathcal{P} \stackrel{\mbox{\tiny $\triangle$}}{=} \mathcal{P}(\bldH)$ is defined as
\[ \mathcal{P} \stackrel{\mbox{\tiny $\triangle$}}{=} \bigcap_{j\in\mathcal{J}} \operatorname{conv}(\mathcal{C}_j) \:, \]
where again $\mathcal{C}_j$ is viewed as a subset of $\mathbb{R}^n$. Now \emph{LP
decoding} of a binary linear code $\mathcal{C}$ with parity-check matrix
$\bldH$ can be expressed as the minimization problem
\begin{equation}\label{eq:lp-decoding}
\hat{\boldsymbol{x}} \stackrel{\mbox{\tiny $\triangle$}}{=} \argmin_{\boldsymbol{x}\in\mathcal{P}}\, \langle \boldsymbol{x},
\boldsymbol{\gamma} \rangle \:,
\end{equation}
where $\mathcal{P} = \mathcal{P}(\bldH)$ denotes the fundamental polytope.
We note that $\operatorname{conv}(\mathcal{C})\subseteq\mathcal{P}$, where the inclusion is usually
proper. However, the number of defining hyperplanes of $\mathcal{P}$ is
typically much smaller than for $\operatorname{conv}(\mathcal{C})$, in particular for LDPC
codes, so that the corresponding linear programming problem becomes
tractable.
If $\mathcal{P}$ is strictly larger than $\operatorname{conv}(\mathcal{C})$ then it may happen that
the decoding rule~(\ref{eq:lp-decoding}) outputs a vertex\footnote{The
set of optimal solutions contains a vertex, and one may assume that
the output is a vertex.} of $\mathcal{P}$ that is not a vertex of
$\operatorname{conv}(\mathcal{C})$, i.e., not a codeword. Such vertices, called
pseudocodewords, are the reason for the suboptimality of LP decoding
with respect to ML decoding.
Note that the fundamental polytope $\mathcal{P}(\bldH)$ is dependent on the
parity-check matrix $\bldH$ rather than the code $\mathcal{C}$ itself, but we
always have $\mathcal{P}(\bldH)\cap \{0,1\}^n = \mathcal{C}$, cf.\ \cite{Feldman,
Feldman-et-al}.
\subsection{The fundamental cone and pseudoweights}
When analyzing LP decoding, we may assume without loss of generality
that the zero codeword $\boldsymbol{0}$ has been sent; then, given this
assumption, the probability of correct LP decoding depends only on the
conic hull of the fundamental polytope rather than on the fundamental
polytope itself (see \cite{Feldman, Feldman-et-al, KV-long-paper}).
The conic hull of the fundamental cone $\mathcal{P}(\bldH)$ is called the
\emph{fundamental cone} $\mathcal{K}(\bldH)$. More concretely, $\mathcal{K}(\bldH)$ is
given as the set of vectors $\boldsymbol{x}\in \mathbb{R}^n$ that satisfy
\begin{equation}\label{eq:fundcone-inequality-1}
\forall j \in \mathcal{J}, \; \forall \ell \in \mathcal{I}_j :\ \:
x_\ell \le \sum_{i \in \mathcal{I}_j \setminus \{ \ell \}} x_i \; ,
\end{equation}
\begin{equation}\label{eq:fundcone-inequality-2}
\forall i \in \mathcal{I} :\ \: x_i \ge 0 \; .
\end{equation}
The vectors $\boldsymbol{x}\in\mathcal{K}(\bldH)$ are called \emph{pseudocodewords}%
\footnote{Some authors consider only the vertices of the fundamental
polytope $\mathcal{P}(\bldH)$ as pseudocodewords, but we will use this more
general definition which includes all vectors of the fundamental
cone $\mathcal{K}(\bldH)$.} of~$\mathcal{C}$ with respect to the parity-check matrix
$\bldH$. Note again that the fundamental cone $\mathcal{K}(\bldH)$ depends on the
parity-check matrix $\bldH$ rather than on the code $\mathcal{C}$ itself. At
the same time, the fundamental cone is independent of the underlying
communication channel.
\begin{example}\label{exa:fund_cone}
Let $\mathcal{C}$ be the $[7,4,3]$ Hamming code with parity-check matrix
\[ \bldH = \mat{
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} } . \]
Then the fundamental cone inequalities read:
{\small
\begin{align*}
& x_1 \le x_2+x_3+x_5 \quad x_2\le x_3+x_4+x_6 \quad x_3\le x_4+x_5+x_7 \\
& x_2 \le x_1+x_3+x_5 \quad x_3\le x_2+x_4+x_6 \quad x_4\le x_3+x_5+x_7 \\
& x_3 \le x_1+x_2+x_5 \quad x_4\le x_2+x_3+x_6 \quad x_5\le x_3+x_4+x_7 \\
& x_5 \le x_1+x_2+x_3 \quad x_6\le x_2+x_3+x_4 \quad x_7\le x_3+x_4+x_5 \\
& 0 \le x_1 \quad 0 \le x_2 \quad 0 \le x_3 \quad 0 \le x_4
\quad 0 \le x_5 \quad 0 \le x_6 \quad 0 \le x_7
\end{align*}}
\end{example}
The influence of a nonzero pseudocodeword on the decoding performance
will be measured by its \emph{pseudoweight}, which depends on the
channel at hand. The BEC, AWGNC, BSC pseudoweights, and max-fractional
weight of a nonzero pseudocodeword $\boldsymbol{x} \in \mathcal{K}(\bldH)$ were defined
in~\cite{FKKR} and~\cite{KV-long-paper} as follows:
\begin{align*}
\mathsf{w}_{\mbox{\tiny BEC}} (\boldsymbol{x}) & \,\stackrel{\mbox{\tiny $\triangle$}}{=}\,
\left| \mbox{supp} ( \boldsymbol{x} ) \right| \; , \\
\mathsf{w}_{\mbox{\tiny AWGNC}} (\boldsymbol{x}) & \,\stackrel{\mbox{\tiny $\triangle$}}{=}\,
\frac{\left( \sum_{i \in \mathcal{I}} x_i \right)^2}{\sum_{i \in \mathcal{I}} x_i^2} \: .
\end{align*}
Let $\boldsymbol{x}'$ be a vector in $\mathbb{R}^n$ with the same components as
$\boldsymbol{x}$ but in non-increasing order. For $i-1 < \xi \le i$, where $1
\le i \le n$, let $\phi(\xi) \stackrel{\triangle}{=} x'_i$. Define
$\Phi(\xi) \stackrel{\mbox{\tiny $\triangle$}}{=} \int_{0}^{\xi} \phi(\xi') \; d \xi'$ and \[
\mathsf{w}_{\mbox{\tiny BSC}}(\boldsymbol{x}) \stackrel{\mbox{\tiny $\triangle$}}{=} 2\, \Phi^{-1} ( \Phi(n)/2 ) \; . \]
Finally, the max-fractional weight of $\boldsymbol{x}$ is defined as
\[\mathsf{w}_{\mbox{\tiny max-frac}} (\boldsymbol{x}) \,\stackrel{\mbox{\tiny $\triangle$}}{=}\,
\frac{\sum_{i \in \mathcal{I}} x_i}{\max_{i \in \mathcal{I}} x_i} \: .\]
Additionally, the pseudoweight of the all-zero vector is usually
defined to be zero, i.e., $\mathsf{w}(\boldsymbol{0}) = 0$, for all four
pseudoweights $\mathsf{w}$, but this is inessential for this paper.
Note that for binary vectors $\boldsymbol{x}\in\{0,1\}^n\setminus\{\boldsymbol{0}\}$
we have
\[ \mathsf{w}_{\mbox{\tiny BSC}}(\boldsymbol{x}) = \mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{x}) =
\mathsf{w}_{\mbox{\tiny BSC}}(\boldsymbol{x}) = \mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{x}) =
\mathsf{w}_{\mathrm{H}}(\boldsymbol{x}) \:, \] where $\mathsf{w}_{\mathrm{H}}(\boldsymbol{x})$
denotes the Hamming weight of $\boldsymbol{x}$.
\begin{example}
Let $\mathcal{C}$ and $\bldH$ be as in Example~\ref{exa:fund_cone}. The
vector $\boldsymbol{x} = (0,0,1,0,1,1,2)$ is a pseudocodeword in $\mathcal{K}(\bldH)$
with weights $\mathsf{w}_{\mbox{\tiny BEC}} (\boldsymbol{x}) = \left| \mbox{supp}
(\boldsymbol{x}) \right| = 4$ and $\mathsf{w}_{\mbox{\tiny AWGNC}} (\boldsymbol{x}) = \left(
\sum_{i \in \mathcal{I}} x_i \right)^2 / \sum_{i \in \mathcal{I}} x_i^2 = 25/7$.
Furthermore, $\mathsf{w}_{\mbox{\tiny BSC}}(\boldsymbol{x}) =
2\,\Phi^{-1}\big((\sum_{i\in\mathcal{I}}x_i)/2\big) = 2\,\Phi^{-1}(5/2) =
3$, where $\boldsymbol{x}'= (2,1,1,1,0,0,0)$, and finally,
$\mathsf{w}_{\mbox{\tiny max-frac}} (\boldsymbol{x}) = \sum_{i \in \mathcal{I}} x_i / \max_{i \in
\mathcal{I}} x_i = 5/2$.
\end{example}
We define the BEC \emph{minimum pseudoweight} of the code $\mathcal{C}$ with
respect to the parity-check matrix $\bldH$ as
\[ \mathsf{w}_{\min}^{\mbox{\tiny BEC}} (\bldH) \,\stackrel{\mbox{\tiny $\triangle$}}{=}\, \min_{\boldsymbol{x}\in\mathcal{K}(\bldH)
\setminus \{ \boldsymbol{0} \} } \mathsf{w}_{\mbox{\tiny BEC}} (\boldsymbol{x}) \; .\] The
quantities $\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} (\bldH) $, $\mathsf{w}_{\min}^{\mbox{\tiny BSC}}
(\bldH) $ and $\mathsf{w}_{\min}^{\mbox{\tiny max-frac}} (\bldH)$ are defined similarly.
We note that the considered pseudoweights are invariant under scaling
by a positive scalar, and that a minimum is indeed attained on
$\mathcal{K}(\bldH)\setminus\{0\}$ (see \cite[Sect.~6]{KV-long-paper}). When
the type of pseudoweight is clear from the context, we sometimes use
the notation $\mathsf{w}_{\min} (\bldH)$. Note that all four minimum
pseudoweights are upper bounded by~$d$, the code's minimum distance.
\subsection{Pseudocodeword redundancy}
Given a code $\mathcal{C}$ we will define the pseudocodeword redundancy as the
minimum number of rows in a parity-check matrix $\bldH$ for $\mathcal{C}$ such
that the corresponding minimum pseudoweight equals the minimum
distance.
So for a binary linear $[n,k,d]$ code $\mathcal{C}$ we define the BEC
\emph{pseudocodeword redundancy} of the code $\mathcal{C}$ as \[
\rho_{\mbox{\tiny BEC}}(\mathcal{C}) \,\stackrel{\mbox{\tiny $\triangle$}}{=}\, \inf\{\#\text{rows}(\bldH) \mid
\ker\bldH=\mathcal{C}\,,\, \mathsf{w}_{\min}^{\mbox{\tiny BEC}}(\bldH)=d\} \: ,\] where
$\inf\varnothing\stackrel{\mbox{\tiny $\triangle$}}{=}\infty$, and similarly we define the
pseudocodeword redundancies $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})$, $\rho_{\mbox{\tiny BSC}}(\mathcal{C})$,
and $\rho_{\mbox{\tiny max-frac}}(\mathcal{C})$ for the AWGNC and BSC pseudoweights, and
the max-fractional weight. When the type of pseudocodeword redundancy
is clear from the context, we sometimes use the notation $\rho(\mathcal{C})$.
We remark that all pseudocodeword redundancies satisfy $\rho(\mathcal{C}) \ge
r\stackrel{\mbox{\tiny $\triangle$}}{=} n-k$.
\begin{example}
Let $\mathcal{C}$ be the $[7,4,3]$ Hamming code. Then:
\begin{gather*}
\rho_{\mbox{\tiny max-frac}}(\mathcal{C}) = 7 \; \ge \;
\rho_{\mbox{\tiny AWGNC}}(\mathcal{C}) = 3 \; \ge \;
\rho_{\mbox{\tiny BEC}}(\mathcal{C}) = 3 \\
\rho_{\mbox{\tiny max-frac}}(\mathcal{C}) = 7 \; \ge \;
\rho_{\mbox{\tiny BSC}}(\mathcal{C}) = 4 \; \ge \;
\rho_{\mbox{\tiny BEC}}(\mathcal{C}) = 3
\end{gather*}
The following matrices $\bldH_3$, $\bldH_4$, and $\bldH_7$ are examples
for parity-check matrices with a minimum number of rows such that
$\mathsf{w}_{\min}^{\mbox{\tiny BEC}}(\bldH_3) = \mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH_3) = 3$,
$\mathsf{w}_{\min}^{\mbox{\tiny BSC}}(\bldH_4) = 3$, and
$\mathsf{w}_{\min}^{\mbox{\tiny max-frac}}(\bldH_7) = 3$ holds.
\begin{gather*}
\bldH_3 = \mat{
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} } \\[1ex]
\bldH_4 = \mat{
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} } \\[1ex]
\bldH_7 = \mat{
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} }
\end{gather*}
These matrices were found by computer search, see
Section~\ref{sec:short_codes}.
\end{example}
We describe the behavior of the pseudocodeword redundancy and the
minimum pseudoweight for a given binary linear $[n,k,d]$ code~$\mathcal{C}$
by introducing four classes of codes:\smallskip
\begin{description}[\IEEEsetlabelwidth{{\bf (class 0)}}]
\item[{\bf (class 0)}] $\rho(\mathcal{C})$ is infinite, i.e., there is no
parity-check matrix~$\bldH$ with $d=\mathsf{w}_{\min}(\bldH)$,
\item[{\bf (class 1)}] $\rho(\mathcal{C})$ is finite, but $\rho(\mathcal{C})>r$,
\item[{\bf (class 2)}] $\rho(\mathcal{C})=r$, but $\mathcal{C}$ is not in
class 3,
\item[{\bf (class 3)}] $d=\mathsf{w}_{\min}(\bldH)$ for \emph{every}
parity-check matrix $\bldH$ of~$\mathcal{C}$.
\end{description}
Note that if a code has infinite pseudocodeword redundancy, then LP
decoding for this code can never achieve the ML decoding performance;
on the other hand, if a code's pseudocodeword redundancy is finite,
its value gives a (very approximate) indication of the LP decoding
complexity required to achieve this bound. Note that this is a
fundamental complexity associated with the code, and not tied to a
particular parity-check matrix. We leave it as a direction for further
research to provide more general definitions which capture the average
complexity-performance tradeoff of LP decoding as more redundant rows
are added to the parity-check matrix.
\subsection{Basic Connections}
The different minimum pseudoweights are related as follows. This
result is taken from~\cite{KV-long-paper}.
\begin{lemma}\label{lemma:relations}
Let $\mathcal{C}$ be a binary linear code with the parity-check matrix
$\bldH$. Then,
\begin{gather*}
\mathsf{w}_{\min}^{\mbox{\tiny max-frac}} (\bldH) \; \le \;
\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} (\bldH) \; \le \;
\mathsf{w}_{\min}^{\mbox{\tiny BEC}} (\bldH) \; , \\
\mathsf{w}_{\min}^{\mbox{\tiny max-frac}} (\bldH) \; \le \;
\mathsf{w}_{\min}^{\mbox{\tiny BSC}} (\bldH) \; \le \;
\mathsf{w}_{\min}^{\mbox{\tiny BEC}} (\bldH) \; .
\end{gather*}
\end{lemma}
As a straightforward corollary we obtain the following theorem, which
relates the different pseudoredundancies.
\begin{theorem}\label{thm:pseudoredundancies}
Let $\mathcal{C}$ be a binary linear code. Then,
\begin{gather*}
\rho_{\mbox{\tiny max-frac}} (\mathcal{C}) \; \ge \; \rho_{\mbox{\tiny AWGNC}} (\mathcal{C}) \; \ge \;
\rho_{\mbox{\tiny BEC}} (\mathcal{C}) \; , \\
\rho_{\mbox{\tiny max-frac}} (\mathcal{C}) \; \ge \; \rho_{\mbox{\tiny BSC}} (\mathcal{C}) \; \ge \;
\rho_{\mbox{\tiny BEC}} (\mathcal{C}) \; .
\end{gather*}
\end{theorem}
\subsection{Related Notions}\label{sec:related}
As mentioned in the introduction, Schwartz and Vardy consider
in~\cite{Schwartz_Vardy} the so-called stopping distance of a binary
linear code given by a parity-check matrix, and the stopping
redundancy of a binary linear code. With \cite[Proposition
51]{KV-long-paper} it is easy to see that the stopping distance equals
the minimum BEC pseudoweight, and thus the stopping redundancy is
equivalent to the BEC pseudocodeword redundancy.
Besides pseudocodewords, the notion of \emph{trapping set}
\cite{trapping} is another concept for analyzing the performance of
binary linear codes under MP decoding. In \cite{LHMH-trapping} the
{\em trapping redundancy} for binary linear codes is introduced as a
generalization of the stopping redundancy, and several upper bounds
are presented.\smallskip
In~\cite{Kashyap} a binary linear code $\mathcal{C}$ is called
\emph{geometrically perfect} if it admits a parity-check matrix $\bldH$
such that the fundamental polytope equals the convex hull of the code,
i.e., $\mathcal{P}(\bldH) = \operatorname{conv}(\mathcal{C})$.
In this case ML decoding can be exactly described as an instance of LP
decoding. Kashyap~\cite[Theorem~VI.2]{Kashyap} gave a
characterization of all geometrically perfect codes: a binary linear
code $\mathcal{C}$ is geometrically perfect if and only if $\mathcal{C}$ does not
contain as a minor\footnote{A \emph{minor} of a code ${\mathcal{C}}$ is any
code obtained from ${\mathcal{C}}$ by a (possibly empty) sequence of
shortening and puncturing operations.} any code equivalent to
certain codes $\mathcal{C}_1$, $\mathcal{C}_2$, $\mathcal{C}_3$ with parameters $[7,3,4]$,
$[10,5,4]$, and $[10,4,4]$, respectively.
It is easy to see that for geometrically perfect codes all four
pseudocodeword redundancies are finite.\smallskip
Smarandache and Vontobel~\cite{Smarandache_Vontobel} define the
\emph{pseudoweight spectrum gap} for a binary linear code $\mathcal{C}$ given
by a parity-check matrix $\bldH$ as follows. The set $\mathcal{M}(\bldH)$ of all
\emph{minimal pseudocodewords} is defined as the set of all vectors
$\boldsymbol{x}\in\mathbb{R}^n$ that lie on an edge of the fundamental cone
$\mathcal{K}(\bldH)$. Now let $\mathcal{M}'(\bldH)$ denote the set of all minimal
pseudocodewords that are not scalar multiples of codewords
$\boldsymbol{c}\in\mathcal{C}$, and let $\mathsf{w}$ be any of the BEC, AWGNC, BSC, or
max-fractional pseudoweight. Then the pseudoweight spectrum gap is
the quantity \[ g(\bldH) \stackrel{\mbox{\tiny $\triangle$}}{=} \min_{\boldsymbol{x}\in\mathcal{M}'(\bldH)} \mathsf{w}(\boldsymbol{x})
- d(\mathcal{C}) \:. \] It is apparent that $g(\bldH)\ge \mathsf{w}_{\min}(\bldH) -
d(\mathcal{C})$, and we have $\mathsf{w}_{\min}(\bldH) = d(\mathcal{C})$ if and only if
$g(\bldH)\ge 0$.
If the pseudoweight spectrum gap $g(\bldH)$ is strictly positive then
the LP decoding performance approaches ML decoding performance as the
signal-to-noise ratio goes to infinity. To date, only few examples of
interesting codes with positive pseudoweight spectrum gap are known;
these include the codes based on the Euclidean plane or the projective
plane \cite[Theorem~8]{Smarandache_Vontobel}.
\section{Pseudoredundancy of Random Codes}\label{sec:random_codes}
In this section we show that for most binary linear codes the AWGNC
and BSC pseudoredundancies are infinite. We begin with the following
lemma.
\begin{lemma}\label{lemma:awgn}
For a binary linear code $\mathcal{C}$ of length $n$, let $d^{\perp}$ be
the minimum distance of the dual code. Then, the minimum AWGNC
pseudoweight of $\mathcal{C}$ (with respect to any parity-check matrix
$\bldH$) satisfies
\begin{equation}\label{eq:pseudo-distance-bound-random}
\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} \le
\frac{(n + d^{\perp} - 2)^2}{(d^{\perp}-1)^2 + (n-1)} \; .
\end{equation}
\end{lemma}
\begin{IEEEproof}
Consider the pseudocodeword $\boldsymbol{x} = (x_1, x_2, \dots, x_n) \stackrel{\mbox{\tiny $\triangle$}}{=}
(d^{\perp} \!\!-\! 1 \,,\, 1 \,,\, \dots, 1 )$. Since $d^{\perp}$ is the
minimum distance of the dual code, every row in $\bldH$ has weight at
least $d^{\perp}$. Therefore, all
inequalities~(\ref{eq:fundcone-inequality-1})
and~(\ref{eq:fundcone-inequality-2}) are satisfied for this $\boldsymbol{x}$,
and so it is indeed a legal pseudocodeword. Finally, observe that
the AWGNC pseudoweight of $\boldsymbol{x}$ is given by the right-hand side of
(\ref{eq:pseudo-distance-bound-random}).
\end{IEEEproof}
In the sequel, we use the term \emph{random code} for a binary linear
code $\mathcal{C}$ whose $k \times n$ generator matrix contains independently
and uniformly distributed random entries from $\mathbb{F}_2$. The following
result is known as the Gilbert-Varshamov bound. If we pick a code by
selecting the generator matrix entries at random, the resulting code
$\mathcal{C}$ has rate $R = k/n$ and relative minimum distance $\delta$, such
that
\[ \delta \ge \mathsf{H}_2^{-1} (1 - R) - \epsilon \; , \] with
probability approaching $1$ as $n \rightarrow \infty$, for any fixed
small $\epsilon > 0$, where $\mathsf{H}^{-1}_2(\cdot)$ is the inverse of
the binary entropy function $\mathsf{H}_2 (p) = - p \log_2 p - (1 - p)
\log_2 (1 - p)$ for $p\in[0\,,\, 1/2]$. A similar result also holds
when the code $\mathcal{C}$ is defined by selecting the parity-check matrix
entries (independently and uniformly) at random.
Let $R = k/n$ be fixed. Then, if we select at random a $k \times n$
matrix over $\mathbb{F}_2$, which corresponds to a code $\mathcal{C}$, the relative
minimum distance of $\mathcal{C}$ is at least $\mathsf{H}_2^{-1} (1 - R) -
\epsilon$ (with probability approaching~$1$ as $n \rightarrow \infty$)
and the relative minimum distance of the dual code of $\mathcal{C}$ is at
least $\mathsf{H}_2^{-1} (R) - \epsilon$ (again, with probability
approaching~$1$ as $n \rightarrow \infty$). By taking the
intersection of these two events, both the code and the dual code have
relative minimum distances which are $\epsilon$-close to the
Gilbert-Varshamov bound with probability approaching~$1$ as $n
\rightarrow \infty$. (The reader can refer to~\cite[Theorems 4.4,
4.5, and 4.10]{Roth-book} and to~\cite[Theorem 8 and Exercise
3]{Guruswami-notes}.)
To this end, we take a random binary linear code $\mathcal{C}$
of arbitrary length $n$ (for $n \rightarrow \infty$) with $R = k/n$.
The dual code $\mathcal{C}^{\perp}$ of $\mathcal{C}$, with probability close to one, has
rate $R^{\perp}=1-R$ and relative minimum distance $\delta^{\perp} = d^{\perp}/n$ that
attains the Gilbert-Varshamov bound
\[ \delta^{\perp} \ge \mu \stackrel{\mbox{\tiny $\triangle$}}{=} \mathsf{H}^{-1}_2 (1 - R^{\perp}) - \epsilon
= \mathsf{H}^{-1}_2(R) - \epsilon \; , \]
Note that~(\ref{eq:pseudo-distance-bound-random}) may be written in
terms of the relative minimum distance $\delta^{\perp}$ of the dual code
as follows:
\begin{equation}
\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} \le
\frac{(1 + \delta^{\perp}\! - 2/n)^2}
{(\delta^{\perp}\!- 1/n)^2 + (1/n-1/n^2)} \; .
\label{eq:pseudo-distance-new}
\end{equation}
Hence, for large $n$, the minimum pseudoweight of the code $\mathcal{C}$ is
bounded from above by $(1 + 1/\delta^{\perp})^2 + \epsilon' \le
(1+1/\mu)^2 + \epsilon'$ for some small $\epsilon' > 0$, and this
bound does not depend on $n$. On the other hand, $\mathcal{C}$ is a random
code and so its minimum distance satisfies the Gilbert-Varshamov
bound, namely
\[ d\ge \left( \mathsf{H}^{-1}_2 (1 - R) - \epsilon \right) \cdot n \;
,\] which increases linearly with $n$ for a fixed $R$. This
immediately establishes the following theorem.
\begin{theorem}\label{thm:non-existence_AWGNC}
Let $0<R<1$ be fixed. For a random binary linear code $\mathcal{C}$ of
length $n$ and rate~$R$, there is, with probability approaching~$1$
as $n$ tends to infinity, a gap between the minimum AWGNC
pseudoweight (with respect to any parity-check matrix) and the
minimum distance. Therefore, the AWGNC pseudoredundancy is infinite
for most codes.
\end{theorem}
\begin{remark}
The result in Theorem~\ref{thm:non-existence_AWGNC} is different
from, but related to, the results in Propositions~49 and
Corollary~50 in~\cite{KV-long-paper}, where it was shown that the
minimum AWGN pseudoweight of ensembles of regular LDPC codes grows
sublinearly in the code length. Indeed, there are three fundamental
differences between our results and~\cite{KV-long-paper}: (i) We do
not assume anything about the density of the parity-check matrix
$\bldH$. We also use the fact that the dual code of the random code
is asymptotically good; for a regular LDPC code this is not
true. (ii) We consider the fundamental cone, which is formed by all
possible linear combinations of the rows of $\bldH$; by contrast, the
authors of~\cite{KV-long-paper} consider only the case when the
column weight of $\bldH$ is smaller than its row weight. (iii) We show
that the minimum pseudoweight of the considered ensemble is bounded
from above by a constant, while in~\cite{KV-long-paper} this
quantity is shown to be bounded by a sublinear function.
\end{remark}\smallskip
The following lemma is a counterpart of Lemma~\ref{lemma:awgn} for the
BSC.
\begin{lemma}\label{lemma:bsc}
Let $\mathcal{C}$ be a binary linear code of length $n$, and let $d^{\perp}$
be the minimum distance of the dual code. Then, the minimum BSC
pseudoweight of $\mathcal{C}$ (with respect to any parity-check matrix
$\bldH$) satisfies
\[ \mathsf{w}_{\min}^{\mbox{\tiny BSC}} \le 2 \lceil n/{d^{\perp}} \rceil \; . \]
\end{lemma}
\begin{IEEEproof}
Consider the pseudocodeword
\[ \boldsymbol{x} = (x_1, x_2, \dots, x_n) \stackrel{\mbox{\tiny $\triangle$}}{=} (\underbrace{
d^{\perp}\!\!-\!1, \dots, d^{\perp}\!\!-\!1}_\tau \,,\,
\underbrace{1, \dots, 1}_{n-\tau}) \; , \]
for some positive integer $\tau$. This vector $\boldsymbol{x}$ is then a
legal pseudocodeword; since $d^{\perp}$ is the minimum distance of the
dual code, every row in $\bldH$ has a weight of at least $d^{\perp}$, and
so, all inequalities~(\ref{eq:fundcone-inequality-1})
and~(\ref{eq:fundcone-inequality-2}) are satisfied by this $\boldsymbol{x}$.
If $\tau(d^{\perp}\!-1) \ge n - \tau$ then by the definition of the BSC
pseudoweight $\mathsf{w}_{\mbox{\tiny BSC}} (\boldsymbol{x}) \le 2 \tau$. This condition is
equivalent to $\tau d^{\perp} \ge n$. Therefore, we set $\tau = \lceil
n/{d^{\perp}} \rceil$. For the corresponding vector $\boldsymbol{x}$, the
pseudoweight is less or equal to $2\tau = 2 \lceil n/{d^{\perp}}
\rceil$.
\end{IEEEproof}
Similarly to the AWGNC case, let $\mathcal{C}$ be a random binary linear code
of length $n$ with $R=k/n$. The parameters $R^{\perp}$
and~$\delta^{\perp}$ of its dual code $\mathcal{C}^{\perp}$ attain with high
probability the Gilbert-Varshamov bound $\delta^{\perp}\ge\mu$.
From Lemma~\ref{lemma:bsc}, for all $n$, the pseudoweight of the
code~$\mathcal{C}$ is bounded from above by
\[2 \lceil n/{d^{\perp}} \rceil < 2/\delta^{\perp} + 2 \le 2/\mu + 2 \; ,\]
which is a constant. On the other hand, $\mathcal{C}$ is a random code and
its minimum distance also satisfies the Gilbert-Varshamov bound, so it
increases linearly with $n$. This proves the following theorem.
\begin{theorem}\label{thm:non-existence_BSC}
Let $0<R<1$ be fixed. For a random binary linear code $\mathcal{C}$ of
length $n$ and rate~$R$, there is, with probability approaching~$1$
as $n$ tends to infinity, a gap between the minimum BSC pseudoweight
(with respect to any parity-check matrix) and the minimum distance.
Therefore, the BSC pseudoredundancy is infinite for most codes.
\end{theorem}
The last theorem disproves the conjecture in~\cite{Kelley_Sridhara}
that the BSC pseudoredundancy is finite for all binary linear
codes.\footnote{We note that a slightly different definition of BSC
pseudoweight was given in~\cite{Kelley_Sridhara}, but the statement
of Lemma~\ref{lemma:bsc} and thus
Theorem~\ref{thm:non-existence_BSC} hold with the same proof also
with respect to this definition.}
\begin{example}
Consider the [23,12] Golay code having minimum distance $d=7$. The
minimum distance of its dual code is $d^{\perp}=8$. We can take a
pseudocodeword $\boldsymbol{x}$ as in the proof of Lemma~\ref{lemma:bsc} with
$\tau = \lceil n/{d^{\perp}} \rceil = 3$. We have $\mathsf{w}_{\mbox{\tiny BSC}}
(\boldsymbol{x}) \le 2 \tau = 6$, thus obtaining that the minimum distance is
not equal to the minimum pseudoweight.
Similarly, for the [24,12] extended Golay code we have $d=d^{\perp}=8$,
and by taking $\tau = \lceil n/{d^{\perp}} \rceil = 3$ we obtain
$\mathsf{w}_{\mbox{\tiny BSC}} (\boldsymbol{x}) \le 2\tau = 6$.
Note however that the presented techniques do not answer the
question of whether these Golay codes have finite AWGNC
pseudoredundancy.
\end{example}
In the context of the extended Golay code we mention that there other
interesting graphical representations of codes than by Tanner graphs;
in particular, a minimal \emph{tail-biting trellis} has been
constructed for the extended Golay code in \cite{TBT}. The
pseudoweights of its pseudocodewords are investigated in \cite{FKKR},
where it is shown that there are pseudocodewords with a BSC
pseudoweight of $6$; on the other hand, as far as we know, it is still
unknown whether there are nonzero pseudocodewords of the tail-biting
trellis with an AWGNC pseudoweight of less than $8$.
We have seen in this section that the AWGNC pseudoredundancy and the
BSC pseudoredundancy of a random binary linear code is infinite. From
Theorem~\ref{thm:pseudoredundancies} it follows that this holds also
for the pseudoredundancy with respect to the max-fractional weight.
\section{Basic Upper Bounds}\label{sec:puncture}
Whereas a random code has infinite pseudoredundancy for the AWGNC and
the BSC, there are several families of codes for which the
pseudoredundancy is finite. Sections~\ref{sec:puncture},
\ref{sec:inheritance}, \ref{sec:row_weight_2}, and~\ref{sec:designs}
deal with upper bounds on the pseudoredunancy for some particular
classes of codes.
We start with this section considering two basic situations, namely
the puncturing of zero coordinates and codes of minimum distance~$2$.
The following results hold with respect to the BEC, AWGNC, and BSC
pseudoweights, and the max-fractional weight.
\begin{lemma}\label{lemma:puncturing}
Let $\mathcal{C}$ be an $[n,k,d]$ code having $t$ zero coordinates, and let
$\mathcal{C}'$ be the $[n-t,k,d]$ code obtained by puncturing~$\mathcal{C}$ at these
coordinates. Then
\[ \rho(\mathcal{C}')\le \rho(\mathcal{C})\le \rho(\mathcal{C}')+t \:. \]
\end{lemma}
\begin{IEEEproof}
For notational purposes, we identify $\mathbb{R}^n$ with $\mathbb{R}^{\mathcal{I}}$, and
for $\boldsymbol{x}\in\mathbb{R}^{\mathcal{I}}$ and some subset $\mathcal{I}'\subseteq\mathcal{I}$ we let
$\boldsymbol{x}|_{\mathcal{I}'}\in\mathbb{R}^{\mathcal{I}'}$ be the projection of $\boldsymbol{x}$ onto the
coordinates in $\mathcal{I}'$.
Let $\mathcal{I}'\subseteq\mathcal{I}$ be the set of nonzero coordinates of the
code~$\mathcal{C}$. To prove the first inequality, let $\bldH$ be a
$\rho\times n$ parity-check matrix for $\mathcal{C}$. Consider its
$\rho\times(n-t)$ submatrix $\bldH'$ consisting of the columns
corresponding to $\mathcal{I}'$. Then $\bldH'$ is a parity-check matrix for
$\mathcal{C}'$, and \[\mathcal{K}(\bldH') = \{\boldsymbol{x}|_{\mathcal{I}'} \mid \boldsymbol{x}\in\mathcal{K}(\bldH),\
\boldsymbol{x}|_{\mathcal{I}\setminus\mathcal{I}'}=\boldsymbol{0}\} \:.\] Therefore,
$\mathsf{w}_{\min}(\bldH')\ge \mathsf{w}_{\min}(\bldH)$, and this proves
$\rho(\mathcal{C}')\le\rho(\mathcal{C})$.
For the second inequality, let $\bldH'$ be a $\rho'\times(n-t)$
parity-check matrix for $\mathcal{C}'$. Now we consider a $(\rho'+t)\times
n$ matrix $\bldH$ with the following properties: The upper
$\rho'\times n$ submatrix of $\bldH$ consists of the columns of $\bldH'$
at positions $\mathcal{I}'$ and of zero-columns at positions
$\mathcal{I}\setminus\mathcal{I}'$, and the lower $t\times n$ submatrix consists of
rows of weight~$1$ that have $1$s at the positions
$\mathcal{I}\setminus\mathcal{I}'$. Then $\mathcal{C}=\ker\bldH$ and
\[ \mathcal{K}(\bldH) = \{ \boldsymbol{x}\in\mathbb{R}^\mathcal{I} \mid \boldsymbol{x}|_{\mathcal{I}'}\in\mathcal{K}(\bldH'),\
\boldsymbol{x}|_{\mathcal{I}\setminus\mathcal{I}'}=\boldsymbol{0}\}\:. \] Consequently,
$\mathsf{w}_{\min}(\bldH) = \mathsf{w}_{\min}(\bldH')$, and this proves
$\rho(\mathcal{C})\le \rho(\mathcal{C}')+t$.
\end{IEEEproof}
\begin{lemma}\label{lemma:distance_two}
Let $\mathcal{C}$ be a code of minimum distance $d\le 2$. Then
$d=\mathsf{w}_{\min}(\bldH)$ for any parity-check matrix $\bldH$ of $\mathcal{C}$,
i.e., $\mathcal{C}$ is in class $3$ (for BEC, AWGNC, BSC, and max-fractional
pseudoweight).
\end{lemma}
\begin{IEEEproof}
By Lemma~\ref{lemma:relations} it suffices to prove this lemma for
the max-fractional weight $\mathsf{w} = \mathsf{w}_{\mbox{\tiny max-frac}}$. Since
$\mathsf{w}(\boldsymbol{x})\ge 1$ holds for all nonzero pseudocodewords, we
always have ${\mathsf{w}_{\min}(\bldH)\ge 1}$, which proves the result in
the case $d=1$.
Let $d=2$ and $\bldH$ be a parity-check matrix for $\mathcal{C}$. Let
$\boldsymbol{x}\in\mathcal{K}(\bldH)$ and let $x_{\ell}$ be the largest coordinate.
Since $d=2$ there is no zero column in $\bldH$ and thus there exists a
row $j$ with $\ell\in\mathcal{I}_j$. Then $x_{\ell}\le
\sum_{i\in\mathcal{I}\setminus\{\ell\}}x_i$, hence
$2x_{\ell}\le\sum_{i\in\mathcal{I}}x_i$, and thus $\mathsf{w}(\boldsymbol{x})\ge 2$. It
follows $\mathsf{w}_{\min}(\bldH)\ge 2$ and the lemma is proved.
\end{IEEEproof}
\section{Constructions of codes from other codes}%
\label{sec:inheritance}
The following results consider the pseudoredundancy of codes obtained
from other codes by the direct sum or the $(\boldsymbol{u}\,\boldsymbol{u})$
construction. They are analogs of Theorems~7 and 8
in~\cite{Schwartz_Vardy}, and Theorems~4.1 and 4.2
in~\cite{Kelley_Sridhara}, for the case of the max-fractional weight
and the AWGNC pseudoweight. Our proofs in each case follow the
exposition of these earlier proofs.
\begin{theorem}\label{thm:concatenation}
Let $\mathcal{C}_1$ and $\mathcal{C}_2$ be $[n_1,k_1,d_1]$ and $[n_2,k_2,d_2]$
binary linear codes, respectively. Then the direct sum $\mathcal{C}_3 = \{
(\boldsymbol{u} \; \boldsymbol{v}) \mid \boldsymbol{u} \in \mathcal{C}_1, \boldsymbol{v} \in \mathcal{C}_2 \}$ is an
$[n_1 + n_2,k_1 + k_2,\min \{ d_1,d_2 \} ]$ code with
\begin{gather*}
\rho_{\mbox{\tiny max-frac}}(\mathcal{C}_3) \le \rho_{\mbox{\tiny max-frac}}(\mathcal{C}_1) +
\rho_{\mbox{\tiny max-frac}}(\mathcal{C}_2) \:, \\
\rho_{\mbox{\tiny AWGNC}}(\mathcal{C}_3) \le \rho_{\mbox{\tiny AWGNC}}(\mathcal{C}_1) +
\rho_{\mbox{\tiny AWGNC}}(\mathcal{C}_2) \:.
\end{gather*}
\end{theorem}
\begin{IEEEproof}
Without loss of generality, we may assume that both $\rho(C_1)$ and
$\rho(C_2)$ are finite, for otherwise the statement to be proved is
trivial. For $i=1,2$, let $\bldH_i$ be a parity-check matrix for
$\mathcal{C}_i$ having $\rho(\mathcal{C}_i)$ rows and such that $\mathsf{w}(\boldsymbol{x}) \ge
d_i$ for all $\boldsymbol{x} \in \mathcal{K}(\bldH_i) \setminus \{ \boldsymbol{0} \}$. Then
\[ \bldH_3 = \left[ \begin{array}{cc}
\bldH_1 & \boldsymbol{0} \\
\boldsymbol{0} & \bldH_2
\end{array} \right] \]
is a parity-check matrix for $\mathcal{C}_3$ with $\rho(\mathcal{C}_1) +
\rho(\mathcal{C}_2)$ rows. Let $\boldsymbol{p} = (\boldsymbol{q} \; \boldsymbol{r}) \in \mathcal{K}( \bldH_3 )
\setminus \{ \boldsymbol{0} \}$, where the vectors $\boldsymbol{q}$ and $\boldsymbol{r}$ in
the concatenation have lengths $n_1$ and $n_2$ respectively. Then,
we may assume $\boldsymbol{q} \in \mathcal{K}( \bldH_1 )\setminus \{ \boldsymbol{0} \}$
and $\boldsymbol{r} \in \mathcal{K}( \bldH_2 )\setminus \{ \boldsymbol{0} \}$, and therefore
$\mathsf{w}(\boldsymbol{q}) \ge d_1$ and $\mathsf{w}(\boldsymbol{r}) \ge d_2$.
(Note that in the case where either $\boldsymbol{q}$ or $\boldsymbol{r}$ is equal to
$\boldsymbol{0}$, the result is trivial since for any $\boldsymbol{q} \ne
\boldsymbol{0}$, $\mathsf{w}(\boldsymbol{q} \; \boldsymbol{0}) = \mathsf{w}(\boldsymbol{q})$
for both the max-fractional weight and the AWGNC pseudoweight.)
We consider the two cases of max-fractional weight and
AWGNC pseudoweight separately.
\emph{Max-fractional weight:} Assume without loss of generality that
$\max \{ q_i \} \ge \max \{ r_i \}$. Then
\begin{multline*}
\mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{p}) = \frac{\sum p_i}{\max \{ p_i \} }
= \frac{\sum q_i + \sum r_i}{\max \{ q_i \} } \\
> \frac{\sum q_i}{\max \{ q_i \} } \ge d_1 \ge \min \{ d_1, d_2 \}
\end{multline*}
which proves the result.
\emph{AWGNC pseudoweight:} Assume without loss of generality that
$\mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{q}) \ge \mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{r})$; this
condition may be written as
\begin{equation}
\bigg( \sum_{i=1}^{n_1} q_i \bigg)^{\!2} \bigg( \sum_{i=1}^{n_2}
r_i^2 \bigg) \ge \bigg( \sum_{i=1}^{n_1} q_i^2 \bigg)
\bigg( \sum_{i=1}^{n_2} r_i \bigg)^{\!2} \:.
\label{eq:ineq_1_thm_AWGNC}
\end{equation}
To establish the result, we need only to prove that
$\mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{p}) \ge \mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{r})$. Now, since
the entries of $\boldsymbol{q}$ and $\boldsymbol{r}$ are nonnegative, we have
\begin{equation}
2 \bigg( \sum_{i=1}^{n_1} q_i \bigg) \bigg( \sum_{i=1}^{n_2} r_i
\bigg) \bigg( \sum_{i=1}^{n_2} r_i^2 \bigg) \ge 0 \; .
\label{eq:ineq_2_thm_AWGNC}
\end{equation}
Adding $\big( \sum_{i=1}^{n_2} r_i \big)^2 \big( \sum_{i=1}^{n_2}
r_i^2 \big)$ to both sides of \eqref{eq:ineq_2_thm_AWGNC} and adding
the resulting inequality to inequality \eqref{eq:ineq_1_thm_AWGNC}
yields
\[
\bigg( \sum_{i=1}^{n_1} q_i + \sum_{i=1}^{n_2} r_i \bigg)^{\!\!2}
\bigg( \sum_{i=1}^{n_2} r_i^2 \bigg)
\ge \bigg( \sum_{i=1}^{n_2} r_i \bigg)^{\!\!2} \bigg( \sum_{i=1}^{n_1}
q_i^2 + \sum_{i=1}^{n_2} r_i^2 \bigg)
\]
which may be rearranged as $\mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{p}) \ge
\mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{r})$, as desired.
\end{IEEEproof}
\begin{theorem}\label{thm:self-concatenation}
Let $\mathcal{C}_1$ be an $[n,k,d]$ binary linear code. Then $\mathcal{C}_2 = \{
(\boldsymbol{u} \; \boldsymbol{u}) \mid \boldsymbol{u} \in \mathcal{C}_1 \}$ is a $[2n, k, 2d]$ code with
\begin{gather*}
\rho_{\mbox{\tiny max-frac}}(\mathcal{C}_2) \le \rho_{\mbox{\tiny max-frac}}(\mathcal{C}_1) + n \:, \\
\rho_{\mbox{\tiny AWGNC}}(\mathcal{C}_2) \le \rho_{\mbox{\tiny AWGNC}}(\mathcal{C}_1) + n \:.
\end{gather*}
\end{theorem}
\begin{IEEEproof}
As before, without loss of generality, we may assume that
$\rho(C_1)$ is finite. Let $\bldH_1$ be a parity-check matrix for
$\mathcal{C}_1$ with $\rho(\mathcal{C}_1)$ rows and such that $\mathsf{w}(\boldsymbol{x}) \ge
d_1$ for all $\boldsymbol{x} \in \mathcal{K}(\bldH_1) \setminus \{ \boldsymbol{0} \}$. Then
\[ \bldH_2 = \left[ \begin{array}{cc}
\bldH_1 & \boldsymbol{0} \\
\bldI_n & \bldI_n
\end{array} \right] \]
is a parity-check matrix for $\mathcal{C}_2$ with $\rho(\mathcal{C}_1) + n$ rows
(here $\bldI_n$ denotes the $n \times n$ identity matrix). Let $\boldsymbol{p}
= (\boldsymbol{q} \; \boldsymbol{r}) \in \mathcal{K}(\bldH_2) \setminus \{ \boldsymbol{0} \}$, where
the vectors $\boldsymbol{q} \in \mathcal{K}(\bldH_1)$ and $\boldsymbol{r}$ in the concatenation
both have length $n$. Then, for $i=1,2,\ldots,n$, from the
fundamental cone inequalities for row $n+i$ we get $q_i \le r_i \le
q_i$, so we have $\boldsymbol{p} = (\boldsymbol{q} \; \boldsymbol{q})$. Now, since $\boldsymbol{q} \in
\mathcal{K}(\bldH_1) \setminus \{ \boldsymbol{0} \}$, we have $\mathsf{w}(\boldsymbol{q}) \ge
d_1$. Since $\mathsf{w}((\boldsymbol{q} \; \boldsymbol{q})) = 2\, \mathsf{w}(\boldsymbol{q})$ for
both the max-fractional weight and the AWGNC pseudoweight, we have
$\mathsf{w}(\boldsymbol{p}) \ge 2d$, and the result follows.
\end{IEEEproof}
\begin{remark}
Theorem~9 in~\cite{Schwartz_Vardy} and Theorem~4.3
in~\cite{Kelley_Sridhara} state that if $\mathcal{C}$ is an $[n,k,3]$ binary
linear code then the extended $[n\!+\!1,k,4]$ code $\mathcal{C}'$ satisfies
$\rho(C')\le 2\rho(\mathcal{C})$, for the BEC pseudoweight and the BSC
pseudoweight, respectively. Regarding the corresponding results for
the case of the max-fractional weight and the AWGNC pseudoweight, we
mention here only that the analogous result in fact does not hold
for the case of the max-fractional weight. As a counterexample,
consider the $[7,4,3]$ Hamming code $\mathcal{C}_1$ which satisfies
$\rho(\mathcal{C}_1) \le 2^3-1 = 7$ (cf.\ Proposition
\ref{prop:Hamming}). On the other hand, the $[8,4,4]$ extended
Hamming code $\mathcal{C}_2$ satisfies $\rho_{\mbox{\tiny max-frac}}(\mathcal{C}_2) = \infty$
(cf.\ Section \ref{sec:short_codes_results}).
\end{remark}
\section{Parity-check matrices with rows of weight~$2$}%
\label{sec:row_weight_2}
In this section we consider the pseudoredundancy of codes with a
parity-check matrix consisting of rows of weight~$2$ and at most one
additional row. The results are then applied to upper-bound the
pseudoredundancy for codes of dimension~$2$. The basic case is dealt
with in the following lemma.
\begin{lemma}\label{lemma:row_weight_2}
Let $\bldH$ be a parity-check matrix of $\mathcal{C}$ such that every row in
$\bldH$ has weight~$2$. Then:
\begin{enumerate}
\item[(a)] There is an equivalence relation on the set $\mathcal{I}$ of column
indices of $\bldH$ such that for a vector $\boldsymbol{x}\in\mathbb{R}^n$ with
non-negative coordinates we have $\boldsymbol{x}\in\mathcal{K}(\bldH)$ if and only if
$\boldsymbol{x}$ has equal coordinates within each equivalence class.
\item[(b)] The minimum distance of $\mathcal{C}$ is equal to its minimum
BEC, AWGNC, BSC, and max-fractional pseudoweights with respect to
$\bldH$, i.e., $d(\mathcal{C}) = \mathsf{w}_{\min}(\bldH)$.
\end{enumerate}
\end{lemma}
\begin{IEEEproof}
For (a), define the required relation $R$ as follows: For
$i,i'\in\mathcal{I}$ let $(i,i')\in R$ if and only if $i=i'$ or there exists
an integer $\ell \ge 1$, column indices $i=i_0, i_1, \dots,
i_{\ell-1}, i_{\ell}=i'\in\mathcal{I}$ and row indices $j_1,\dots,j_l\in\mathcal{J}$
such that
\[ \{ i_0, i_1 \} = \mathcal{I}_{j_1} \, , \,
\{ i_1, i_2 \} = \mathcal{I}_{j_2} \, , \, \dots \, , \,
\{ i_{\ell-1}, i_{\ell} \} = \mathcal{I}_{j_{\ell}} \; . \]
This is an equivalence relation, and it defines equivalence
classes over $\mathcal{I}$. It is easy to check that
inequalities~(\ref{eq:fundcone-inequality-1}) imply that
$\boldsymbol{x}\in\mathcal{K}(\bldH)$ if and only if $x_i=x_{i'}$ for any $(i,i')\in
R$.
In order to prove (b), we note that the minimum (BEC, AWGNC, BSC or
max-fractional) pseudoweight is always bounded above by the minimum
distance of $\mathcal{C}$, so we only have to show that the minimum
pseudoweight is bounded below by the minimum distance.
Let $\mathcal{S} = \{S_1, S_2, \dots, S_t\}$ be the set of equivalence
classes of $R$, and let $d_S = |S|$ for $S\in\mathcal{S}$. It is easy to
see that the minimum distance of $\mathcal{C}$ is $d = \min_{S\in\mathcal{S}}d_S$
(since the minimum weight nonzero codeword of $\mathcal{C}$ has non-zeros in
the coordinates corresponding to a set $S\in\mathcal{S}$ of minimal size and
zeros everywhere else).
Now let $\boldsymbol{x}\in\mathcal{K}(\bldH)$. Since the coordinates $x_i$, $i\in\mathcal{I}$,
depend only on the equivalence classes, we may use the notation
$x_S$, $S\in\mathcal{S}$. Let $x_T$, $T\in\mathcal{S}$, be the largest
coordinate. Then:
\[ \mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{x}) = \frac{\sum_{i \in \mathcal{I}} x_i}{x_T} \ge
\frac{\sum_{i\in T}x_i}{x_T} = |T|=d_T \ge d \:. \] Therefore,
$\mathsf{w}_{\min}^{\mbox{\tiny max-frac}} (\bldH) \ge d$, and by using
Lemma~\ref{lemma:relations}, we get
$\mathsf{w}_{\min}^{\mbox{\tiny BEC}}(\bldH)\ge d$, $\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH)\ge
d$, and $\mathsf{w}_{\min}^{\mbox{\tiny BSC}}(\bldH)\ge d$.
\end{IEEEproof}
The following proposition is a stronger version of
Lemma~\ref{lemma:row_weight_2}.
\begin{proposition}\label{prop:row_weight_2}
Let $\bldH$ be an $m \times n$ parity-check matrix of $\mathcal{C}$, and
assume that the $m\!-\!1$ first rows in $\bldH$ have weight~$2$.
Denote by $\widehat{\bldH}$ the $(m\!-\!1) \times n$ matrix consisting
of these rows, consider the equivalence relation of
Lemma~\ref{lemma:row_weight_2} (a) with respect to $\widehat{\bldH}$,
and assume that $\mathcal{I}_m$ intersects each equivalence class in at most
one element. Then, the minimum distance of $\mathcal{C}$ is equal to its
minimum BEC, AWGNC, BSC, and max-fractional pseudoweights with
respect to $\bldH$, i.e., $d(\mathcal{C}) = \mathsf{w}_{\min}(\bldH)$.
\end{proposition}
\begin{IEEEproof}
Let $\mathcal{S}$ be the set of classes of the aforementioned equivalence
relation on $\mathcal{I}$, and let $d_S=|S|$ for $S\in\mathcal{S}$. Let
\[ \mathcal{S}' = \{ S \in \mathcal{S} \mid\: |S \cap \mathcal{I}_m| = 1 \} \:. \] Also
let $\mathcal{S}'' = \mathcal{S} \setminus \mathcal{S}'$, so that $S \cap \mathcal{I}_m =
\varnothing$ for all $S \in \mathcal{S}''$.
Let $\boldsymbol{x}\in\mathcal{K}(\bldH) \setminus \{ \boldsymbol{0} \}$. As before, since
the coordinates $x_i$, $i\in\mathcal{I}$, depend only on the equivalence
classes, we may use the notation $x_S$, $S\in\mathcal{S}$. The fundamental
cone constraints \eqref{eq:fundcone-inequality-1} and
\eqref{eq:fundcone-inequality-2} may then be written as $x_S \ge 0$
for all $S \in \mathcal{S}$ and
\begin{equation}
\forall R \in \mathcal{S}' :\ x_R \le \sum_{S \in \mathcal{S}'\setminus\{R\}} x_S \; ,
\label{eq:fundcone_ineq_S'}
\end{equation}
respectively, and the max-fractional weight of $\boldsymbol{x} \in \mathcal{K}(\bldH)
\setminus \{ \boldsymbol{0} \}$ is given by
\begin{equation}
\mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{x})
= \frac{\sum_{S \in \mathcal{S}} d_S x_S}{\max_{S \in \mathcal{S}} x_S} \; .
\label{eq:new_maxfrac}
\end{equation}
Suppose $\boldsymbol{x}\in\mathcal{K}(\bldH) \setminus \{ \boldsymbol{0} \}$ has minimal
max-fractional weight. Let $x_T$ be its largest coordinate. First
note that if there exists $R \in \mathcal{S}''\setminus\{T\}$ with $x_R > 0$,
setting $x_R$ to zero results in a new pseudocodeword with lower
max-fractional weight, which contradicts the assumption that $\boldsymbol{x}$
achieves the minimum. Therefore $x_R = 0$ for all $R \in
\mathcal{S}''\setminus\{T\}$. We next consider two cases.
\emph{Case 1:} $T \in \mathcal{S}''$. If there exists $R \in \mathcal{S}'$ with $x_R >
0$, setting all such $x_R$ to zero results in a new pseudocodeword
with lower max-fractional weight, which contradicts the minimality of
the max-fractional weight of $\boldsymbol{x}$. Therefore $x_T$ is the only
positive coordinate of $\boldsymbol{x}$, and by \eqref{eq:new_maxfrac} the
max-fractional weight of $\boldsymbol{x}$ is $d_T$.
\emph{Case 2:} $T \in \mathcal{S}'$. In this case $x_R = 0$ for all $R \in
\mathcal{S}''$. From inequality~(\ref{eq:fundcone_ineq_S'}) for $R=T$ we
obtain
\[ x_T \le \sum_{S\in\mathcal{S}'\setminus\{T\}} x_S \:. \]
With $d_0 \stackrel{\mbox{\tiny $\triangle$}}{=} \min_{S\in\mathcal{S}'\setminus\{T\}}d_S$ it follows that
\[ d_0 x_T \le \sum_{S\in\mathcal{S}'\setminus\{T\}} d_0 x_S \le
\sum_{S\in\mathcal{S}'\setminus\{T\}} d_Sx_S \:. \] Consequently, \[
(d_T+d_0)x_T \le \sum_{S\in\mathcal{S}} d_Sx_S \:, \] and thus
$\mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{x}) \ge d_T+d_0$. We conclude that the minimum
max-fractional weight is given by
\[ \mathsf{w}_{\min}^{\mbox{\tiny max-frac}} (\bldH) = \min \left\{ \min_{S, T \in \mathcal{S}',
S \neq T} \{d_S + d_T\} \; , \; \min_{S \in \mathcal{S}''} \{d_S \}
\right\} \:. \] But this is easily seen to be equal to the minimum
distance $d$ of the code.
Finally, by using Lemma~\ref{lemma:relations}, we obtain that
$\mathsf{w}_{\min}^{\mbox{\tiny BEC}}(\bldH)=d$, $\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH)=d$ and
$\mathsf{w}_{\min}^{\mbox{\tiny BSC}}(\bldH)=d$.
\end{IEEEproof}
\begin{remark}
The requirement that all $i \in \mathcal{I}_m$ belong to different
equivalence classes of $\widehat{\bldH}$ in
Proposition~\ref{prop:row_weight_2} is necessary. Indeed, consider
the matrix
\[ \bldH = \mat{
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
} \:. \]
One can see that there are two equivalence classes for
$\widehat{\bldH}$: $S_1 = \{ 1, 2, 3 \}$, $S_2 = \{ 4 \}$. The minimum
distance of the corresponding code $\mathcal{C}$ is $4$ (since $(1, 1, 1, 1)$
is the only nonzero codeword). However, $\boldsymbol{x} = (1, 1, 1, 3) \in
\mathcal{K}(\bldH)$ is a pseudocodeword of max-fractional weight $2$.
\end{remark}
\begin{corollary}\label{corollary:dimension_two}
Let $\mathcal{C}$ be a code of length~$n$ and dimension~$2$. Then
$\rho(\mathcal{C})=n-2$, i.e., $\mathcal{C}$ is of class at least~$2$ (for BEC,
AWGNC, BSC, and max-fractional pseudoweight).
\end{corollary}\pagebreak
\begin{IEEEproof}
We consider two cases.
{\em Case 1: $\mathcal{C}$ has no zero coordinates.}
Let $\boldsymbol{c}_1$ and $\boldsymbol{c}_2$ be two linearly independent codewords
of $\mathcal{C}$. Define the following subsets of $\mathcal{I}$:
\begin{eqnarray*}
S_1 & \stackrel{\mbox{\tiny $\triangle$}}{=} & \{ i \in \mathcal{I} \mid
i \in \mbox{supp}(\boldsymbol{c}_1)
\mbox{ and } i \notin \mbox{supp}(\boldsymbol{c}_2) \} \; \\
S_2 & \stackrel{\mbox{\tiny $\triangle$}}{=} &
\{ i \in \mathcal{I} \mid
i \notin \mbox{supp}(\boldsymbol{c}_1)
\mbox{ and } i \in \mbox{supp}(\boldsymbol{c}_2) \} \; \\
S_3 & \stackrel{\mbox{\tiny $\triangle$}}{=} & \{ i \in \mathcal{I} \mid
i \in \mbox{supp}(\boldsymbol{c}_1)
\mbox{ and } i \in \mbox{supp}(\boldsymbol{c}_2) \} .
\end{eqnarray*}
The sets $S_1$, $S_2$, and $S_3$ are pairwise disjoint. Since $\mathcal{C}$
has no zero coordinates, $\mathcal{I} = S_1 \cup S_2 \cup S_3$. The
ordering of the elements in $\mathcal{I}$ implies an ordering on the
elements in each of $S_1$, $S_2$, and $S_3$. Assume that $S_1 = \{
i_1, i_2, \cdots, i_{|S_1|}\}$ and $i_1 < i_2 < \cdots < i_{|S_1|}$.
If $S_1 \neq \varnothing$, let $m_1 = i_1$ be the minimal element in
$S_1$, and define an $(|S_1| - 1) \times n$ matrix $\bldH_1 =
(H^1_{j,\ell})$ as follows:
\[ H^1_{j,\ell} = \left\{ \begin{array}{cl}
1 & \mbox{ if } i_j = \ell \mbox{ or } i_{j+1} = \ell \; , \\
& \qquad \qquad j = 1, 2, \cdots, |S_1|-1 \; , \\
0 & \mbox{ otherwise } \; .
\end{array} \right. \]
Similarly, define $(|S_2| - 1) \times n$ and $(|S_3| - 1) \times n$
matrices $\bldH_2$ and $\bldH_3$, with respect to $S_2$ and $S_3$.
(Some of the $S_i$s might be equal to $\varnothing$, in which case
the corresponding $\bldH_i$ is not defined.) Let $m_2$
and $m_3$ be minimal elements of $S_2$ and $S_3$, respectively (if $S_2
\neq \varnothing$ and $S_3 \neq \varnothing$).
{\em Subcase 1-a: One of $S_1$, $S_2$, $S_3$ is empty.} Without
loss of generality we may assume that $S_3 = \varnothing$, i.e.,
that $c_1$ and $c_2$ have disjoint support; indeed, if for example
$S_1 = \varnothing$, then $\mathrm{supp}(c_1) \subseteq
\mathrm{supp}(c_2)$ and we can replace $c_2$ by $c_1 + c_2$. Define
an $(n-2) \times n$ matrix $\bldH$ by $\bldH^T \stackrel{\mbox{\tiny $\triangle$}}{=} [ \, \bldH_1^T \; |
\; \bldH_2^T \,]$. It is easy to see that all rows of $\bldH$ are
linearly independent, and so its rank is $n-2$. It is also
straightforward that for all $\boldsymbol{c} \in \mathcal{C}$ we have $\boldsymbol{c} \in
\ker(\bldH)$. Therefore, $\bldH$ is a parity-check matrix of~$\mathcal{C}$. The
matrix $\bldH$ has a form as in Lemma~\ref{lemma:row_weight_2}, and
thus $\rho(\mathcal{C}) = n-2$.
{\em Subcase 1-b: Neither of $S_1$, $S_2$, $S_3$ is empty.}
Define a $1 \times n$ matrix $\bldH_4 = (H^4_{j,\ell})$, where
\[ H^4_{1,\ell} = \left\{ \begin{array}{cl}
1 & \mbox{ if } S_j \neq \varnothing \mbox{ and } m_j = \ell \\
& \qquad \qquad \mbox{ for } j = 1, 2, 3 \; , \\
0 & \mbox{ otherwise } \; .
\end{array} \right. \]
Additionally, define an $(n-2) \times n$ matrix $\bldH$ by $\bldH^T
\stackrel{\mbox{\tiny $\triangle$}}{=} [ \, \bldH_1^T \; | \; \bldH_2^T \; | \; \bldH_3^T \; | \; \bldH_4^T
\,]$. Similarly to the previous case, all rows of $\bldH$ are
linearly independent, its rank is $n-2$. For all $\boldsymbol{c} \in \mathcal{C}$ we
have $\boldsymbol{c} \in \ker(\bldH)$. Therefore, $\bldH$ is a parity-check
matrix of~$\mathcal{C}$.
The matrix $\bldH$ has a form as in
Proposition~\ref{prop:row_weight_2} (where $S_1$, $S_2$, and $S_3$
are corresponding equivalence classes over $\mathcal{I}$), and therefore
$\rho(\mathcal{C}) = n-2$.
{\em Case 2: $\mathcal{C}$ has $t>0$ zero coordinates.}
Consider a code $\mathcal{C}'$ of length $n-t$ obtained by puncturing~$\mathcal{C}$
in these $t$ zero coordinates. From Case 1 (with respect to $\mathcal{C}'$),
$\rho(C') = n-t-2$. By applying the rightmost inequality in
Lemma~\ref{lemma:puncturing}, we have $\rho(C) \le n-2$. Since
$k=2$, we conclude that $\rho(C) = n-2$.
\end{IEEEproof}
\section{Codes Based on Designs}\label{sec:designs}
Among the codes with finite pseudoredundancy an interesting class of
codes is based on designs. In this section we consider {\em partial
designs}, which include the common BIBDs (also called $2$-designs).
We present a principal lower bound on the minimum pseudoweight for
codes, when the parity-check matrix is the block-point incidence
matrix of a partial design. We apply this bound to the Hamming codes
and the simplex codes and deduce that their pseudoredundancy is
finite.
\begin{definition}\label{prop:design_incidence_matrix}
A \emph{partial $(w_c,\lambda)$ design} is a block design consisting
of an $n$-element set $\mathcal{V}$ (whose elements are called
\emph{points}) and a collection of $m$ subsets of $\mathcal{V}$ (called
\emph{blocks}) such that every point is contained in exactly $w_c$
blocks and every $2$-element subset of $\mathcal{V}$ is contained in at most
$\lambda$ blocks. The \emph{incidence matrix} of a design is an $m
\times n$ matrix $\bldH$ whose rows correspond to the blocks and whose
columns correspond to the points, and that satisfies $H_{j,i} = 1$
if block $j$ contains point~$i$, and $H_{j,i} = 0$ otherwise.
If each block contains the same number $w_r$ of points and every
$2$-element subset of $\mathcal{V}$ is contained in exactly $\lambda$
blocks, the design is said to be an $(n,w_r,\lambda)$ \emph{balanced
incomplete block design} (BIBD), or \emph{$2$-design}.
\end{definition}
In the following we avoid the trivial cases $n\le 1$ and $\lambda=0$.
For a BIBD we have $n\, w_c = m\, w_r$ and also
\[ w_c\,(w_r-1) = \lambda\,(n-1) \]
(see, e.g., \cite[p.~60]{MacWilliams_Sloane}), so $(n,w_r,\lambda)$
determines the other parameters $w_c$ and $m$ by
\begin{equation}\label{eq:design_constraint}
w_c = \frac{n-1}{w_r-1}\,\lambda
\quad\text{and}\quad
m = \frac{n\,(n-1)}{w_r\,(w_r-1)}\,\lambda \:.
\end{equation}
Note that \cite{Vasic_Milenkovic} and \cite{Kashyap_Vardy} consider
parity-check matrices based on BIBDs; these matrices are the transpose
of the incidence matrices defined here.
We have the following general result for codes based on partial
$(w_c,\lambda)$ designs.
\begin{theorem}\label{thm:pseudoweight_bound}
Let $\mathcal{C}$ be a code with parity-check matrix $\bldH$, such that a
subset of the rows of $\bldH$ forms the incidence matrix for a partial
$(w_c,\lambda)$ design. Then the minimum max-fractional weight of
$\mathcal{C}$ with respect to $\bldH$ is lower bounded by
\begin{equation}
\mathsf{w}_{\min}^{\mbox{\tiny max-frac}} \ge 1 + \frac{w_c}{\lambda} \; .
\label{eq:new_lower_bound}
\end{equation}
For the case of an $(n,w_r,\lambda)$ BIBD, the lower bound in
(\ref{eq:new_lower_bound}) may also be written as
\[ \mathsf{w}_{\min}^{\mbox{\tiny max-frac}} \ge 1 + \frac{n-1}{w_r-1} \:; \]
the alternative form follows directly from
(\ref{eq:design_constraint}).
\end{theorem}
\begin{IEEEproof}
Consider the subset of the rows of $\bldH$ which forms the incidence
matrix for a partial $(w_c,\lambda)$ design. Let $\boldsymbol{x}$ be a
nonzero pseudocodeword and let $x_\ell$ be a maximal coordinate of
$\boldsymbol{x}$ ($\ell \in \mathcal{I}$). For all $j \in \mathcal{J}_\ell$, sum
inequalities (\ref{eq:fundcone-inequality-1}). We have
\begin{equation*}
w_c x_{\ell} \le \lambda\! \sum_{i \in \mathcal{I} \setminus \{ \ell \} } x_i \; ,
\end{equation*}
and thus
\begin{equation}\label{eq:basic_step_for_bound_bsc}
\left( 1 + \frac{w_c}{\lambda} \right) x_{\ell} \le
\sum_{i \in \mathcal{I}} x_i \; .
\end{equation}
The result now easily follows from the definition of
$\mathsf{w}_{\min}^{\mbox{\tiny max-frac}}$.
\end{IEEEproof}
\begin{theorem}\label{thm:pseudoweight_bound_2}
Let $\mathcal{C}$ be a code with parity-check matrix $\bldH$, such that a
subset of the rows of $\bldH$ forms the incidence matrix for a partial
$(w_c,\lambda)$ design. Then,
\begin{align*}
\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} & \ge 1 + \frac{w_c}{\lambda} \; , \\
\mathsf{w}_{\min}^{\mbox{\tiny BSC}} & \ge 1 + \frac{w_c}{\lambda} \; .
\end{align*}
\end{theorem}
\begin{IEEEproof}
Apply Lemma~\ref{lemma:relations} and
Theorem~\ref{thm:pseudoweight_bound}.
\end{IEEEproof}
Results similar to Theorem~\ref{thm:pseudoweight_bound} and
Theorem~\ref{thm:pseudoweight_bound_2} were also presented and proven
by Xia and Fu \cite{Xia_Fu} in the AWGNC case.
\begin{remark}\label{rem:spectrum_gap}
Under the conditions of Theorem~\ref{thm:pseudoweight_bound_2}, if
$\boldsymbol{x}\in\mathcal{K}(\bldH)$ is a nonzero pseudocodeword such that
$\mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{x}) = 1+\frac{w_c}{\lambda}$ holds then it
follows that $\boldsymbol{x}$ is a scalar multiple of a binary vector. This
can be easily seen by considering the proof of the inequality
$\mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{x})\ge \mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{x})$ (see
\cite[Lemma 44]{KV-long-paper}) and examining when equality
$\mathsf{w}_{\mbox{\tiny AWGNC}}(\boldsymbol{x}) = \mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{x})$ holds.
Furthermore it can be shown that in this case $\boldsymbol{x}$ is actually a
scalar multiple of a codeword (see \cite[Theorem 3]{Xia_Fu}). It
follows that the AWGNC pseudocodeword spectrum gap is positive,
provided that $d(\mathcal{C}) = 1+\frac{w_c}{\lambda}$ holds.
\end{remark}
Another tool for proving lower bounds on the minimum AWGNC
pseudoweight is provided by the following eigenvalue-based lower bound
by Vontobel and Koetter \cite{KV-lower-bounds}.
\begin{proposition}[cf.\ \cite{KV-lower-bounds}]\label{prop:KV_bound}
The minimum AWGNC pseudoweight for a $(w_c,w_r)$-regular
parity-check matrix $\bldH$ whose corresponding Tanner graph is
connected is bounded below by
\begin{equation}\label{eq:KV_bound}
\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} \ge n \cdot \frac{ 2w_c - \mu_2 }
{\mu_1 - \mu_2 } \; ,
\end{equation}
where $\mu_1$ and $\mu_2$ denote the largest and second largest
eigenvalue (respectively) of the matrix $\bldL \stackrel{\mbox{\tiny $\triangle$}}{=} \bldH^T \bldH$;
here, $\bldL$ and the matrix multiplication are to be considered over
the reals.
\end{proposition}
In the case where $\bldH$ is equal to the incidence matrix for an
$(n,w_r,\lambda)$ BIBD, the bound of Proposition \ref{prop:KV_bound}
becomes
\begin{equation}\label{eq:KV_bound_2}
\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} \ge 1 + \frac{ w_c }{ \lambda } \; ,
\end{equation}
so that in this case the bound of Proposition~\ref{prop:KV_bound}
coincides with that of Theorem \ref{thm:pseudoweight_bound_2} (for the
case of the AWGNC only).
To see why (\ref{eq:KV_bound}) becomes (\ref{eq:KV_bound_2}), denote
the column $i\in\mathcal{I}$ of $\bldH$ by $\boldsymbol{h}_i$ and denote the matrix $\bldL
= \left( L_{i,\ell} \right)_{i,\ell \in \mathcal{I}} = \bldH^T \bldH$. From the
properties of a BIBD we get
\[ L_{i,\ell} = \boldsymbol{h}_i^T \boldsymbol{h}_{\ell} = \begin{cases}
w_c & \text{if } i = \ell \:, \\
\lambda & \text{if } i \ne \ell \:.
\end{cases} \] Now, $\bldL$ has largest eigenvalue $\mu_1 = w_r w_c$ and
only one other eigenvalue $\mu_2 = w_c-\lambda$, whose multiplicity is
$n-1$, since one can write $\bldL = \lambda\boldsymbol{1} + (w_c-\lambda)
\boldsymbol{I}$, where $\boldsymbol{1}$ and $\boldsymbol{I}$ denote the all-ones and the
identity matrices, respectively. Now we have $2w_c-\mu_2 =
w_c+\lambda$ and $\mu_1-\mu_2 = w_rw_c - w_c + \lambda = n\lambda$, so
that $n \cdot \frac{ 2w_c - \mu_2 } {\mu_1 - \mu_2 } = 1 +
\frac{w_c}{\lambda}$.\smallskip
\begin{remark}
Prominent examples for codes based on designs are codes based on
Euclidean or projective geometries, in particular the $[{4^s-1}\,,
{4^s-3^s}\,, {2^s+1}]$ code based on the Euclidean plane ${\rm
EG}(2, 2^s)$ as well as the $[{4^s+2^s+1}\,, {4^s-3^s+2^s}\,,
{2^s+2}]$ code based on the projective plane ${\rm PG}(2, 2^s)$
(see~\cite{Kou_Lin_Fossorier, Smarandache_Vontobel}).
Theorem~\ref{thm:pseudoweight_bound_2} and
Remark~\ref{rem:spectrum_gap} apply to these codes, as their
standard parity-check matrices form the incidence matrix for a
partial design with parameters $(w_c,\lambda)=(2^s,1)$ and
$(w_c,\lambda)=({2^s+1}\,,1)$, respectively\footnote{In the latter
case the partial design is even a BIBD with parameters
$({4^s\!+\!2^s\!+\!1}\,, {2^s\!+\!1}\,, 1)$.}; in particular these
codes have finite pseudoredundancy.
\end{remark}
We next apply the bounds of Theorems~\ref{thm:pseudoweight_bound}
and~\ref{thm:pseudoweight_bound_2} to some other examples of codes
derived from designs.
\begin{proposition}\label{prop:Hamming}
For $m \ge 2$, the $[{2^m-1}\,, {2^m-1-m}\,, 3]$ Hamming code has
BEC, AWGNC, BSC, and max-fractional pseudocodeword redundancies \[
\rho(\mathcal{C}) \le 2^m-1 \:. \]
\end{proposition}
\begin{IEEEproof}
For $m \ge 2$, consider the binary parity-check matrix $\bldH$ whose
rows are exactly the nonzero codewords of the dual code $\mathcal{C}^{\perp}$,
in this case the $[{2^m-1}\,,m\,,{2^{m-1}}\,]$ simplex code. This
$\bldH$ is the incidence matrix for a BIBD with parameters
$(n,w_r,\lambda)$ = $({2^m-1}\,, {2^{m-1}}\,, {2^{m-2}})$.
Theorem~\ref{thm:pseudoweight_bound} gives $\mathsf{w}_{\mbox{\tiny max-frac}}
(\boldsymbol{x}) \ge 3$, leading to $\rho_{\mbox{\tiny max-frac}} (\mathcal{C}) \le 2^m-1$.
The result for BEC, AWGNC, and BSC follows by applying
Theorem~\ref{thm:pseudoredundancies}.
\end{IEEEproof}
In the next example, we consider simplex codes. Straightforward
application of the previous reasoning does not lead to the desired
result. However, more careful selection of the matrix $\bldH$, as
described below, leads to a new bound on the pseudoredundancy.
\begin{proposition}\label{prop:simplex}
For $m \ge 2$, the $[{2^m-1}\,, m\,, {2^{m-1}}]$ simplex code has
BEC, AWGNC, BSC, and max-fractional pseudocodeword redundancies
\begin{eqnarray*}
\rho(\mathcal{C}) \le \frac{(2^m-1)\,(2^{m-1}-1)}{3} \; .
\end{eqnarray*}
\end{proposition}
\begin{IEEEproof}
For $m \ge 2$, consider the binary parity-check matrix $\bldH$ whose
rows are exactly the codewords of the dual code $\mathcal{C}^{\perp}$ (in this
case the $[{2^m-1}\,, {2^m-1-m}\, ,3]$ Hamming code) with Hamming
weight equal to $3$. This $\bldH$ is the incidence matrix for a BIBD
with parameters $(n,w_r,\lambda)$ = $({2^m-1}\,, 3\,, 1)$.
Theorem~\ref{thm:pseudoweight_bound} gives
$\mathsf{w}_{\min}^{\mbox{\tiny max-frac}} \ge 2^{m-1}$.
Note that the number of codewords of weight~$3$ in the $[{2^m-1}\,,
{2^m-1-m}\,, 3]$ Hamming code equals ${(2^m-1)}{(2^{m-1}-1)}/3$.
One can show this, e.g., by considering the full sphere-packing of
the perfect Hamming code and observing that each codeword of weight
$3$ covers exactly $3$ vectors of weight $2$, of which there are
$(2^m-1)(2^m-2)/2$ in total.
Next, we justify the claim that $\bldH$ is a parity-check matrix of
$\mathcal{C}$. A theorem of Simonis~\cite{Simonis} states that if there
exists a linear $[n,k,d]$ code then there also exists a linear
$[n,k,d]$ code whose codewords are spanned by the codewords of
weight~$d$. Since the Hamming code is unique for the parameters
$[{2^m-1}\,, {2^m-1-m}\,, 3]$, this implies that the Hamming code
itself is spanned by the codewords of weight $3$, so the rowspace of
$\bldH$ equals $\mathcal{C}$.
The result for BEC, AWGNC, and BSC follows again by applying
Theorem~\ref{thm:pseudoredundancies}.
\end{IEEEproof}
We remark that the bounds of Propositions~\ref{prop:Hamming}
and~\ref{prop:simplex} are sharp at least for the case $m=3$ and the
max-fractional weight, see Section~\ref{sec:short_codes_results}.
The following proposition proves that the AWGNC, BSC, and
max-fractional pseudocodeword redundancies are finite for all codes
$\mathcal{C}$ with minimum distance at most $3$.
\begin{proposition}\label{prop:D3_codes}
Let $\mathcal{C}$ be a $[n,k,d]$ code with $d\le 3$. Then
$\rho_{\mbox{\tiny max-frac}}(\mathcal{C})$ is finite. Moreover, we have
$\rho_{\mbox{\tiny max-frac}}(\mathcal{C}) = n-k$ in the case $d \le 2$.
\end{proposition}
\begin{IEEEproof}
By using Lemma~\ref{lemma:distance_two} we may assume $d=3$. Denote
by $\bldH$ the parity-check matrix whose rows consist of \emph{all}
codewords of the dual code of $\mathcal{C}$. Note that for a code of minimum
distance $d$, a parity-check matrix $\bldH$ consisting of all rows of
the dual code $\mathcal{C}^{\perp}$ is an orthogonal array of strength
$d-1$. In the present case $d=3$, and this implies that in any pair
of columns of $\bldH$, all length-$2$ binary vectors occur with equal
multiplicities (cf.\ \cite[p. 139]{MacWilliams_Sloane}). Thus the
matrix $\bldH$ is an incidence matrix for a partial block design with
parameters $(w_c,\lambda) = (2^{r-1},2^{r-2})$, where $r=n-k$.
Therefore for this matrix $\bldH$ the code has minimum (AWGNC, BSC, or
max-fractional) pseudoweight at least $1 + w_c/\lambda = 3$, and it
follows that the pseudocodeword redundancy is finite for any code
with $d=3$.
\end{IEEEproof}
We remark that Proposition~\ref{prop:D3_codes} implies the results
for the Hamming codes (Proposition~\ref{prop:Hamming}). However, we
present the two proofs, since they use different methods.
We have considered in this section several families of codes based on
designs, which have finite pseudocodeword redundancy. As noted in
Section~\ref{sec:related}, finiteness of pseudoredundancy would also
follow if one can show that the codes are geometrically perfect.
However, this is not the case for the examined codes in general. For
example, the $[{2^m-1} \,, {2^m-1-m}\,, 3]$ Hamming code is not
geometrically perfect for $m\ge 4$; this follows from the
characterization of geometrically perfect codes, as the $[7,3,4]$
simplex code can be obtained from the Hamming code by repeated
shortening, when $m\ge 4$.
\section{The Pseudocodeword Redundancy for Codes of Small Length}%
\label{sec:short_codes}
In this section we compute the AWGNC, BSC, and max-fractional
pseudocodeword redundancies for all codes of small length. By
Lemma~\ref{lemma:distance_two} it is sufficient to examine only codes
with minimum distance at least~$3$. Furthermore, in light of
Lemma~\ref{lemma:puncturing} we will consider only codes without zero
coordinates, i.e., codes that have a dual minimum distance of at
least~$2$. Finally, we point out to
Corollary~\ref{corollary:dimension_two} for codes of dimension~$2$, by
which we may focus on codes with dimension at least~$3$.
\subsection{The Algorithm}
To compute the pseudocodeword redundancy of a code $\mathcal{C}$ we have to
examine all possible parity-check matrices for the code $\mathcal{C}$, up to
equivalence. Here, we say that two parity-check matrices $\bldH$ and
$\bldH'$ for the code $\mathcal{C}$ are \emph{equivalent} if $\bldH$ can be
transformed into $\bldH'$ by a sequence of row and column permutations.
In this case, $\mathsf{w}_{\min}(\bldH) = \mathsf{w}_{\min}(\bldH')$ holds for
the BEC, AWGNC, BSC, and max-fractional pseudoweights. The enumeration
of codes and parity-check matrices can be described by the following
algorithm.
\begin{algorithm}\hrulefill\smallskip
\textbf{Input:} Parameters $n$~(code length), $k$~(code dimension),
$\rho$~(number of rows of the output parity-check matrices), where
$\rho\ge r\stackrel{\mbox{\tiny $\triangle$}}{=} n-k$.
\textbf{Output:} For all codes of length $n$, dimension $k$, minimum
distance $d\ge 3$, and without zero coordinates, up to code
equivalence: a list of all $\rho\times n$ parity-check matrices, up to
parity-check matrix equivalence.
\begin{enumerate}
\item Collect the set $X$ of all $r\times n$ matrices such that
\begin{itemize}
\item they have different nonzero columns, ordered
lexicographically,
\item there is no non-empty $\mathbb{F}_2$-sum of rows which has weight~$0$
or~$1$ {\sl (this way, the matrices are of full rank and the
minimum distance of the row space is at least~$2$)}.
\end{itemize}
\item Determine the orbits in $X$ under the action of the group
$\operatorname{GL}_r(2)$ of invertible $r\times r$ matrices over $\mathbb{F}_2$ {\sl
(this enumerates all codes with the required properties, up to
equivalence; the codes are represented by parity-check matrices)}.
\item For each orbit $X_{\mathcal{C}}$, representing a code $\mathcal{C}$:
\begin{enumerate}
\item Determine the suborbits in $X_{\mathcal{C}}$ under the action of the
symmetric group $S_r$ {\sl (this enumerates all parity-check
matrices without redundant rows, up to equivalence)}.
\item For each representative $\bldH$ of the suborbits, collect all
matrices enlarged by adding $\rho-r$ different redundant rows that
are $\mathbb{F}_2$-sums of at least two rows of $\bldH$. Let
$X_{\mathcal{C},\rho}$ be the union of all such $\rho\times n$ matrices.
\item Determine the orbits in $X_{\mathcal{C},\rho}$ under the action of the
symmetric group $S_{\rho}$, and output a representative for each
orbit.
\end{enumerate}
\end{enumerate}
\noindent\hrulefill\smallskip
\end{algorithm}
This algorithm was implemented in the C programming language. The
minimum pseudoweights for the various parity-check matrices were
computed by using Maple 12 and the Convex package~\cite{maple-convex}.
\subsection{Results}\label{sec:short_codes_results}
\begin{table}
\caption{The Number of Binary $[n,k,d]$ Codes
with~$d\ge 3$~and~without~Zero~Coordinates}
\label{table:no-codes}
{\centering
\begin{tabular}{r|rcccc}
& $k=1$ & $2$ & $3$ & $4$ & $5$ \\\hline
& & & & & \vspace{-2mm} \\
$n=5$ & $1$ & $1$ & & & \\
$6$ & $1$ & $3$ & $1$ & & \\
$7$ & $1$ & $4$ & $4$ & $1$ & \\
$8$ & $1$ & $6$ & $10$ & $5$ & \\
$9$ & $1$ & $8$ & $23$ & $23$ & $5$
\end{tabular}\\}%
\end{table}
We considered all binary linear codes up to length~$n$ with minimum
distance $d\ge 3$ and without zero coordinates, up to code
equivalence. The number of those codes for given length~$n$ and
dimension~$k$ is shown in Table~\ref{table:no-codes}.
\subsubsection{AWGNC pseudoweight}
The following results were found to hold for all codes of length $n\le
9$.
\begin{itemize}\setlength{\parsep}{0pt}\setlength{\itemsep}{4pt}
\item There are only two codes $\mathcal{C}$ with $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})>r$,
i.e., in class $0$ or $1$ for the AWGNC.
\begin{itemize}
\item The $[8,4,4]$ extended Hamming code is the shortest code
$\mathcal{C}$ in class 1. We have $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})=5>4=r$ and out
of $12$ possible parity-check matrices (up to equivalence) with one
redundant row there is exactly one matrix $\bldH$ with
$\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH) = 4$, namely
\[ \bldH = \mat{
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
}\:. \] There is exactly one matrix $\bldH$ with
$\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH) = 25/7$, and for the remaining
matrices $\bldH$ we have $\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH) = 3$.
For this code, also $\rho_{\mbox{\tiny BEC}}(\mathcal{C})=5>4$, and it is the only
code of length $n\le 9$ with $\rho_{\mbox{\tiny BEC}}(\mathcal{C})>r$.
\item Out of the four $[9,4,4]$ codes there is one code $\mathcal{C}$ in
class 1. We have $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})=6>5=r$ and out of $2526$
possible parity-check matrices (up to equivalence) with one
redundant row there are $13$ matrices $\bldH$ with
$\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH) = 4$.
\end{itemize}
\item For all codes $\mathcal{C}$ of minimum distance $d\ge 3$ and for all
parity-check matrices $\bldH$ of $\mathcal{C}$ we have
$\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH)\ge 3$; in particular, if $d=3$,
then $\mathcal{C}$ is in class~$3$ for the AWGNC.
\item For the $[7,3,4]$ simplex code there is (up to equivalence) only
one parity-check matrix $\bldH$ without redundant rows such that
$\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH) = 4$, namely
\[ \bldH = \mat{
\mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} \\
\mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} & \mbox{\footnotesize \!$0$\!\! \normalsize} & \mbox{\footnotesize \!$\mathbf{1}$\!\! \normalsize} \\
}\:. \] It is the only parity-check matrix with constant row weight
$3$.
\end{itemize}
\subsubsection{BSC pseudoweight}
We computed the pseudocodeword redundancy for the BSC for all codes of
length $n\le 8$.
\begin{itemize}\setlength{\parsep}{0pt}\setlength{\itemsep}{4pt}
\item The shortest codes with $\rho_{\mbox{\tiny BSC}}(\mathcal{C})>r$, i.e., in class~$0$
or $1$ for the BSC, are the $[7,4,3]$ Hamming code $\mathcal{C}$ and its
dual code $\mathcal{C}^{\bot}$, the $[7,3,4]$ simplex code. We have
$\rho_{\mbox{\tiny BSC}}(\mathcal{C})=4>3$ and $\rho_{\mbox{\tiny BSC}}(\mathcal{C}^{\bot})=5>4$.
\item There are two codes of length~$8$ with
$\rho_{\mbox{\tiny BSC}}(\mathcal{C})>r$. These are the $[8,4,4]$ extended Hamming
code, for which $\rho_{\mbox{\tiny BSC}}(\mathcal{C})=6>4$ holds, and one of the three
$[8,3,4]$ codes, which satisfies $\rho_{\mbox{\tiny BSC}}(\mathcal{C})=6>5$.
\end{itemize}
\subsubsection{Max-fractional weight}
We computed the pseudocodeword redundancy with respect to the
max-fractional weight for all codes of length $n\le 8$.
\begin{itemize}\setlength{\parsep}{0pt}\setlength{\itemsep}{4pt}
\item The shortest code with $\rho_{\mbox{\tiny max-frac}}(\mathcal{C})>r$ is the unique
$[6,3,3]$ code $\mathcal{C}$. We have $\rho_{\mbox{\tiny max-frac}}(\mathcal{C})=4>3$.
\item There are two codes of length~$7$ with $\rho_{\mbox{\tiny max-frac}}(\mathcal{C})>r$.
These are the $[7,4,3]$ Hamming code and the $[7,3,4]$ simplex code,
which have both pseudocodeword redundancy~$7$. In both cases, there
is, up to equivalence, a unique parity-check matrix $\bldH$ with seven
rows that satisfies $d(\mathcal{C})=w_{\min}^{\mbox{\tiny max-frac}}(\bldH)$.
This demonstrates that Propositions~\ref{prop:Hamming}
and~\ref{prop:simplex} are sharp for the max-fractional weight, and
that the parity-check matrices constructed in the proofs are unique
in this case.
\item For the $[8,4,4]$ extended Hamming code $\mathcal{C}$ we have
$\rho_{\mbox{\tiny max-frac}}(C)=\infty$, and thus the code is in class~$0$ for
the max-fractional weight. It is the shortest code with infinite
max-fractional pseudoredundancy.
(It can be checked that $\boldsymbol{x}=[1,1,1,1,1,1,1,3]$ is a
pseudocodeword in $\mathcal{K}(\bldH)$, where the rows of $\bldH$ consist of all
dual codewords; since $\mathsf{w}_{\mbox{\tiny max-frac}}(\boldsymbol{x}) = \frac{10}{3}
< 4$, we have $\mathsf{w}_{\min}^{\mbox{\tiny max-frac}}(\bldH) < 4$.)
\item There are two other codes of length~$8$ with
$\rho_{\mbox{\tiny max-frac}}(\mathcal{C})>r$, namely two of the three $[8,3,4]$ codes,
having pseudocodeword redundancy~$6$ and~$8$, respectively.
\end{itemize}
\subsubsection{Comparison}
Comparing the results for the AWGNC and BSC pseudoweights, and the
max-fractional weight, we can summarize the results as follows.
\begin{itemize}\setlength{\parsep}{0pt}\setlength{\itemsep}{4pt}
\item For the $[7,4,3]$ Hamming code $\mathcal{C}$ we have
$\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})=r=3$\,, $\rho_{\mbox{\tiny BSC}}(\mathcal{C})=4$\,, and
$\rho_{\mbox{\tiny max-frac}}(\mathcal{C})=7$.
\item For the $[7,3,4]$ simplex code $\mathcal{C}$ we have
$\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})={r=4}$\,, $\rho_{\mbox{\tiny BSC}}(\mathcal{C})=5$\,, and
$\rho_{\mbox{\tiny max-frac}}(\mathcal{C})=7$.
\item For the $[8,4,4]$ extended Hamming code $\mathcal{C}$ we have
$\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})=5$\,, $\rho_{\mbox{\tiny BSC}}(\mathcal{C})=6$\,, and
$\rho_{\mbox{\tiny max-frac}}(\mathcal{C})=\infty$. This code $\mathcal{C}$ is the shortest
one such that $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})>r$, and also the shortest one
such that $\rho_{\mbox{\tiny max-frac}}(\mathcal{C})=\infty$.
\item If $d\ge 3$ then for \emph{every} parity-check matrix $\bldH$ we
have $\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH)\ge 3$. This is not true for
the BSC and the max-fractional weight.
\end{itemize}
These observations show that there is some significant difference
between the various types of pseudocodeword redundancies.
It is also interesting to note that the $[7,4,3]$ Hamming code is
geometrically perfect, while the $[7,3,4]$ code and the $[8,4,4]$ code
are not (cf.\ Section~\ref{sec:related}).
\section{Cyclic Codes Meeting the Eigenvalue Bound}%
\label{sec:kv_bound}
In this last section we apply the eigenvalue-based lower bound on the
minimum AWGNC pseudoweight by Vontobel and Koetter
\cite{KV-lower-bounds}, see Proposition~\ref{prop:KV_bound}. We
investigate for which cyclic codes of short length this bound is sharp
with respect to the minimum Hamming distance, for in this case, the
codes have finite AWGNC pseudoredundancy.
We consider binary cyclic codes with full circulant parity-check
matrices, defined as follows: Let $\mathcal{C}$ be a binary cyclic code of
length $n$ with \emph{check polynomial} $h(x)=\sum_{i\in\mathcal{I}}h_ix^i$
(cf.\ \cite{MacWilliams_Sloane}, p.~194). Then the \emph{full
circulant parity-check matrix} for~$\mathcal{C}$ is the $n\times n$ matrix
$\bldH=(H_{j,i})_{i,j\in \mathcal{I}}$ with entries $H_{j,i}=h_{j-i}$. Here,
all the indices are modulo $n$, so that $\mathcal{I}=\{0,1,\dots,n-1\}$.
Since such a matrix is $w$-regular, where $w=\sum_{i\in\mathcal{I}}h_i$, we
may use the eigenvalue-based lower bound of
Proposition~\ref{prop:KV_bound} to examine the AWGNC pseudocodeword
redundancy: If the right hand side equals the minimum distance $d$ of
the code~$\mathcal{C}$, then $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})\le n$.
Note that the largest eigenvalue of the matrix $\bldL = \bldH^T \bldH$ is
$\mu_1=w^2$, since every row weight of $\bldL$ equals
$\sum_{i,j\in\mathcal{I}}h_ih_j=w^2$. Consequently, the eigenvalue bound is
\[ \mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} \ge n \cdot \frac{2w-\mu_2}{w^2-\mu_2}
\:, \] where $\mu_2$ is the second largest eigenvalue of $\bldL$. We
remark further that $\bldL=(L_{j,i})_{i,j\in\mathcal{I}}$ is a symmetric
circulant matrix, with $L_{j,i}=\ell_{j-i}$ and
$\ell_i=\sum_{k\in\mathcal{I}}h_kh_{k+i}$. The eigenvalues of $\bldL$ are thus
given by \[ \lambda_j = \sum_{i\in\mathcal{I}}\ell_i\zeta_n^{ij} =
\operatorname{Re}\sum_{i\in\mathcal{I}}\ell_i\zeta_n^{ij} =
\sum_{i\in\mathcal{I}}\ell_i\cos(2\pi ij/n) \] for $j\in\mathcal{I}$, where
$\zeta_n=\exp(2\pi{\bf i}/n)$ is the $n$-th primitive root of unity
and ${\bf i}^2=-1$ (see, e.g., \cite{Davis-book}, Theorem~3.2.2).
We also consider quasi-cyclic codes of the form given in the following
remark. This code construction is only introduced for completeness
towards classifying the results; the resulting codes are not
interesting for applications, as the minimum Hamming distance is at
most~$2$ for $m\ge 2$.
\begin{remark}\label{rem:qc}
Denote by $\boldsymbol{1}_m$ the $m\times m$ matrix with all entries equal to
$1$. If $\bldH$ is a $w$-regular circulant $n\times n$ matrix then
the Kronecker product $\tilde{\bldH} \stackrel{\mbox{\tiny $\triangle$}}{=} \bldH\otimes\boldsymbol{1}_m$ will
be a $w$-regular circulant $mn\times mn$-matrix and defines a
quasi-cyclic code. We have
\[ \tilde{\bldL} = \tilde{\bldH}^T\tilde{\bldH} = \bldH^T\bldH \otimes
\boldsymbol{1}_m^T\boldsymbol{1}_m = \bldL \otimes (m\boldsymbol{1}_m)\:,\] and the eigenvalues
of $m\boldsymbol{1}_m$ are $m^2$ and $0$. Thus, the largest eigenvalues of
$\tilde{\bldL}$ are $\tilde{\mu}_1 = m^2\mu_1 = m^2w^2$ and
$\tilde{\mu}_2 = m^2\mu_2$, and the eigenvalue bound of
Proposition~\ref{prop:KV_bound} becomes
\[ \mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} \ge mn \cdot
\frac{2mw-m^2\mu_2}{m^2w^2-m^2\mu_2} =
n \cdot \frac{2w-m\mu_2}{w^2-\mu_2} \:. \]
\end{remark}
We carried out an exhaustive search on all cyclic codes $\mathcal{C}$ up to
length $n\le 250$ and computed the eigenvalue bound in all
cases where the Tanner graph of the full circulant parity-check matrix
is connected, by using the following algorithm.\smallskip
\begin{algorithm}
\hrulefill\smallskip
\textbf{Input:} Parameter $n$ (code length).
\textbf{Output:} For all divisors of $x^n-1$, corresponding to cyclic
codes $\mathcal{C}$ with full circulant parity-check matrix, such that the
Tanner graph is connected: the value of the eigenvalue bound.\medskip
\begin{enumerate}
\item Factor $x^n-1$ over~$\mathbb{F}_2$ into irreducibles, using Cantor and
Zassenhaus' algorithm (cf.\ \cite{vzGG-book}, Section~14.3).
\item For each divisor $f(x)$ of $x^n-1$:
\begin{enumerate}
\item Let $f(x)=\sum_ih_ix^i$ and $\bldH=(h_{j-i})_{i,j\in\mathcal{I}}$.
\item Check that the corresponding Tanner graph is connected (i.e.,
that the greatest common divisor of the indices $i$ with $h_i=1$
together with $n$ is~$1$).
\item Compute the eigenvalues of~$\bldL=\bldH^T\bldH$: Let
$\ell_i=\sum_{k\in\mathcal{I}}h_kh_{k+i}$ and for $j\in\mathcal{I}$ compute
$\sum_i\ell_i\cos(2\pi ij/n)$.
\item Determine the second largest eigenvalue $\mu_2$ and output
$n\cdot(2\ell_0-\mu_2)/(\ell_0^2-\mu_2)$.
\end{enumerate}
\end{enumerate}
\noindent\hrulefill\smallskip
\end{algorithm}
This algorithm was implemented in the C programming language.
Tables~\ref{table:KV_bound-2} and~\ref{table:KV_bound-1} give a
complete list of all cases in which the eigenvalue bound equals the
minimum Hamming distance $d$, for the cases $d=2$ and $d\ge 3$,
respectively. In particular, the AWGNC pseudoweight equals the
minimum Hamming distance in these cases and thus we have for the
pseudocodeword redundancy $\rho_{\mbox{\tiny AWGNC}}(\mathcal{C})\le n$. All examples of
minimum distance $2$ are actually quasi-cyclic codes as in
Remark~\ref{rem:qc} with parity-check matrix $\tilde{\bldH} =
\bldH\otimes\boldsymbol{1}_2$. We list here the constituent code given by the
parity-check matrix~$\bldH$.
\begin{table}
\caption{Binary Cyclic Codes up to Length 250
with $d=2$ Meeting~the~Eigenvalue~Bound}\label{table:KV_bound-2}
{\centering
\begin{tabular}{ccl}
parameters & $w$-regular & constituent code\vspace{.5mm} \\\hline
& & \vspace{-2mm} \\
$[2n,2n\!-\!m,2]$ & $2^m$ & Hamming c., $n=2^m\!-\!1$, $m=2\dots 6$ \\
\!\!\!$[2n,2n\!-\!m\!-\!1,2]$\!\!\! &
\!\!\!$2^m\!-\!2$\!\!\! & Hamming c. with overall
parity-check \\
$[42,32,2]$ & $10$ & projective geometry code $PG(2,4)$ \\
$[146,118,2]$ & $18$ & projective geometry code $PG(2,8)$ \\
$[170,153,2]$ & $42$ & a certain $[85,68,6]$ $21$-regular code \\
& & (the eigenvalue bound is 5.2) \\
\end{tabular}\\}%
\end{table}
\begin{table}
\caption{Binary Cyclic Codes up to Length 250
with $d\ge 3$ Meeting~the~Eigenvalue~Bound}\label{table:KV_bound-1}
{\centering
\begin{tabular}{ccl}
parameters & $w$-regular & comments\vspace{.5mm} \\\hline
& & \vspace{-2mm} \\
$[n,1,n]$ & $2$ & repetition code, $n=3\dots 250$ \\
$[n,n\!-\!m,3]$ & \!\!$2^{m-1}$\!\!
& Hamming c., $n=2^m\!-\!1$, $m=3\dots 7$ \\
$[7,3,4]$ & $3$ & dual of the $[7,4,3]$ Hamming code \\
$[15,7,5]$ & $4$ & Euclidean geometry code EG(2,4) \\
$[21,11,6]$ & $5$ & projective geometry code PG(2,4) \\
$[63,37,9]$ & $8$ & Euclidean geometry code EG(2,8) \\
$[73,45,10]$ & $9$ & projective geometry code PG(2,8) \\
\end{tabular}\\}%
\end{table}
We conclude this section by proving a result which was observed by the
experiments.
\begin{lemma}
Let $m\ge 3$ and let $\mathcal{C}$ be the intersection of a Hamming code of
length $n=2^m-1$ with a simple parity-check code of length $n$,
which is a cyclic $[n\,, {n-m-1}\,, 4]$ code. Consider its full circulant
parity-check matrix $\bldH$. Then \[ \mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH)\ge
3+\frac{1}{2^{m-2}-1}>3 \:. \]
In particular, if $m=3$ then $\mathcal{C}$ is the $[7,3,4]$ code and the
result implies $\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}}(\bldH)=4$ and $\rho_{\mbox{\tiny AWGNC}}
(\mathcal{C}) \le 7$.
\end{lemma}
\addtolength{\textheight}{-8cm}
\begin{IEEEproof}
Let $\bldH$ be the $w$-regular full circulant parity-check matrix for
$\mathcal{C}$. We claim that $w=2^{m-1}\!-\!1$. Indeed, each row $\boldsymbol{h}$
of $\bldH$ is a codeword of the dual code $\mathcal{C}^{\bot}$, and since
$\mathcal{C}^{\bot}$ consists of the codewords of the simplex code and their
complements, the weight of $\boldsymbol{h}$ and thus $w$ must be
$2^{m-1}\!-\!1$, $2^{m-1}$, or $2^m\!-\!1$. But $w$ cannot be even,
for otherwise all codewords of $\mathcal{C}^{\bot}$ would be of even weight.
As $w=2^m\!-\!1$ is clearly impossible, it must hold
$w=2^{m-1}\!-\!1$.
Next, we show that the second largest eigenvalue of
$\bldL=\bldH^T\bldH=(L_{j,i})_{i,j\in\mathcal{I}}$ equals $\mu_2=2^{m-2}$.
Indeed, let $\boldsymbol{h}_1$ and $\boldsymbol{h}_2$ be different rows of $\bldH$,
representing codewords of $\mathcal{C}^{\bot}$. As their weight is equal,
their Hamming distance is even, and thus it must be $2^{m-1}$.
Hence, the size of the intersection of the supports of $\boldsymbol{h}_1$ and
$\boldsymbol{h}_2$ is $2^{m-2}\!-\!1$. This implies that $L_{i,i}=w$ and
$L_{j,i}=2^{m-2}\!-\!1$, for $i\ne j$. Consequently, $\bldL$ has an
eigenvalue of multiplicity $n-1$, namely
$w-(2^{m-2}\!-\!1)=2^{m-2}$, and thus $\mu_2$ must be $2^{m-2}$.
Finally, we apply Proposition~\ref{prop:KV_bound} to get
\begin{align*}
\mathsf{w}_{\min}^{\mbox{\tiny AWGNC}} &\ge (2^m\!-\!1)\,
\frac{2\,(2^{m-1}\!-\!1)-2^{m-2}}{(2^{m-1}\!-\!1)^2-2^{m-2}}\\
&= \frac{(2^m\!-\!1)\,\big(2\,(2^{m-1}\!-\!1)-2^{m-2})}{(2^m\!-\!1\big)\,
(2^{m-2}\!-\!1)}\\
&= \frac{3\,(2^{m-2}\!-\!1)+1}{2^{m-2}-1} \:,
\end{align*}
which proves the result.
\end{IEEEproof}
\section*{Acknowledgments}
The authors are indebted to the anonymous reviewers for their
worthwhile suggestions and recommendations, which significantly
improved the exposition of the work, and also to Pascal Vontobel for
carefully handling the manuscript. In addition, they would like to
thank Nigel Boston, Christine Kelley, and Olgica Milenkovic for helpful
discussions.
|
\section{Introduction and basic notions}
The theory of functions with values in a Clifford algebra that are annihilated by the Dirac operator represents one possibility to generalize classical complex function theory to higher dimensions and can be regarded as a counterpart to function theory of several complex variables. In the classical case dealing with functions in one complex variable and in the context of several complex variables theory, the study of the properties of the Bergman metric and of the related Szeg\"o metric has been a subject of interest over the past decades. These metrics turned out to be a useful tool in differential geometry, too. More concretely, using special properties of the Szeg\"o metric it was possible to establish a close connection between the completeness of the Bergman metric and the smoothness of the boundary. See for instance works of Diederich, Forn\ae{}ss, Herbort, Kobayashi, Krantz, Pflug, and others ( \cite{Died70,Koba59,Koba62,JPZ00,BP98,Herb99,DFH84,Kran01}) in the context of function theory of several complex variables. It is of great interest to investigate on which domains the Szeg\"o metric is complete. Recent years publications like \cite{Herb99,JPZ00,KK03} show that there is still both interest and need for research in the field of several complex variables, also in the study basic properties of these metrics (\cite{ForLee2009}). Publications for example by Chen \cite{Chen04,Chen03,Chen99}, Nikolov \cite{Niko03} and Herbort \cite{Herb99,Herb04} examined the Bergman completeness on different kinds of manifolds and domains with varying smoothness of the boundary, while Blumberg, Nikolov, Pflug and others studied estimates and behaviour of both the Bergman kernel and the Bergman metric on different types of domains, see for example \cite{Blum05,NP03,NP03_2}.
\par\medskip\par
The strong results in the context of one and several complex variables theory provide a motivation to investigate whether one can establish similar results in the context of Clifford algebra valued functions. In order to start we first need to briefly recall some basic terms and notions.
\par\medskip\par
Let $\{e_1,e_2,\ldots,e_m\}$ be the standard basis of the
Euclidean vector space ${\mathbb{R}}^m$. Further, let ${\mathcal{C}l}_{0m}$
be the associated real Clifford algebra in which
$$e_i e_j + e_j e_i = - 2 \delta_{ij} e_0,\quad i,j=1,\cdots,m,$$
holds. Here $\delta_{ij}$ is Kronecker symbol. Each
element $a \in {\mathcal{C}l}_{0m}$ can be represented in the form $ a =
\sum_A a_A e_A$ with $a_A \in {\mathbb{R}}$, $A \subseteq
\{1,\cdots,m\}$, $e_A = e_{l_1} e_{l_2} \cdots e_{l_r}$, where $1
\le l_1 < \cdots < l_r \le m,\;\; e_{\emptyset} = e_0 = 1$. The
scalar part of $a$, denoted by $\mathop{\rm Sc}\nolimits(a)$, is defined as the $a_0$
term. The Clifford conjugate of $a$ is defined by $\overline{a} =
\sum_A a_A \overline{e}_A$, where $\overline{e}_A =
\overline{e}_{l_r} \overline{e}_{l_{r-1}} \cdots
\overline{e}_{l_1}$ and $\overline{e}_j = - e_j$ for
$j=1,\cdots,m,\; \overline{e}_0 = e_0 = 1$.
For the space of paravectors ${\mathbb{R}} \oplus {\mathbb{R}}^n \subset {\mathcal{C}l}_{0m}$, which consists of elements of the form $z = z_0 + z_1 e_1 + z_2 e_2 + \cdots +
z_m e_m$, we also write $\mathbb{R}^{m+1}$ for simplicity. For such an element $z$, the term $\sum_{i=1}^m z_i e_i$ is called the vector part of $z$ and is denoted by $\overrightarrow{z}$ or $Vec(z)$. The standard scalar product
between two Clifford numbers $a,b \in {\mathcal{C}l}_{0m}$ is defined by
$\langle a,b \rangle := \mathop{\rm Sc}\nolimits(a\overline{b})$. This induces a pseudo
norm on the Clifford algebra, viz $\|a\| = ( \sum\limits_A
|a_A|^2)^{1/2}$.
\par\medskip\par
In the Euclidean flat space $\mathbb{R}^{m+1}$, the associated Dirac operator has the simple form
\begin{equation}
\label{cr} D_z = \frac{\partial }{\partial z_0} +
\sum\limits_{j=1}^m \frac{\partial }{\partial z_j} e_j \quad \quad
z_j \in {\mathbb{R}} \;\;j=0,\ldots,m.
\end{equation}
In this particular context it is often called the generalized Cauchy-Riemann operator.
\par\medskip\par
Next let $G \subseteq {\mathbb{R}}^{m+1}$ be an open set. A real
differentiable function $f: G \rightarrow \mathcal{C}l_{0m}$ that
satisfies inside of $G$ the system $D_z f = 0$ (or $ f D_z
=0$) is called left (right) monogenic with respect
to the paravector variable $z$, respectively. In the
two-dimensional case (i.e. $m=1$) this differential operator coincides with the
classical complex Cauchy-Riemann operator. In this sense, the set
of monogenic functions can also be regarded as a higher dimensional
generalization of the set of complex-analytic functions. For
details, see for example \cite{DSS} and elsewhere. The first order
operator $D_z$ factorizes the Euclidean Laplacian, viz
$D_z\overline D_z=\Delta_z$.
\par\medskip\par
An important example of a separable Hilbert space of Clifford
valued monogenic functions that satisfies in particular the
Bergman condition $\|f(z)\|\leq C(z)\|f\|_{L^2}$ is the closure of
the set of functions that are monogenic inside a domain of
${\mathbb{R}}^{m+1}$ and square integrable on the boundary. This
space is called the Hardy space of monogenic functions. Each Hardy
space of monogenic Clifford algebra valued functions is a Banach
space which is endowed with the Clifford-valued inner product $
(f,g) = \int_{\partial G} \overline{f(z)} g(z) dS$, where $f$ and
$g$ denote
elements of the Hardy space. Basic contributions to the study of these function spaces have already been conducted in the 1970s, see for example \cite{BD78}.\\
The reproducing kernel $K_G$ of the Hardy space of monogenic functions, called the Szeg\"o
kernel, is uniquely defined and satisfies $f(z) = \int_{\partial G}
{K_{G}(z,w)} f(w) dV $ for any function $f$ that belongs to that Hardy space.
The function $K_G(z,z)$ represents a non-negative real valued function (see, for example, \cite{BDS,GHS} or below). This in turn can be used to define a Hermitian metric on the domain $G$. This metric then is called the Szeg\"o metric.
\par\medskip\par
A reason why we focus on this paper on the study of the Szeg\"o metric and not on a generalization of the Bergman metric is that we are interested in metrics which are invariant under conformal transformations. Although we do not get a genuine total invariance, the Szeg\"o metric turns out to exhibit an invariance up to a simple scaling factor. This is a consequence of the well-known transformation formula between two Hardy spaces of monogenic functions, see for instance \cite{Cno02,Cno94}. This transformation formula expresses the isometry of two Hardy spaces of monogenic functions in the case that both domains can be transformed into each other by a conformal map. As a consequence of Liouville's theorem, this is exactly the case when both domains can be transformed into each other by a M\"obius transformation. A M\"obius transformation is generated by reflections at spheres and hyperplanes. Following classical works like for instance,\cite{Ahlf86}, we recall that a M\"obius transformation in $\mathbb{R}^{m+1}$ can be expressed in the simple form $T(z) = (az+b)(cz+d)^{-1}$ when $a,b,c,d$ are Clifford numbers that satisfy the constraints
\begin{enumerate}
\item[(i)] $a,b,c,d$ are products of paravectors from
$\mathbb{R}^{m+1}$
\item[(ii)] \ $\tilde{H} := a\tilde{d} - b\tilde{c} = 1 $
\item[(iii)] \ $ac^{-1}, \, c^{-1}d \in \mathbb{R}^{m+1} $ \, for \, $ c \neq 0 \, $ and \, $ bd^{-1} \in \mathbb{R}^{m+1} $ \, for \,$ c=0 $ .
\end{enumerate}
\par\medskip\par
Recently some attempts have been made to introduce a Bergman
metric in the context of hypercomplex function theory, see e.g.
\cite{Cerv09}. In the very interesting work \cite{Cerv09} the
author studied generalizations of the famous Bergman
transformation formula. However, the transformation formula
presented in \cite{Cerv09} does not represent an isometry between
the two Bergman spaces. We do not have an isometry as a
consequence of the appearance of the extra factor
$\frac{1}{|z|^2}$. That factor is only equal to $1$ on the
boundary of the unit ball and not in the interior of it. Due to
the non-existence of an isometry of two Bergman spaces of two
conformally equivalent domains it seems very unlikely that one can
establish a total analogy of the classical transformation formula
for the Bergman kernel when dealing with Clifford algebras that
are different from $\mathbb{C}$. This is one motivation why we look in
this paper at the hypercomplex Szeg\"o kernel instead.
\par\medskip\par
The existence of an isometric transformation formula for the hypercomplex Szeg\"o kernel is a fundamental advantage.
In Section~2 of this paper we depart from that transformation formula in order to establish the invariance of the hypercomplex Szeg\"o metric (up to a scaling factor) under conformal transformations. This invariance in turn is crucially applied in Section~3 to show that the curvature of this metric is negative on bounded domains. Finally, in Section~4, this property again is used to establish the completeness of this metric on bounded domains. These results can be regarded as a close analogy to the results from one and several complex variables theory.
\section{Simple consequence of the transformation formula}
The starting point of our consideration is the following transformation formula of the Szeg\"o kernel, cf. for example \cite{Cno94,Krau98,Cno02}:
\begin{theorem}
Let $G, G^* \, \subseteq \mathbb{R}^{m+1}$ be two bounded domains with smooth boundaries. Let $T: G \longrightarrow G^*$ be a M\"obius transformation with coefficients $a,b,c,d$, which maps $G$ conformally onto $G^*$. By $K_{G}$ and $K_{G^*}$ we denote the Szeg\"o kernels belonging to $H^2(\partial G,\mathcal{C}l_{0m})$ resp. $H^2(\partial G^*,\mathcal{C}l_{0m})$.
We further write $z^*:=T(z)$ for $z \in G$. Then we have the following transformation formula:
\[
K_{G}(z,\zeta)= \frac{\overline{cz+d}}{\vert cz+d\vert^{m+1}} K_{G^*}(z^*,\zeta^*) \frac{c\zeta + d}{\vert c\zeta+d \vert^{m+1}}.
\]
\end{theorem}
To proceed we follow the theory of \cite{Kran04} and introduce in the same spirit
\begin{definition}
\begin{enumerate}
\item A Hermitian metric on a domain $G$ is a continuous function $\lambda$ on $G$, which is positive up to isolated zeros.
\item The length of a piecewise continuously differentiable path $C$ in $G$ with parameterisation $\gamma : [a,b] \to G$ with respect to a Hermitian metric $\lambda$ is defined by
\[
L_{\lambda}(C) = \int_{C} \lambda(z) \vert dz \vert.
\]
From now on, we denote Hermitian metrics by
\[
ds = \lambda(z) |dz|
\]
and call $ds$ the metric's \textit{line element}.
\end{enumerate}
\end{definition}
\begin{definition}
The distance function $d: G \times G \to \mathbb{R}$ that is associated to a Hermitian metric $ds$ is defined by
\[
d(z_1, z_2) = \inf \left\lbrace L_{ds}(C) : C \text{ path in G from } z_1 \text{ to } z_2 \right\rbrace.
\]
\end{definition}
The proof that $d$ is a distance function is analogue to the complex case, which is described in \cite{FL}.
\begin{definition}
A Hermitian metric $ds = \lambda(z) |dz|$ is called \textit{regular}, if $\lambda$ is at least twice continuously differentialble and does not have any zeros.
\end{definition}
Finally, one introduces the Gau{\ss}ian curvature as usual by
\begin{definition}
Let $ds = \lambda(z) |dz|$ be a regular metric. Then the function
\[
\mathfrak{G}_{ds}(z):=-\frac{1}{\lambda(z)^2} \Delta \log \lambda(z)
\]
is called the (Gau\ss ian) curvature of the metric $ds$.
\end{definition}
This naturally leads to the following definition of the Szeg\"o metric in the context of Hardy spaces of monogenic functions:
\begin{definition}
Let $G \subset \mathbb{R}^{m+1}$ be a domain and $K_G$ the Szeg\"o kernel of this domain. The Hermitian metric, which is defined by
\[
ds_G := K_G (z,z) \vert dz \vert,
\]
is called the \it{Szeg\"o metric}.
\end{definition}
To establish that $ds_G$ actually is a metric we only have to
verify that $K_G (z,z)$ is a positive real valued expression. Since $K_G$ is the reproducing kernel of the Hardy space, we can conclude
\[
K_G(z,z) = (K_G(z, \cdot), K_G(z, \cdot)) = \Vert K(z, \cdot) \Vert,
\]
where $\Vert \cdot \Vert$ is the usual $L_2$-norm induced by the inner product on $L^2(G)$. $K_G(z,\cdot)$ cannot be the zero function, because otherwise the Hardy space would be degenerated. Therefore, $K_G(z,z)$ is real valued and positive for any $z$. Thus, $ds_G$ is indeed a well-defined metric.
\par\medskip\par
Our starting point is to prove.
\begin{theorem}
The Szeg\"o metric is invariant under M\"obius transformations up to the automorphy factor
\[
\vert cz + d \vert^{2m-2}.
\]
\end{theorem}
\begin{proof}
Let $T$ be a M\"obius transformation, expressed through $\left( \begin{array}{lr} a & b \\ c & d \end{array} \right) $ with Clifford numbers $a, b, c, d$ satisfying the usual constraints as mentioned above.\\
Then
\begin{eqnarray*}
ds_{T(G)} (z) & = & K_{T(G)}(T(z),T(z)) \vert dT(z) \vert \\
& = & K_{T(G)}(T(z),T(z)) \frac{1}{\vert cz + d \vert^{2}} \vert dz \vert \\
& = & \vert cz + d \vert^{2m-2} \underbrace{K_{T(G)}(T(z),T(z))}_{\in \mathbb{R}} \frac{\overline{cz + d}}{\vert cz + d \vert^{m+1}} \frac{cz + d}{\vert cz + d \vert^{m+1}} \vert dz \vert\\
& = & \vert cz + d \vert^{2m-2} \frac{\overline{cz + d}}{\vert cz + d \vert^{m+1}} K_{T(G)}(T(z),T(z)) \frac{cz + d}{\vert cz + d \vert^{m+1}} \vert dz \vert \\
& = & \vert cz + d \vert^{2m-2} K_G(z,z) \vert dz \vert,
\end{eqnarray*}
where we apply the transformation formula for the Szeg\"o kernel
in the last line.
\end{proof}
\begin{remark}
In the complex case, i.e. $m=1$, the automorphic factor simplifies to $1$ and the metric thus is completely invariant under M\"obius transformations.
\end{remark}
For the sake of readability and better comparability to the complex case, we introduce the following notation:\\
For a continuously differentiable function $f$ we define
\[
f' := \bar{D} f .
\]
Here, an in all that follows, $D$ stands for $D_z :=\frac{\partial }{\partial x_0} + \sum\limits_{i=1}^m e_i \frac{\partial }{\partial x_i}$. We will leave out the index when the variable of differentiation is unambiguous.\\
Next we need the following generalized product role:
\begin{theorem}\label{produktregel}
Let $f$ and $g \in C^1(\Omega, \mathcal{C}l_{0,m})$. Then
\[
D(fg) = (Df)g + \bar{f}(Dg) + 2R(f)(\partial_0 g) + 2 \sum_{i=1}^m \sum_{ \begin{array}{c} \scriptstyle A \subseteq \{1,\ldots, m\}, \\ \scriptstyle i \notin A : |A| \equiv_4 0, 1 \\ \scriptstyle i \in A : |A| \equiv_4 2, 3 \end{array} } f_A (\partial_i g)
\]
and
\[
(fg)D = (fD)\bar{g} + f(gD) + 2(f \partial_0)R(g) + 2 \sum_{i=1}^m \sum_{\begin{array}{c} \scriptstyle A \subseteq \{1,\ldots, m\}, \\ \scriptstyle i \notin A : |A| \equiv_4 0, 1 \\ \scriptstyle i \in A : |A| \equiv_4 2, 3 \end{array}} g_A (\partial_i f),
\]
as well as
\[
\overline{D}(fg) = (\overline{D}f)g + \bar{f}(\overline{D}g) + 2R(f)(\partial_0 g) - 2\sum_{i=1}^m \sum_{\begin{array}{c} \scriptstyle A \subseteq \{1,\ldots, m\}, \\ \scriptstyle i \notin A : |A| \equiv_4 0, 1 \\ \scriptstyle i \in A : |A| \equiv_4 2, 3 \end{array}} f_A (\partial_i g)
\]
and
\[
(fg)\overline{D} = (f\overline{D})\bar{g} + f(g\overline{D}) + 2(f \partial_0)R(g) - 2 \sum_{i=1}^m \sum_{\begin{array}{c} \scriptstyle A \subseteq \{1,\ldots, m\}, \\ \scriptstyle i \notin A : |A| \equiv_4 0, 1 \\ \scriptstyle i \in A : |A| \equiv_4 2, 3 \end{array}} g_A (\partial_i f),
\]
where we use the notation $R(z):=z - \mathop{\rm Sc}\nolimits(z)$ for a Clifford number $z$.
\end{theorem}
\begin{proof}
The proof follows almost along the same lines as the proof for the case of quaternionic vectors, which can be found in \cite{GS90}. Therefore, we will only give the additional, necessary arguments for the proof of the first equation (the other three cases can then be treated analogously).\\
By applying simple calculations, we obtain that
\begin{eqnarray*}
\partial_0 (fg) &=& (\partial_0 f)g + f(\partial_0 g) \\
&=& (\partial_0 f)g + \overline{f}(\partial_0 g) + 2\;R(f)(\partial_0 g).
\end{eqnarray*}
Furthermore, notice that
\begin{equation*}
e_i e_A = \left\lbrace \begin{array}{l}
(-1)^{|A|} e_A e_i \; \text{for} \; i \notin A \\
(-1)^{|A|-1} e_A e_i \; \text{for} \; i \in A \\
\end{array} \right.
\end{equation*}
holds for any $i \in \{1,\ldots , m \}$ and any $A \subseteq \{1,\ldots , m \}$.\\
In view of
\begin{equation*}
\overline{e_A} = \left\lbrace \begin{array}{l}
e_A \; \text{for} \; |A| \equiv_4 0, 3 \\
-e_A \; \text{for} \; |A| \equiv_4 1, 2 \\
\end{array} \right. ,
\end{equation*}
we obtain $e_i e_A = \overline{e_A} e_i$ for all choices of $A$ except for $i \notin A , |A| \equiv_4 0, 1$ and $i \in A , |A| \equiv_4 2, 3$. For details concerning the calculations with imaginary units see also \cite{BDS}.
\end{proof}
From now on, we denote by
\[
\partial^z_i, \qquad i \in \{0, \ldots, m \}
\]
the real partial derivative with respect to the $i$-th component and the variable $z$.
\section{The curvature of the Szeg\"o metric}
The main aim of this section is to show that the Szeg\"o metric is
negative on bounded domains. This provides us with a generalization of the well-known analogous statement in classical complex analysis, as presented for instance in \cite{FL}. This result will then further be applied in the following section where we prove that the Szeg\"o metric is complete.
To proceed in this direction we first need to prove the following
\begin{lemma}\label{drei}
Let $G$ be a domain and $\left( \phi_k(z) \right)_{k \in \mathbb{N}}$ be an orthonormal basis of the Hardy space over $G$. Then
\[
\sum_{k=1}^{\infty} \vert \phi'_k(z) \vert^2 < \infty .
\]
\end{lemma}
\begin{proof}
Let $K=K_G$ be the Szeg\"o kernel of $G$. Due to the monogenicity resp. anti-monogenicity in the components and the compact convergence of the series expansion, we can infer that
\begin{equation}\label{eins}
\overline{D} K(z,w) = \sum_{k=1}^{\infty} \phi'_k(z) \overline{\phi_k(w)} < \infty ,
\end{equation}
and the series is again compactly convergent, as a consequence of Weierstra\ss ' convergence theorem .\\
For an arbitrary $k \in \mathbb{N}$ we have
\begin{eqnarray}\label{zwei}
( \overline{D_z} + \overline{D_w} ) \phi_k (z) \phi_k (w) (D_z + D_w) &=& \phi'_k (z) \overline{\phi_k (w)}D_z + \overline{D_w} \phi_k (z) \overline{\phi_k (w)} D_z \\
&& + \overline{D_w} \phi_k (z) \overline{\phi'_k (w)} + \phi'_k (z) \phi'_k (w) ,
\end{eqnarray}
as well as (in view of Theorem \ref{produktregel})
\begin{eqnarray*}
\phi'_k (z) \overline{\phi_k (w)}D_z &=& (\phi'_k(z)D_z) \phi_k (w) + \phi'_k (z) (\underbrace{\overline{\phi_k (w)} D_z}_{=0} ) \\
&& + 2 ( \partial^z_0 \phi'_k (z) \overline{\phi_k (w)} ) + 2 \sum_{i=1}^m \sum_{\begin{array}{c} \scriptstyle A \subseteq \{1,\ldots, m\}, \\ \scriptstyle i \notin A : |A| \equiv_4 0, 1 \\ \scriptstyle i \in A : |A| \equiv_4 2, 3 \end{array}} ((\phi'_k (z)) \partial^z_i) (\overline{\phi_k (w)} )_A \\
\bar{D_w} \phi_k (z) \overline{\phi'_k (w)} &=& \underbrace{(\bar{D}_w \phi_k (z))}_{=0} + \overline{ \phi_k (z) } \underbrace{(\bar{D}_w) \overline{ \phi'_k (z) } }_{= \overline{\phi'_k (w)D_w}=0} \\
&& + 2 Vec(\phi_k (z)) (\partial_0^w \overline{\phi'_k (w)}) + 2 \sum_{i=1}^m \sum_{A} (\phi_k (z))_A \partial_i^w \overline{ \phi'_k (w) }\\
\bar{D}_w \phi_k (z) \overline{\phi_k (w)} D_z &=& \left[ (\bar{D}_w \phi_k (z)) \overline{\phi_k (w)} + \overline{\phi_k (z)} \overline{\phi_k (w) D_w} \right. \\
&& \left. + 2 Vec(\phi_k (z)) (\partial^w_0 \overline{\phi_k (w))} + 2 \sum_{i=1}^m \sum_{A} (\phi_k (z))_A \partial_i^w \overline{\phi_k (w)} \right] D_z\\
&=& \left[ + 2 Vec(\phi_k (z)) (\partial^w_0 \overline{\phi_k (w))} + 2 \sum_{i=1}^m \sum_{A} (\phi_k (z))_A \partial_i^w \overline{\phi_k (w)} \right] D_z,\\
\end{eqnarray*}
where each sum is taken over the same set of $A$s.\\
Because of
\[
\lim_{w \to z } \sum_{k=1}^{\infty} ( \overline{D_z} + \overline{D_w} ) \phi_k (z) \phi_k (w) (D_z + D_w) = \overline{D_z} K(z,z) D_z < 0
\]
and in view of (\ref{zwei}) it suffices to show that the sums in the expanded terms of
\begin{eqnarray*}
(\phi'_k(z)D_z) \phi_k (w)\\
\phi'_k (z) \overline{\phi_k (w)}D_z, \\
\overline{D_w} \phi'_k (w) \overline{\phi'_k (w)}, \\
\overline{D}_w \phi_k (z) \overline{\phi_k (w)} D_z, \\
\end{eqnarray*}
converge in each case.
Both the series expansion of $K$ and of (\ref{eins}) are convergent and well-defined. As $K$ is the Szeg\"o kernel, $K$ is infinitely many times continuously real differentiable, as well as (\ref{eins}). Thus, we are able to differentiate partially with respect to any component in any order as many times as we want or can apply the Cauchy-Riemann operator on $K$ and (\ref{eins}). As both series converge compactly, we can switch these actions with the summation due to Weierstra\ss ' Theorem. So the summation of the terms above yields to a convergent series in each case.
\end{proof}
From now on, we will leave out the index and just write $K$ for the Szeg\"o kernel, if it is clear which is the corresponding domain of the kernel.\\
We introduce further notation
\[
K_z := \overline{D}_z K(z, \cdot)
\]
\[
K_{\bar{z}} := K(z, \cdot) D_z
\]
\begin{lemma}\label{vier}
Let $G$ be a bounded domain and $K$ be the Szeg\"o kernel of $G$. Then
\[
K(z,z)( K(z,z) K_{\bar{z}z}(z,z) - K_z(z,z) K_{\bar{z}}(z,z)) > 0 .
\]
\end{lemma}
\begin{beweis}
First, we show that $K_z(z,\cdot)$ is again an element of the Hardy space.\\
For that matter:\\
As $K$ is monogenic in $z$, so is $K_z$. Furthermore, we have in
view of (\ref{eins}):
\begin{eqnarray*}
\int_{G} \vert K_z (z, w) \vert^2 dV_w &=& \int_{G} \vert \sum_{k=1}^{\infty} \phi'_k(z) \overline{\phi_k(w)} \vert^2 dV_w \\
&\leq& \int_{G} \sum_{k=1}^{\infty} \vert \phi'_k(z) \overline{\phi_k(w)} \vert^2 dV_w \\
&\leq & \int_{G} \left( \sum_{k=1}^{\infty} \vert \phi'_k(z) \vert^2 \right) \cdot \underbrace{\left( \sum_{k=1}^{\infty} \vert \overline{\phi_k(w)} \vert^2 \right)}_{=K(w,w)} dV_w \\
&=& \sum_{k=1}^{\infty} \vert \phi'_k(z) \vert^2 \cdot \int_{G}
K(w,w) dV_w .
\end{eqnarray*}
Here, we applied the H\"older inequality and in the last line we
applied Lemma \ref{drei} stating that $\sum_{k=1}^{\infty} \vert
\phi'_k(z) \vert^2 = K_{\bar{z}z} < \infty$.
As $K(z,z)$ is $L^2$-integrable in both components (because $K_z(z,\cdot)$ is an element of the Hardy space for any $z$ ), the assertion turns out to be true.\\
We consider the reproducing property:
\[
f(z) = \int_{\partial G} K(z,w) f(w) dS_w .
\]
Since $K_z(z,w)$ is again an element of the Hardy space, we can interchange the order of the conjugated $D$-operator with the integral and we obtain
\begin{eqnarray*}
\bar{D}_z f(z) &=& \int_{\partial G} \bar{D}_z K(z,w) f(w) dS_w \\
&=& \int_{\partial G} \overline{K(z,w) D_z } f(w) dS_w \\
&=&\int_{\partial G} \overline{K_{\bar{z}}(z,w)} f(w) dS_w \\
&=& \langle K_{\bar{z}}(z,w), f(w) \rangle_w .
\end{eqnarray*}
Now let
\[
\tilde{M} := K_z(z,z) K(z,w) - K(z,z) K_{\bar{z}}(z,w) .
\]
Then
\begin{eqnarray*}\label{norm_M}
\langle \tilde{M}, \tilde{M} \rangle_w &=& \vert K_z(z,z) K(z,w) \vert^2 + \vert K(z,z) K_{\bar{z}}(z,w) \vert^2 \\
&& - K_z(z,z) K(z,z) \langle K(z,w), K_{\bar{z}}(z,w) \rangle \\
&& - K(z,z) \langle K_{\bar{z}}(z,w), K(z,w) \rangle \overline{K_z(z,z)} \\
&=& K_z(z,z)^2 \underbrace{K(z,z)}_{\langle K(z,w), K(z,w) \rangle} + K(z,z)^2 \langle K_{\bar{z}}(z,w) K_{\bar{z}}(z,w) \rangle \\
&& - K_z(z,z) K(z,z) \langle K(z,w), K_{\bar{z}}(z,w) \rangle \\
&& - K(z,z) \langle K_{\bar{z}}(z,w), K(z,w) \rangle \overline{K_z(z,z)} \\
&=& K_z(z,z)^2 K(z,z) + K(z,z)^2 K_{z\bar{z}}(z,z) \\
&& - K_z(z,z) K(z,z) K_{\bar{z}}(z,z) - K(z,z)K_z(z,z) \overline{\underbrace{K_z(z,z)}_{\in \mathbb{R}}} \\
&=& K(z,z)^2 K_{z\bar{z}}(z,z) - K(z,z) K_z(z,z) K_{\bar{z}}(z,z) \\
&=& K(z,z) \left( K(z,z) K_{z\bar{z}}(z,z) - K_z(z,z) K_{\bar{z}}(z,z) \right) \\
\end{eqnarray*}
Suppose that $\tilde{M} \equiv 0$. Then we obtain that
\begin{eqnarray*}
0 = \langle \tilde{M}, w- z \rangle_w & = & \langle K_z(z,z) K(z,w) - K(z,z) K_{\bar{z}}(z,w), w- z \rangle_w \\
&=& K_z(z,z) \langle K(z, w), w- z \rangle - K(z,z) \langle K_{\bar{z}}(z, w), w- z \rangle \\
&=& K_z(z,z)(z - z) - K(z,z) \left[ \left. \overline{D}_w (w - z) \right|_{w=z} \right] \\
&=& (n+1) K(z,z) > 0 .
\end{eqnarray*}
This is a contradiction.
\end{beweis}
With these preliminary results we are able to prove the main result of this section:
\begin{theorem}\label{thm4}
Let $G$ be a bounded domain. Then the curvature of the Szeg\"o metric is negative on $G$.
\end{theorem}
\begin{proof}
According to the definition of the Gau{\ss}ian curvature it suffices to show that
\[
\Delta \log K^2 (z,z) > 0.
\]
Since $K(z,z)$ is real, the product formula for the $D$-operator simplifies and we determine that
\begin{eqnarray*}
\Delta \log K^2 (z,z) &=& \bar{D} D K(z,z) \\
&=& \bar{D} \frac{1}{K(z,z)} K_z(z,z) \\
&=& \dfrac{K_{\bar{z}z}(z,z) K(z,z) - K_z(z,z) K(z,z) \bar{D} }{K(z,z)^2} \\
&+& \dfrac{K_{\bar{z}z}(z,z) K(z,z) - K_z(z,z) K_{\bar{z}}(z,z) }{K(z,z)^2} \\
\end{eqnarray*}
Here we applied in the last line that $K(z,z) \in \mathbb{R}$ According to
Lemma \ref{vier}, the last expression is positive, whereby the
theorem is proven.
\end{proof}
\section{The completeness result}
In order to show that the Szeg\"o metric is complete on bounded domains we will introduce first another metric which will be called the Szeg\"o-Caratheodory metric and show first that the latter one is complete.
The following two propositions are needed for the proof of the completeness of the
Szeg\"o-Carath\'{e}odory metric. Proofs of analogous statements for the case dealing with functions in several complex variables can be found for instance in \cite{Kran04}, chapter 3. Although this book deals with the function theory in several complex variables, the proofs only make use of arguments from real differential calculus in $\mathbb{R}^m$ and can be directly transferred due to the isometric isomorphy between $\mathbb{R}^{2^m}$ and $\mathcal{C}l_{0m}$ as normed vector spaces.
\begin{theorem}\label{cara_help1}
If $U$ is a domain with $C^2$ boundary, then there is an open neighbourhood $W$ of $\partial U$ such that if $z \in U \cap W$, then there is a unique point $P = P(z) \in \partial U$ which has the shortest Euclidean distance to $z$.
\end{theorem}
\begin{theorem}\label{cara_help2}
Let $U \subseteq \mathbb{C}$ be a bounded domain with $C^2$ boundary. Then there is an $r_0 > 0$ such that for each $P \in \partial U$ there is a disc $D(C(P),r_0)$ of radius $r_0$ which is externally tangent to $\partial U$ at $P$. There is also a disc $D(C'(P),r_0)$ which is internally tangent to $\partial U$ at $P$. This disc has furthermore the property that $D(C(P),r_0) \cap \partial U = \{P\}$ and $D(C'(P),r_0) \cap \partial U = \{P\}$.\\
\end{theorem}
We introduce
\begin{definition}
Let $G$ be a domain. The associated Hermitian metric defined by
\[
d_C(z) := \sup \{ |\overline{D} f (z)| ; f \in H^2(G), f(z)=0, \Vert f \Vert = 1 \}
\]
is called the Szeg\"o-Carath\'{e}odory metric. Here, and in all that follows $\|\cdot\|$ denotes the norm in the Hardy space $H^2$ that is induced by the inner product defined previously.
\end{definition}
\begin{remark}
The function $d_C$ is well-defined and positive. For a finite domain $G$ and $z_0 = \sum_{i=0}^m z^0_i e_i \in G$ we consider the function
\[
\mathfrak{Z}_1(z) := z_1 - z_0 e_1 - z^0_1 + z^0_0 e_1 .
\]
As one can easily verify, $\mathfrak{Z}_1$ is an element of the Hardy space over $G$ and has $z_0$ as a zero. Furthermore, we have
\[
\overline{D} \mathfrak{Z}_1 = - 2 e_1 \neq 0.
\]
As now $\frac{\mathfrak{Z}_1}{\Vert \mathfrak{Z}_1 \Vert}$ falls under the definition of the set from the definition of $d_C$, we have $d_C(z) > 0$. Then again holds for each function $f$ of the Hardy space
\begin{eqnarray*}
(\overline{D} f)(z) & = & \overline{D_z} \int_{\partial G} K(z,w) f(w) dS_w\\
& = & \int_{\partial G} \overline{D_z} K(z,w) f(w) dS_w\\
& = & \langle \overline{D_z} K(z,\cdot), f \rangle \\
& \leq & \underbrace{\Vert \overline{D_z} K(z,\cdot) \Vert}_{< \infty} \underbrace{\Vert f \Vert}_{=1} \\
& = &\Vert \overline{D_z} K(z,\cdot) \Vert < \infty ,
\end{eqnarray*}
so in particular the supremum is really finite. Here, again H\"older's inequality has been used in the estimate.
\end{remark}
Next we need to prove the following intertwining property of the operator $\overline{D}$:
\begin{theorem}\label{kettenregel}
Let $G \subseteq \mathbb{R}^m$ be a domain and $f \in C^1(G)$. Then we have that for each M\"obius transformation $ M =\left( \begin{array}{cc} a & b \\ c & d \\ \end{array} \right) $
\[
\left|\overline{D} (\frac{\overline{cz + d}}{|cz+d|^m} f)(M<z>)\right| = \left| \frac{\overline{cz + d}}{|cz+d|^{m+2}} \left( \overline{D}f\right)(M<z>) \right|,
\]
if $f(M<z>) = 0$.
\end{theorem}
\begin{proof}
We apply the product formula (\ref{produktregel}). As we restrict to consider functions $f$ with $f(M<z>)=0$ only, we can neglect the term $\left(\overline{D}\frac{\overline{cz+d}}{|cz+d|^m}\right) f(M<z>)$. This leaves only with terms depending on partial derivatives of $f$. If we now compare the product rules for the $D$ and the $\overline{D}$ operator, we see that the terms including $\partial_0 f$ are identical, while the terms involving $\partial_i f$, $i>0$, are identical up to a multiplication with $(-1)$. As the proposition has already been proven by Ryan in \cite{Ryan93} with the $D$ operator replacing $\overline{D}$, we can verify the equation for the $\partial_0 f$ terms using the identity
\[
Dg + \overline{D} g = 2 \partial_0 g, \qquad g \in C^1(G).
\]
After that we can show the identity also holds for the terms involving $\partial_i f$, $i>0$,
using
\[
Dg - \overline{D} g = \sum_{i=1}^m \partial_i g, \qquad g \in C^1(G).
\]
\end{proof}
The following theorem generalizes the distance decreasing property of the classical Carath\'{e}odory metric, compare for example with \cite{Kran04,Kran08}.
\begin{theorem}\label{metrik_ungleichung}
For each M\"obius transformation $ M =\left( \begin{array}{cc} a & b \\ c & d \\ \end{array} \right) $ and each domain $G$ we have
\[
\left|\frac{\overline{cz + d}}{|cz+d|^m}\right| d_C^{G}(M<z>) \leq d_C^{M<G>}(z).
\]
\end{theorem}
\begin{proof}
\begin{eqnarray*}& &
|\frac{\overline{cz + d}}{|cz+d|^m}| d_C(M<z>)\\ & = & |\frac{{cz + d}}{|cz+d|^m}| \sup \{ |\overline{D}(f(M<z>))|; f \in H^2, f(M<z>)=0 \}\\
& = & \sup \{ |\frac{{cz + d}}{|cz+d|^m}(\overline{D}f)(M<z>))|; f \in H^2, f(M<z>)=0 \} \\
& = & \sup \{ |\overline{D}(\frac{{cz + d}}{|cz+d|^m}f(M<z>))|; f \in H^2, f(M<z>)=0 \}\\
& \leq & \sup \{ |\overline{D}f(w))|; f \in H^2(M<G>), f(w)=0 \}
\end{eqnarray*}
\end{proof}
With this tool in hand we can establish
\begin{theorem}
The Szeg\"o-Carath{\'e}odory metric is complete on each finite domain $G$ with a $C^2$-smooth boundary.
\end{theorem}
\begin{proof}
Let $z \in G$ and $P$ be the point in $\partial G$ with minimal Euclidian distance to $z$. Let furthermore $\delta$ be the Euclidean distance of $z$ to the boundary. According to (\ref{cara_help1}), (\ref{cara_help2}) there exists a $r_0 > 0$ and $C(P) \in \mathbb{R}^{m+1} \setminus G$ so that $D(C(P),r_0) \cap \partial G = \{P\}$.\\
We now define the maps
\[
i_P: z \mapsto \frac{r_0 \overline{z} - \overline{C(P)}}{|z - C(P)|^2} + C(P)
\]
and
\[
j_P: z \mapsto \frac{1}{r_0} \left( z - C(P) \right) .
\]
Both $i_P$ and $j_P$ are apparently M\"obius tranformations.\\
For their composition it holds
\begin{eqnarray*}
(j_P \circ i_P)(z) & = & \frac{1}{r_0} \left( r_0^2 \frac{\overline{z}-\overline{C(P)}}{|z - C(P)|^2} + C(P) - C(P) \right) \\
& = & r_0 \frac{\overline{z}-\overline{C(P)}}{|z - C(P)|^2} \\
& = & r_0 \left( z - C(P) \right)^{-1}
\end{eqnarray*}
From the choice of $z$ and $C(P)$ we gain for the absolute value
\[
|(j_P \circ i_P)(z)| = r_0 \frac{1}{\delta + r_0} = 1 - \frac{\delta}{\delta + r_0}.
\]
We now consider the automorphic factor belonging to the M\"obius tranformation $j_P \circ i_P$ according to (\ref{kettenregel}):
\[
\left| r_0 \frac{\overline{z}-\overline{C(P)}}{|z - C(P)|^{m+1}} \right| = r_0 \frac{1}{(\delta + r_0)^{m}}
\]
Applying now Theorem \ref{metrik_ungleichung} leads to
\begin{equation}\label{2te_ungl}
d_C^G(z) \geq \left| r_0 \frac{\overline{z}-\overline{C(P)}}{|z - C(P)|^{m+1}} \right| d_C^{\mathbb{D}}((j_P \circ i_P)(z)) = \left| r_0 \frac{\overline{z}-\overline{C(P)}}{|z - C(P)|^{m+1}} \right| d_C^{\mathbb{D}}(\frac{r_0}{z - C(P)}).
\end{equation}
We thereby denote the unit ball in $\mathbb{R}^{m+1}$ with $\mathbb{D}$.\\
To estimate the Szeg\"o-Carath{\'e}odory metric on the unit ball, we define a test function.\\
Let $\mathfrak{K}$ be the Cauchy kernel and $w_0 \in \mathbb{R}^{m+1} \setminus G$ so that $|z-w_0|=2\delta$. then it holds that
\[
D\mathfrak{K}( \cdot - w_0) = 0
\]
on $G$ and $\mathfrak{K}$ is welldefined on $G$.\\
We now set
\[
\mathfrak{K}_2 (z) := \overline{D_z} \left( \mathfrak{K} ( \cdot - w_0) \right) - \overline{D_z} \left( \mathfrak{K} ( \frac{r_0}{\delta + r_0} - w_0) \right) .
\]
By construction we have $\mathfrak{K}_2 \in H^2(G, \mathcal{C}_{0m})$ and $\mathfrak{K}_2 (\frac{r_0}{\delta + r_0} - w_0) = 0$.\\
Applying this property leads to
\[
d_C^{\mathbb{D}}(\frac{r_0}{\delta + r_0}) \geq \left| \left( \overline{D} \mathfrak{K}_2 \right) (\frac{r_0}{\delta + r_0}) \right|.
\]
We obtain that
\begin{eqnarray}
\overline{D}^2 & = & (\partial_0 + \overrightarrow{D})^2 \\
& = & \partial_0^2 - \partial_0 \overrightarrow{D} - \overrightarrow{D} \partial_0 - \sum_{i=1}^{m-1} \partial^2_i,
\end{eqnarray}
so
\[
\Delta \mathfrak{K}_2 = 0 ,
\]
and therefore
\[
- \sum_{i=1}^{m} \partial^2_i \mathfrak{K}_2 = \partial_0^2 \mathfrak{K}_2
\]
As $\mathfrak{K}_2 \in C^{\infty}(G)$, we get
\[
\partial_0 \overrightarrow{D} \mathfrak{K}_2 = \overrightarrow{D} \partial_0 \mathfrak{K}_2 .
\]
Because of $D \mathfrak{K}_2 = 0$, we have
\[
\overrightarrow{D} \mathfrak{K}_2 = - \partial_0 \mathfrak{K}_2.
\]
From this property we can deduce that
\begin{eqnarray*}
\overline{D}^2 \mathfrak{K}_2 (z) & = & 4 \partial_0^2 \mathfrak{K}_2 (z) \\
& = & 4 \partial_0 \frac{|z-w_0|^{2} - (m+1) (\overline{z-w_0})(z_0 - {w_0}_0)}{|z-w_0|^{m+3}} \\
& = & 4 \Bigg[ \frac{-(m+1)(z_0 - {w_0}_0)}{|z-w_0|^{m+3}}\\ & & \quad\quad - (m+1)\frac{-\overrightarrow{z-w_0}}{|z-w_0|^{m+3}} + (m+3)\frac{\overline{z-w_0}(z_0 - {w_0}_0)^2}{|z-w_0|^{m+5}} \Bigg] \\
& = & 4 \left[ -(m+1) \frac{ \overline{z-w_0} }{|z-w_0|^{m+3}} \right].
\end{eqnarray*}
As only small values of $\delta$ are relevant in our case, we can assume that $\delta<1$ without loss of generality. Under this assumption, we can estimate
\[
|\overline{D}^2 \mathfrak{K}_2 (z)|
\]
from below through
\[
C_1 \frac{1}{|z-w_0|^{m+2}} = C_1 \frac{1}{(2\delta)^{m+2}}
\]
with a constant $C_1 > 0$.\\
Furthermore, we can estimate $\Vert \overline{D} \mathfrak{K}_2 (z) \Vert$ from below through $\tilde{C}_1 \frac{1}{(2 \delta)^{m+1}}$.\\
Together with (\ref{2te_ungl}) we then have
\begin{equation}\label{scm_endabschaetzung}
d_C^G(z) \geq r_0 \frac{1}{(\delta + r_0)^{m}} C_1 \frac{1}{(2\delta)^{m+2}} \frac{1}{\tilde{C}_1 \frac{1}{(2 \delta)^{m+1}}} \geq C_2(r_0) \frac{1}{\delta}
\end{equation}
with a number $C_2 > 0$ which only depends on $r_0$ .\\
Let now $E$ be a set which is bounded in the Szeg\"o-Carath{\'e}odory metric, then
\[
d_C^G(z) < M
\]
for all $z \in E$ for an $M>0$. According to (\ref{scm_endabschaetzung}) this yields to
\[
\delta \geq C_3(r_0) M.
\]
So there exists a positive minimal distance of $E$ to $\partial G$. Due to the boundedness of $G$ $E$ then is relatively compact. Thus, the Szeg\"o-Carath{\'e}odory metric is complete on $G$.
\end{proof}
To show the completeness of the Szeg\"o metric, we first prove an inequality between it and the Szeg\"o-Carath\'{e}odory metric on a bounded domain, which will be essential for the final proof. The proof of the completeness of the Bergman metric in the theory of several complex variables makes use of a similar inequality, which can be found in \cite{Koba62,Hahn76,Hahn77,Hahn78} and \cite{Kran01}.
\begin{theorem}\label{relation_s_c_metriken}
Let $C_G$ be the Szeg\"o-Carath{\'e}odory metric of a domain $G$. Then
\begin{enumerate}
\item[a)] $ds_G \geq C_G$
\item[b)] The metric $ds_G^*$ given by $ \Delta \log S(z,z) |dz|$ satisfies the inequality
\[
ds_G^* \geq C_G.
\]
\end{enumerate}
\end{theorem}
\begin{proof}
Let furthermore be $f'(z):=\overline{D} f$.\\
Let $z \in G$ $f \in H^2$ with $f(z)=0$ and $\| f \| \leq 1$. We further define
\[
M:=K(z,z)K_{\bar{z}}(z, \cdot) - K_{z}(z, z)K(z, \cdot) .
\]
It holds that
\[
0 = f(z) = \langle f, K(z, \cdot)\rangle .
\]
From this property we may infer that
\begin{eqnarray*}
|f'(z)| &=& \frac{1}{K(z,z)} \langle f, K(z,z)K_{\bar{z}}(z, \cdot) - K_{z}(z, z)K(z, \cdot)\rangle \\
&\leq & \underbrace{\| f \|}_{\leq 1} \frac{1}{K(z,z)} \| M \| \\
&\stackrel{(\ref{norm_M})}{\leq}& \frac{\sqrt{K(z,z)(K(z,z)K_{z\bar{z}}(z,z) - K_z(z,z)K_{\bar{z}}(z,z)) }}{K(z,z)} \\
&=& \frac{\sqrt{(K(z,z)K_{z\bar{z}}(z,z)) - K_z(z,z)K_{\bar{z}}(z,z) }}{\sqrt{K(z,z)}} \\
&=&\frac{\sqrt{M'(z)}}{\sqrt{K(z,z)}} \\
&=& \sqrt{\Delta \log K^2 (z,z)}.
\end{eqnarray*}
In the second line, we applied H\"older's inequality. In the last line we notice that the argument of the square root is really positive. This is a consequence of Theorem~\ref{thm4}.
Next, each
function $f \in H^2(G, \mathcal{C}l_{0m} )$ with $\Vert f \Vert
\leq 1$ satisfies
\begin{eqnarray*}
|f(z)| &=& |\langle f, K(z, \cdot) \rangle| \\
& \leq & \Vert f \Vert \Vert K(z, \cdot) \Vert \\
&\leq& \sqrt{\langle K(z, \cdot), K(z, \cdot) \rangle}\\
&=& \sqrt{K(z,z)}.\\
\end{eqnarray*}
Again, in the second line we applied H\"older's inequality. Notice that we require the result of lemma \ref{vier} for the root to be well-defined.\\
Furthermore, we have with $g(w):= \frac{M(w)}{\Vert M \Vert}$ that
$\Vert g \Vert \leq 1$ and
\begin{eqnarray*}
d_C(z) &\geq & |g'(z)| \\
&=& |\frac{M'(z)}{\Vert M \Vert}| \\
&=& \frac{\Vert M \Vert^2}{\Vert M \Vert K(z,z)} \\
&=& \frac{\Vert M \Vert}{K(z,z)}
\end{eqnarray*}
A simple calculation shows that $\Vert M \Vert = \Vert M' \Vert$.\\
From this equality it follows as a consequence of $DM'=0$ that
\[
\frac{M'(z)}{\Vert M \Vert} \leq \sqrt{K(z,z)},
\]
and thus
\[
\Vert M \Vert \frac{1}{K(z,z)} \leq \sqrt{K(z,z)}.
\]
So, powers of both metrics can be estimated from below through the Szeg\"o-Carath{\'e}odory metric. But from this property we may draw the conclusion that each set $E$ which is bounded in these metrices is also bounded in the Szeg\"o-Carath{\'e}odory metric. As this metric is complete, $E$ must be relatively compact in $G$. Thus both these metrices are also complete and the propostion is proven.
\end{proof}
From the completeness of the Szeg\"o-Carath{\'e}odory metric we can finally establish our main result of this section:
\begin{theorem}
The Szeg\"o metric is complete.
\end{theorem}
\begin{proof}
In regard of Theorem \ref{relation_s_c_metriken}, we have for all $z, w \in G$
\[
d_C(z,w) \leq d_S(z, w).
\]
So each $d_S$-bounded set $H \subset G$ is simultaneously
$d_C$-bounded. As the Szeg\"o-Carath{\'e}odory metric is complete,
$H$ must be relatively compact in $G$. With this, each
$d_S$-bounded set is also relatively compact in $G$ and $d_S$ thus
complete.
\end{proof}
Perspectives:
We have already mentioned that in the theory of several complex variables there exists a strong connection between the completeness of the metric and the smoothness of the boundary of the domain. So it is an obvious point of investigation for further works if and how these propositions can be transferred to the Clifford case, as our hitherto existing results require a strong smoothness ($C^2$-boundary).
\par\medskip\par Acknowledgements. The second author gratefully acknowledges the fruitful und helpful discussions with Dr. Denis Constales from Ghent University on the topic of this paper which lead to the successful development of it.
|
\section{Introduction}
Neutrinos, being neutral particles, cannot couple directly to photons.
They can couple indirectly via weak interaction and other effects, potentially
leading to significant observational effects in astrophysics~\cite{Adams-1963,Bandyopadhyay:1969ck}.
The photon-neutrino scattering cross section predicted by the Standard Model is, however,
exceedingly small and likely of little practical importance in astrophysics,
see e.g.~\cite{Dicus:1993iy}.
Field theories on noncommutative (NC) spaces offer a different interaction channel
for neutrinos and photons:
Such theories have been introduced as effective
models\footnote{These models should be understood as effective theories
and are not necessarily renormalizable.}
for the quantum geometric structure of spacetime
that is generically expected in any reasonable quantum theory of gravity
(including string theory). The presence of gauge fields influences such
noncommutative structures and neutrinos propagating in
the modified background feel the effect. The non-zero star commutator
\begin{equation}
[ A_\mu\stackrel{\star}{,}\Psi]= A_\mu\star\Psi-\Psi\star A_\mu.
\label{1.1}
\end{equation}
is a striking illustration of this effect in noncommutative $U_{\star}(1)$ theory:
Fermions
that are neutral or even sterile in the commutative limit can nevertheless
directly couple to a $U_{\star}(1)$
gauge field. In this article we shall investigate such noncommutativity induced
interactions using methods that are non-perturbative in the noncommutativity
parameters~\cite{Schupp:2008fs}.
Studies in noncommutative particle
phenomenology \cite{Hinchliffe:2002km}, aiming at predicting possible
experimental signatures and estimating bounds on space-time
noncommutativity from existing experimental data, started around the time that
string theory indicated that noncommutative gauge theory could
be one of its low-energy effective
theories \cite{Seiberg:1999vs}. Early attempts with simple models based on star products
(with expressions like the one given above) soon ran into several serious difficulties:
(i)~Such theories have no local gauge invariant quantities,
(ii)~fields like $ A_\mu$ do not
transform covariantly under coordinate changes~\cite{Jackiw:2001jb} and
(iii)~there are unphysical restrictions on representations and charges.
These shortcomings are overcome in an approach based on Seiberg-Witten (SW) maps,
which enables one to deform commutative gauge theories with essentially
arbitrary gauge group and
representation~\cite{Madore:2000en,Jurco:2000ja,Bichl:2001gu,Jurco:2000fb,Jurco:2001my,Jurco:2001rq}.
Since this approach also fixes problems of non-covariance under coordinate changes,
we shall refer to this class of theories as \emph{covariant noncommutative field theories}.%
\footnote{All these models are consistent, contrary to slightly misleading statements in
a recently published ``no-go theorem'' \cite{Chaichian:2009uw},
where a mal-constructed model is shown to lead to contradictions.}
After some initial enthusiasm, renormalizability of these theories turned out
to be a delicate issue. In \cite{Wulkenhaar:2001sq,Grimstrup:2002af} a Dirac fermion 4-vertex was
identified as one of the major culprits in this issue. This vertex is absent
in theories with chiral fermions \cite{Buric:2007ix,Martin:2009sg}.
Particularly well behaved are theories where
the gauge fields couple via a star commutator (\ref{1.1}) to fermions,
as is the case in our model: Cancelations between fermion and boson loops lead
to softened UV/IR coupling \cite{Matusis:2000jf}. Regardless of the question of
renormalizability, SW-map based models serve as effective theories for some
of the quantum geometric effects expected in more fundamental theories such
as string theory and quantum gravity.
In the Seiberg-Witten map approach, noncommutative fields $ A_\mu$, $ \Psi$,\ldots and
gauge transformation parameters $ \Lambda$ are interpreted as non-local, enveloping algebra-valued
functions of their commutative counterparts $a_\mu$, $\psi$, $\lambda$ and of the
noncommutative parameters $\theta^{\mu\nu}$, in such a way that
ordinary gauge transformations of the commutative fields induce noncommutative
gauge transformations of the noncommutative fields.
This procedure allows the construction of noncommutative extensions of
important particle physics models like the NC Standard Model (NCSM) and GUT models
\cite{Bichl:2001cq,Calmet:2001na,Behr:2002wx,Aschieri:2002mc,Melic:2005fm,Melic:2005am,Martin:2011un},
as well as various follow-on studies with NC modifications of particle physics
\cite{Ohl:2004tn,Ohl:2004ke,Melic:2005su,Alboteanu:2005gj,
Alboteanu:2006hh,Alboteanu:2007bp,Alboteanu:2007by,Buric:2007qx}.
In more recent development, it has been found that SW expanded
models at first order in~$\theta$ are well-behaved regarding
anomalies and renormalizability. For example the NCSM \cite{Calmet:2001na} at
$\theta$-order appears to be anomaly free \cite{Martin:2002nr,Brandt:2003fx},
has remarkably well-behaved one-loop quantum corrections
\cite{Buric:2006wm} and breaks Lorentz symmetry; see also
\cite{Bichl:2001cq,Latas:2007eu,Martin:2009vg,Tamarit:2009iy,Buric:2010wd}.
To avoid strong backgrounds from known processes, considerable
efforts in noncommutative phenomenology has been directed at
interactions which are suppressed in Standard Model settings.
One important candidate or this type is the aforementioned tree-level coupling of neutrinos
with photons or, more precisely, plasmons.
Such interactions have already been studied in the framework of
noncommutative gauge theories defined by Seiberg-Witten
maps \cite{Schupp:2002up, Minkowski:2003jg}.
There, like in almost all other studies of covariant NC field theory, an expansion and cut-off
in powers of the noncommutativity parameters $\theta^{\mu\nu}$
was used for computational simplicity.
Such an expansion corresponds to an expansion in momenta (derivatives) and restrict the range of
validity to energies well below the noncommutativity scale $\Lambda_{\rm NC}$.
This is usually no problem for experimental predictions because the noncommutativity parameters
$\theta^{ij}=c^{ij}/\Lambda_{\rm NC}^2$ are in general considered to be
small. There exists, however, exotic processes like ultra high energy cosmic
rays \cite{Horvat:2010sr} in which the interacting energy scale runs
higher than the current experimental
bound on the noncommutative scale $\Lambda_{\rm NC}$. Here the previously available approximate
results are inapplicable. To overcome the $\theta$-expansion and cut-off approximation,
we are using in this article
$\theta$-exact expressions and expand in powers of the coupling constant as in ordinary gauge theory.
The $\theta$-exact approach has been inspired by exact formulas for the Seiberg-Witten
map \cite{Jurco:2001my,Mehen:2000vs,Liu:2000mja,Okawa:2001mv}.
For arbitrary non-Abelian gauge theories the $\theta$-exact approach is still a challenging problem,
in particular in loop computations and at higher orders in the coupling constant.
Since perturbative renormalization and UV/IR mixing are still fairly poorly understood,
it is not clear how to
interpret the quantum corrections and to relate them to
observations \cite{Abel:2006wj,Horvat:2010km}.
Another interesting venue for applications of
$\theta$-exact methods is the investigation of quantum corrections in
covariant noncommutative quantum field theories.
Recently it was suggested~\cite{Vilar:2009er} that noncommutative QED
might be renormalizable by adding proper counter-terms. A covariant $\theta$-exact
version of this theory could be a very
interesting object for future studies.
First $\theta$-exact results have been published in the investigation of
UV/IR mixing in covariant NC gauge theory~\cite{Schupp:2008fs}
and later in the context of NC photon-neutrino phenomenology, namely scattering of
ultra high energy cosmic ray neutrinos on nuclei \cite{Horvat:2010sr}.
Those topics were off-limits in the old $\theta$-expansion method.
\section{Model}
In this section we recall some basic facts about covariant noncommutative gauge theory based
on Seiberg-Witten maps. We then review the derivation of $\theta$-exact interaction terms
and apply the method to the computation of neutrino-photon tri-particle vertices.
We close with a comment on an alternative covariant vertex.
Throughout the article we shall concentrate on \emph{Abelian} noncommutative gauge theory.
It is straight-forward to formulate field theories on noncommutative spaces by inserting
a star product~$\star$ between all fields in the action. This introduces ordering ambiguities
and it breaks ordinary gauge invariance (because local gauge transformations do not commute
with star products).
In analogy to the introduction of covariant derivatives in gauge theory,
the star product can be promoted to a gauge-field dependent covariant star product~$\star'$.
Together with a gauge-field dependent covariant integral measure this leads to a noncommutative
gauge theory.
Alternatively and in fact equivalently~\cite{Jurco:2000fb,Jurco:2001my}
it is possible to retain the original star product and instead promote all fields
to noncommutative fields and the gauge transformations to noncommutative gauge transformations.
In this construction the ``noncommutative fields'' are obtained via
Seiberg-Witten maps~\cite{Seiberg:1999vs} and their generalizations from the original
``commutative fields''.
With some field-ordering ``fine-tuning'', it is possible to obtain noncommutative models,
were neutrinos and other neutral fermion fields do not couple to photons
-- the minimal NC Standard Model is an example of this type. More generically, however,
electrically neutral matter fields will be promoted via (hybrid) Seiberg-Witten maps
to noncommutative fields that couple via star commutator to photons and transform in
the adjoint representation of $U_{\star}(1)$ -- this is the case for phenomenologically
promising NC GUTs.
The inclusion of all gauge covariant coupling terms is furthermore a prerequisite
for reasonable UV behavior.
Taking all this into account we eventually arrive at the following model of a
Seiberg-Witten type noncommutative $U_{\star}(1)$ gauge theory
\footnote{In this section we set the
coupling constant $e=1$, to restore the coupling constant one simply
substitute $a_\mu$ by $ea_\mu$, then divide the gauge-field
Lagrangian by $e^2$.} (for more details, see the discussion at the end of the section)
\begin{equation}
S=\int\left(-\frac{1}{4}F^{\mu\nu}F_{\mu\nu}+i\bar\Psi
\left(\slashed{D}-m_{\nu}\right)\Psi\right) \, d^4x
\label{swqed:action}
\end{equation}
with $F_{\mu\nu}=\partial_\mu
A_\nu-\partial_\nu A_\mu-i[A_\mu\stackrel{\star}{,}A_\nu]$ and
\begin{equation}
D_\mu\Psi=\partial_\mu\Psi-i[A_\mu\stackrel{\star}{,}\Psi] \,.
\label{swqed:actionD}
\end{equation}
All the fields in this action are images under (hybrid) Seiberg-Witten maps
of the corresponding commutative fields $a_\mu$ and $\psi$.
In the original work of Seiberg and Witten and in virtually all subsequent applications,
these maps are understood as (formal) series in powers of the noncommutativity parameter
$\theta^{\mu\nu}$. Physically, this corresponds to an expansion in momenta and is valid
only for low-energy phenomena. Here we shall not subscribe to this point of view and instead
interpret the noncommutative fields as valued in the enveloping algebra of
the underlying gauge group.
This naturally corresponds to an expansion in powers of the gauge field $a_\mu$
and hence in powers of
the coupling constant. At each order in $a_\mu$ we shall determine $\theta$-exact expressions.
The expansion in powers of the commutative (gauge) field content is motivated
from the obvious fact that in
perturbative quantum field theory one can sort the interaction
vertices by the number of external legs and this is equivalent to the
number of field operators in the corresponding interacting terms.
For any specific process and loop order there exists an upper limit on
the number of external legs. So if one expands the
noncommutative fields with respect to the formal power of the
commutative fields which are the primary fields in the theory up to
an appropriate order, the relevant vertices in a specific diagram
will automatically be exact to all orders of $\theta$.
In tree-level neutrino-photon coupling processes
only vertices of the form $a\bar\psi\psi$ contribute, therefore an
expansion to lowest nontrivial order in $a_\mu$ (but all orders in $\theta$) is enough.
There are at least three known methods for $\theta$-exact computations:
The closed formula derived using deformation quantization based on
Kontsevich formality maps~\cite{Jurco:2001my}, the relationship
between open Wilson lines in the commutative and noncommutative
picture~\cite{Mehen:2000vs,Okawa:2001mv}, and direct recursive
computations using consistency conditions. For the lowest nontrivial order a
direct deduction from the recursion and consistency relations
\begin{gather}
\delta_\Lambda A_\mu \equiv \partial_\mu\Lambda +
i[\Lambda \stackrel{\star}{,}A_\mu]
\nonumber \\= A_\mu[a_\mu+\delta_\lambda
a_\mu]-A_\mu[a_\mu] + \mathcal O(\lambda^2)\,,
\\
\delta_\Lambda\Psi \equiv i[\Lambda \stackrel{\star}{,}\Psi]
\nonumber \\
=\Psi[a_\mu+\delta_\lambda a_\mu,
\psi+\delta_\lambda\psi]
-\Psi[a_\mu,\psi] + \mathcal O(\lambda^2)\,,
\\
\Lambda[[\lambda_1,\lambda_2],a_\mu]
=[\Lambda[\lambda_1,a_\mu]\stackrel{\star}{,}
\Lambda[\lambda_2,a_\mu]]
\nonumber \\
+i\delta_{\lambda_1}\Lambda[\lambda_2,a_\mu]-i\delta_{\lambda_2}
\Lambda[\lambda_1,a_\mu]\,,
\label{SWrecurs}
\end{gather}
with the ansatz\footnote{Notation: Capital letters denote noncommutative objects,
small letters denote commutative objects, hatted capital letters denote differential
operator maps from the latter to the former.}
\begin{gather}
\Lambda=\hat\Lambda[a_\mu]\lambda=(1+\hat\Lambda^1[a_\mu]
+\hat\Lambda^2[a_\mu]+\mathcal O(a^3))\lambda\,,
\\
\Psi=\hat\Psi[a_\mu]\psi=(1+\hat\Psi^1[a_\mu]+\hat\Psi^2[a_\mu]+\mathcal
O(a^3))\psi\,,
\label{ansatz:linear}
\end{gather}
is already sufficient. Here $\hat\Psi[a_\mu]$ and $\hat\Lambda[a_\mu]$ are gauge-field
dependent differential operators that we shall now determine:
Starting
with the fermion field $\Psi$, at lowest order we have
\begin{equation}
i[\lambda\stackrel{\star}{,}\psi]=\hat\Psi[\partial\lambda]\psi\,.
\label{fermiconsiscondi}
\end{equation}
Writing the star commutator explicitly as
\begin{equation}
\begin{split}
[f\stackrel{\star}{,}g]&=f(x)(e^{i\frac{\partial_x\theta\partial_y}{2}}
-e^{-i\frac{\partial_x\theta\partial_y}{2}})g(y)\bigg|_{x=y}
\\&=2if(x)\sin(\frac{\partial_x\theta\partial_y}{2})g(y)\bigg|_{x=y}
\\&=i\theta^{ij}\bigg(\frac{\partial f(x)}{\partial x^i}\bigg)
\frac{\sin(\frac{\partial_x\theta\partial_y}{2})}
{\frac{\partial_x\theta\partial_y}{2}}
\bigg(\frac{\partial g(y)}{\partial y^{i}}\bigg)
\bigg|_{x=y}\,.
\end{split}
\label{starcommut}
\end{equation}
we observe that
\begin{equation}
\hat\Psi[a_\mu]=-\theta^{ij}a_i\star_2\partial_j
\label{fermionstar2}
\end{equation}
will fulfill the consistency relation. The generalized star product
$\star_2$~\cite{Mehen:2000vs} that appears here
is defined as follows
\begin{align}
\begin{split}
f\star_2 g&=f(x)\frac{\sin\frac{\partial_x\wedge
\partial_y}{2}}{\frac{\partial_x\wedge
\partial_y}{2}}g(y)\bigg|_{x=y}.
\end{split}
\label{star2}
\end{align}
The gauge transformation $\Lambda$ can be worked out similarly, namely
\begin{equation}
\begin{split}
0&=[\lambda_1\stackrel{\star}{,}\lambda_2]
+i\hat\Lambda[\partial\lambda_1]\lambda_2
-i\hat\Lambda[\partial\lambda_2]\lambda_1
\\
&=\frac{1}{2}([\lambda_1\stackrel{\star}{,}\lambda_2]
-[\lambda_2\stackrel{\star}{,}\lambda_1])
+i\hat\Lambda[\partial\lambda_1]\lambda_2
-i\hat\Lambda[\partial\lambda_2]\lambda_1\,,
\end{split}
\label{consisconfgt}
\end{equation}
and hence
\begin{equation}
\hat\Lambda^1=-\frac{1}{2}\theta^{ij}a_i\star_2\partial_j\,.
\label{star2gt}
\end{equation}
The gauge field $a_\mu$ requires slightly more work. The lowest order terms in its consistency
relation are
\begin{equation}
-\partial_\mu(\frac{1}{2}\theta^{ij}a_i\star_2\partial_j
\lambda)-i[\lambda\stackrel{\star}{,}a_\mu]
=A^2_\mu[a_\mu+\partial_\mu\lambda]-A^2_\mu[a_\mu]\,,
\label{conscondgf}
\end{equation}
where $A^2$ is the $a^2$ order term in the expansion of $A$ as power
series of $a$. The left hand side can be rewritten as
$-\frac{1}{2}\theta^{ij}\partial_\mu
a_i\star_2\partial_j\lambda-\frac{1}{2}
\theta^{ij}a_i\star_2\partial_\mu\partial_j\lambda
-\theta^{ij}\partial_i\lambda\star_2\partial_j a_\mu$,
where the first term comes from
$-\frac{1}{2}\theta^{ij}\partial_\mu a_i\star_2 a_j$, while the third one
comes from $-\theta^{ij}a_i\star_2\partial_j a_\mu$. After a gauge transformation,
the sum of the first and third terms equals the second term. Ultimately, we obtain
\begin{eqnarray}
A_\mu&=&a_\mu-\frac{1}{2}\theta^{ij}a_i\star_2(\partial_j
a_\mu+f_{j\mu})+\mathcal O(a^3)\,, \label{exactA}
\\
\Psi&=&\psi-\theta^{ij}a_i\star_2\partial_j\psi+\mathcal
O(a^2)\psi\,, \label{exactPsi}
\\
\Lambda&=&\lambda-\frac{1}{2}\theta^{ij}a_i\star_2\partial_j\lambda+\mathcal
O(a^2)\lambda\,, \label{exactLambda}
\end{eqnarray}
with $f_{\mu\nu}$ being the commutative field strength
$f_{\mu\nu}=\partial_\mu a_\nu-\partial_\nu a_\mu$.
Expanding the action \eqref{swqed:action} in the terms of the
commutative fields, one gets the following $\theta$-exact cubic terms up to first order in $a_\mu$:
\begin{equation}
\begin{split}
{\cal L}&=\bar\psi\gamma^\mu[a_\mu\stackrel{\star}{,}\psi]
-(\theta^{ij}a_i
\star_2 \partial_j\bar\psi)(i\slashed\partial - m_\nu)\psi\\&\quad-\bar\psi
(i\slashed\partial - m_\nu)(\theta^{ij}
a_i\star_2\partial_j\psi)+\bar\psi\mathcal{O}(a^2)\psi\,.
\label{pf:int}
\end{split}
\end{equation}
Here $\psi$ is a Dirac-type massive or massless (i.e. Weyl) neutrino field.
To extract Feynman rules in an appropriate form, we use the
arithmetic property
$i\theta^{ij}\partial_i f\star_2\partial_j g=[f\stackrel{\star}{,}g]$ to obtain
the effective neutrino-photon Lagrangian density
\begin{equation}
\begin{split}
{\cal L}
&=-(\theta^{ij}a_i\star_2\partial_j\bar\psi)(i\slashed\partial - m_\nu)\psi
\\-&\bar\psi(i\slashed\partial - m_\nu)(\theta^{ij}a_i\star_2\partial_j\psi)
\\+&i\bar\psi\gamma^{\mu}(\theta^{ij}\partial_i
a_\mu\star_2\partial_j\psi)+\mathcal O(a^2\bar\psi\psi)\,.
\label{calL}
\end{split}
\end{equation}
The rest of the derivation resembles that of NCQED without a Seiberg-Witten
map. Just like the Moyal-Weyl star product turns into an
exponential function in momentum space, the generalized star
product $\star_2$ turns into a function
\begin{equation}
F(q,k)=\frac{\sin \frac{q\theta k}{2}}{\frac{q\theta k}{2}},
\end{equation}
where $q$ and $k$ are the momenta of the fields involved in
the product. We notice that for tri-field interaction,
$4$-momentum conservation $q=k-k'$ renders the function $F$ independent
of the order of momenta involved:
$F(q,k)=F(k,q)=F(k,k')=F(k',k)$.
We can hence pull out $F$ as an universal factor.
In the end we obtain the following $\theta$-exact Feynman rule for the
neutrino-photon tri-particle vertex:
\begin{eqnarray}
\Gamma^{\mu}&=& i F(q,k) \left[(\slashed k - m_{\nu})\tilde q^{\mu}
+(q\theta k)\gamma^{\mu} -\slashed q\tilde k^{\mu}\right]\,,
\label{Feynrule}
\end{eqnarray}
with the shorthand notations $q\theta k \equiv q_i \theta^{ij} k_j$ and $\tilde k^\mu = \theta^{\mu j} k_j$.
If we compare this with to the first order in $\theta$ vertex that was used in previous work, we
see that only the factor $F(q,k)$ is new. Consequently, it is this factor
that leads to modifications in $\theta$-exact
computations.
Interestingly, the first term in (\ref{Feynrule}),
\begin{equation}
\Gamma^{\mu}_\text{alt} = i F(q,k) (\slashed k - m_{\nu})\tilde
q^{\mu} \label{simplevertex} \,,
\end{equation}
is already consistent with gauge invariance on its own. This simplified vertex defines an alternative
NC theory of neutrino-photon interaction that is attractive for computations beyond tree level.
In this article we will use the full vertex which is more natural from the point of view of
NC gauge theory as it is derived from a covariant derivative.
The two choices of NC vertices is ultimately related to a choice of generalized SW map in
the construction of the NC theory. In the remainder of
this section we shall explore this construction in more detail.
For simplicity of presentation we shall set $e=1$ and focus on the massless case.
We start with the action for a neutral massless free fermion field
\begin{equation}
S = \int \bar\psi\gamma^\mu\partial_\mu\psi \, d^4x =\int \bar\psi\star \gamma^\mu\partial_\mu\psi\, d^4x \,,
\end{equation}
where, as indicated, a Moyal-Weyl type star product can be inserted or removed by partial integration.
Following the method of constructing a covariant NC gauge theory outlined at the beginning of this section,
we lift the factors in the action via (generalized) Seiberg-Witten maps $\hat \Psi[a_\mu]$, $\hat \Phi[a_\mu]$
to noncommutative status as follows:
\begin{equation}
S=\int \hat\Psi(\bar\psi)\gamma^\mu \hat\Phi(\partial_\mu\psi) \, d^4x
= \int \hat\Psi(\bar\psi)\star \gamma^\mu \hat\Phi(\partial_\mu\psi) \, d^4x \,.
\end{equation}
Now if the SW maps $\hat\Psi$, $\hat\Phi$ and a corresponding map $\hat\Lambda$ for the gauge parameter $\lambda$
satisfy
\begin{equation}
\begin{split}
\delta_\lambda(\hat\Psi(\bar\psi))=i[\hat\Lambda(\lambda)\stackrel{\star}{,}\hat\Psi(\bar\psi)],\\
\delta_\lambda(\hat\Phi(\partial_\mu\psi))
=i[\hat\Lambda(\lambda)\stackrel{\star}{,}\hat\Phi(\partial_\mu\psi)],
\end{split}
\end{equation}
we will have a noncommutative action that is
gauge invariant under infinitesimal commutative gauge
transformations $\delta_\lambda$ and reduces to the free fermion action in the commutative limit $\theta\to 0$.
The appropriate map $\hat\Psi$ is the one (\ref{exactPsi}) that we have already derived:
\begin{equation}
\hat\Psi(\psi)=\psi-\theta^{ij}a_i\star_2\partial_j\psi+\mathcal
O(a^2)\psi.
\end{equation}
Recalling that we are dealing with neutral fields, i.e.\ $\delta \psi = 0$ and $\delta (\partial_\mu\psi) = 0$,
we notice that we can in principle use the same map also for
$\hat\Phi$:
\begin{equation}
\hat\Phi_\text{alt}(\partial_\mu\psi)=\hat\Psi(\partial_\mu\psi)
=\partial_\mu\psi-\theta^{ij}a_i\star_2(\partial_j\partial_\mu\psi)+\mathcal
O(a^2)\psi \,.
\end{equation}
This construction is quite unusual from the point of gauge theory, as it yields a covariant derivative
term without introducing a covariant derivative. In any case the resulting action
\begin{equation}
\begin{split}
S_\text{alt}&=\int \Big(
i\bar\psi\gamma^\mu\partial_\mu\psi-i\left(\theta^{ij}\partial_j\bar\psi
\star_2 a_i\right)\gamma^\mu\partial_\mu\psi
\\+&i\bar\psi\gamma^{\mu}\left(\theta^{ij}
a_i\star_2\partial_\mu\partial_j\psi\right) \Big) \, d^4x \; + \mathcal{O}(a^2)
\end{split}
\end{equation}
is consistent and gauge invariant.
The corresponding photon-fermion interaction vertex
\begin{equation}
\Gamma^{\mu}_\text{alt}=i F(q,k) \tilde q^\mu\slashed k
\end{equation}
is surprisingly simple and therefore quite attractive for loop-level computations. The vertex satisfies
\begin{equation}
q_\mu\Gamma^{\mu}_\text{alt}=(q\cdot \tilde q)i F(q,k) \slashed k=0
\,.
\end{equation}
There is, however, a second choice for $\hat\Phi$:
\begin{equation}
\begin{split}
&\hat\Phi_{}(\partial_\mu\psi)=D^\star_\mu \hat\Psi(\psi)=\partial_\mu
\hat\Psi(\psi)-i[A_\mu\stackrel{\star}{,}\hat\Psi(\psi)]
\\
=&\partial_\mu\psi-\theta^{ij}a_i\star_2\partial_j\partial_\mu\psi+\theta^{ij}f_{i\mu}\star_2\partial_j\psi
+O(a^2)\psi \,,
\end{split}
\end{equation}
based on the well-known NC QED-type covariant derivative. This second choice of SW map
differs from the first one by the gauge invariant term
$\theta^{ij}f_{i\mu}\star_2\partial_j\psi$, indicating a freedom in the choice of Seiberg-Witten map.
The second choice leads to the vertex
\begin{equation}
\Gamma^{\mu}_{}=iF(q,k)\left[\slashed k \tilde q^{\mu} + (q\theta
k)\gamma^{\mu} -\slashed q\tilde k^{\mu}\right].
\end{equation}
In general one can chose any superposition of the two SW maps $\hat\Phi_\text{alt}$ and $\hat\Phi_{}$,
but in this article we shall focus on the second choice as it is more natural from the point of view of gauge theory.
\section{Applications}
\subsection{Plasmon decay into $\bar\nu\nu$ pairs}
Our first phenomenological application of the new neutrino-photon vertex
(\ref{Feynrule}) will include a decay of transverse plasmon modes
into neutrino pairs,
which we then compare with the result obtained with perturbative methods (to first
order in $\theta$) in \cite{Schupp:2002up}.
Starting from the tree-level vector-like coupling to photons (\ref{Feynrule}),%
\footnote{A few parenthetical remarks are in order here.
In a different
model \cite{Ettefaghi:2007zz}, it is claimed that
Dirac masses for neutrinos are not consistent with NC gauge invariance
of the Yukawa terms. Our model features a vector-like
interaction, since both $\nu_L$ and $\nu_R$ are singlets under the residual
$U(1)_Q$ and
Dirac mass terms for neutrinos are
allowed. The tree-level coupling to photons is experienced by both
neutrino chiralities even above the electroweak symmetry breaking scale.
Note also that the vertex (\ref{Feynrule}) is zero
for Majorana neutrinos, unless transition electromagnetic moments are
invoked. In all applications we will work with neutrinos with definite
chiralities, i.e.\ Weyl neutrinos, as the neutrino mass can be neglected.
The interaction (\ref{Feynrule}) is generation-independent, and hereafter the
coupling constant is restored.}
standard $\gamma$-matrix techniques yield the amplitude squared for the process
$\gamma_{pl} \to
\bar\nu\nu$ summed over polarizations:
\begin{equation}
\left|M_{\rm NC}(\gamma_{pl} \to \bar\nu\nu)\right|^2
=4e^2(F(q,k))^2(q\theta k)^2\left(q^2+2m_\nu^2\right)\,. \label{M2}
\end{equation}
In the plasmon rest frame the total rate of decay into massless neutrinos involves the following phase
space integral over the outgoing neutrino momenta
\begin{equation}
\label{crosssection}
\begin{split}
&{\Gamma_{\rm NC}(\gamma_{pl} \to \bar\nu_{L \choose R} \nu_{L \choose R})}
=\frac{\alpha\,\omega_{pl}}{4\pi}
\int\limits_{0}^{\pi}\sin\vartheta
d\vartheta\int\limits_{0}^{2\pi}d\phi\sin^2 \frac{q\theta k}{2}\\&
=\frac{1}{4}\alpha \,\omega_{pl}\int\limits_{-1}^{1}dx
\bigg[1-(\cos Ax) J_0(B\sqrt{1-x^2})\bigg]\,,
\end{split}
\end{equation}
were $J_0$ is the zeroth order Bessel function of the first kind,
and\footnote{Here the dimensionless normalized matrix elements $c^{0i}$ are defined by
$\theta^{0i}=c^{0i}/\Lambda^2_{\rm NC}$, with $\sum\limits_{i=1}^3 |c^{0i}|^2=1$.}
\begin{equation}
A\equiv\frac{c_{03}\,\omega^2_{pl}}{2\Lambda^2_{\rm NC}},\quad
B\equiv\frac{\omega^2_{pl}}{2\Lambda^2_{\rm NC}}\sqrt{c^2_{01}+c^2_{02}}\,.
\label{AB}
\end{equation}
The plasma frequency $\omega_{pl}$ is defined as the frequency of plasmons
at $|\vec{q}| = 0$. In the regime where the motion of background electrons
is irrelevant, i.e.\ $q_0 >2m_e$ and $|\vec{q}| > m_e$, the dispersion
relation for transverse and longitudinal waves can be calculated
analytically, giving (see e. g. \cite{Grasso:1993bp})
\begin{equation}
\omega^2_{pl} = {\cal R}e \,\Pi_{T}(q_{0}, |\vec{q}| = 0) = \frac{e^2 T^2}{9}\,,
\label{plasdisprel}
\end{equation}
where ${\cal R}e \,\Pi_{T}$ is the transverse part of the
one-loop contribution to the photon self-energy at
finite temperature/density and $T$ is the temperature.
The integral \eqref{crosssection} can be solved analytically (see Appendix):
\begin{equation}
{\Gamma_{\rm NC}(\gamma_{pl} \to \bar\nu_{L \choose R} \nu_{L \choose R})}
=\frac{\alpha}{2} \,\omega_{pl}
\left(1-\frac{\sin\xi}{\xi}\right)~,
\label{NCrate}
\end{equation}
where $\xi={\omega_{pl}^2}/{(2\Lambda_{\rm NC}^2)}$ and
$\alpha$ is the fine structure constant.
The Standard Model (SM) neutrino-penguin-loop decay rate for transverse
plasmons (of energy $E_{\gamma}$)
into neutrinos is proportional
to $\omega_{pl}^6/E_{\gamma}$ \cite{Altherr:1993hb}.
Comparing the SM rate to our NC rate (\ref{NCrate}) for a plasmon at rest, and taking into account that
our NC photon-neutrino interaction (\ref{Feynrule})
has equal strength for both neutrino chiralities we obtain the ratio
\begin{equation}
\begin{split}
R\equiv&\frac{\sum_{\mbox{\rm\scriptsize flavors}}{\Gamma_{\rm NC}(\gamma_{pl}\to\bar\nu_L\nu_L+\bar\nu_R\nu_R)}}
{{\sum_{\mbox{\rm\scriptsize flavors}}\Gamma_{\rm SM}(\gamma_{pl} \to \bar\nu_L \nu_L)}}
\\=& \frac{3\cdot48 \pi^2\alpha^2}
{(c_{\nu_e}^2+c_{\nu_{\mu}}^2+c_{\nu_{\tau}}^2)\,G_F^2\,\omega_{pl}^4}
\left(1-\frac{\sin\xi}{\xi}\right)\,,
\label{nrate}
\end{split}
\end{equation}
see Figure~\ref{fig:RvsOmegaLambda}. For $\nu_e$, we have $c_{\rm v}=\frac{1}{2} + 2\sin^2\Theta_W$,
while for $\nu_\mu$ and $\nu_\tau$ we have $c_{\rm v}=-\frac{1}{2} + 2\sin^2\Theta_W$.
At small $\omega_{pl}$ and reasonably low NC scale $\Lambda_{\rm NC}$ the NC rate clearly dominates,
while for large $\omega_{pl}$ the Standard Model rate dominates
regardless of the value of $\Lambda_{\rm NC}$.
\begin{figure}[top]
\centerline{\includegraphics[width=8.5cm,angle=0]{RvsOmegaLambda.eps}}
\caption{The plot of the scale of noncommutativity $\Lambda_{\rm NC}$ versus
the plasmon frequency $\omega_{pl}$ with $R=1$. }
\label{fig:RvsOmegaLambda}
\end{figure}
Comparing to the previous result \cite{Schupp:2002up}, we note
an overall suppression factor $1-\frac{\sin\xi}{\xi}$ here, which,
depending on the parameters, may assume any value between 0 (for $\xi = 0$) and $1.22$ (for $\xi \approx 4.5$).
The previous result was perturbative in $\theta$, corresponding to an expansion of
the suppression factor as a power series of $\xi$
\begin{equation}
1-\frac{\sin\xi}{\xi}=\frac{1}{6}\xi^2-\frac{1}{120}\xi^4+\mathcal
O(\xi^6) \,.
\label{sinxi}
\end{equation}
The perturbative approach clearly fails, when $\xi$ is considerably larger than one,
while the new $\theta$-exact results remain valid.
Using
Eqs.~(\ref{NCrate}) and (\ref{sinxi}) for small $\xi$, we recover the old result \cite{Schupp:2002up},
thus showing consistency of the new computation.
Keeping just the first term in (\ref{sinxi}), the
ratio $R$ becomes independent of the plasma frequency $\omega_{pl}$. This
feature is reflected in the approximately constant $R=1$
contour for small $\omega_{pl}$ shown in Fig.~\ref{fig:RvsOmegaLambda} and
gives a lower bound $\Lambda_{\rm NC} \stackrel{>}{\sim} 70$ GeV,
in agreement with \cite{Schupp:2002up}, where a bound on
$\Lambda_{\rm NC}$ was derived from the requirement that NC contributions to plasmon
decay in stars should not go beyond the Standard Model predictions.
In practice, the plasmon frequency in stars,
$\omega_{pl} \simeq 10$ keV, is too low to see
the effect of the modified interaction (\ref{pf:int}). In the following, we shall explore other examples.
\subsection{Neutrino charge radius}
We make use of the full $\theta$-exact expression for the plasmon decay rate to recompute
also NC neutrino charge radii for both chiralities. Noting \cite{Minkowski:2003jg}
\begin{equation}
\Gamma(\gamma_{pl} \rightarrow {\bar\nu}_{\rm L}\nu_{\rm L})=
\frac{\alpha}{144}\frac{q^6}{E_{\gamma}}
\left|\langle r^2_{\nu}\rangle\right|^2\,
\label{Rateradii}
\end{equation}
and using
(\ref{NCrate})
we find that the $\theta$-exact NC induced neutrino charge radius is given by
\begin{equation}
\left|\langle r^2_{\nu}\rangle\right|
=\lim_{\omega_{pl} \rightarrow 0}
\frac{6\sqrt2}{\omega_{pl}^2}\sqrt{1-\frac{\sin\xi}{\xi}}\,.
\label{NCradii}
\end{equation}
The limit $\omega_{pl} \to 0$ implies, through the dispersion
relation $q^2 = \omega_{pl}^2$, the familiar $q^2 \to 0$ limit
entering the definition of the charge radius. It is interesting to note
that the expansion in the plasma frequency coincides with an
expansion in $\theta$, because the effect of the full $\theta$-exact interaction enters only through the
parameter $\xi$. Therefore, the limit $\omega_{pl} \to 0$ picks up only the
first term in (\ref{sinxi}) and that corresponds to the
first-order-in-$\theta$ result, as stated before. This lucky coincidence implies
that there are no $\theta$-exact corrections to
the first-order-in-$\theta$ charge radius that was obtained earlier
\cite{Minkowski:2003jg}:
\begin{equation}
|\langle r^2_{\nu} \rangle| = \frac{\sqrt{3}}{\Lambda_{\rm NC}^2} \;.
\label{radius}
\end{equation}
Note that the $\theta$-parameter -- when interpreted as the length scale of
the fuzziness of spacetime which
arises as a consequence of space-space uncertainty relations -- directly runs
up a charge radius for a neutrino (by giving spatial extent to a point
particle), as expected.
With (\ref{radius}) at hand, one can immediately place a constraint on
$\Lambda_{\rm NC}$ by employing a very stringent bound on $\langle r^2_{\nu_R}\rangle$ based on
SN1987A \cite{Grifols:1989vi}.
With $\langle r^2_{\nu_R}\rangle \, \stackrel{<}{\sim} \, 2 \times 10^{-33} \rm cm^2$,
and using (\ref{radius}) one obtains $\Lambda_{\rm NC} \stackrel{>}{\sim} 0.6$ TeV.
\subsection{Big Bang Nucleosynthesis (BBN)}
Over the past decades, BBN has established itself as one of the most
powerful available probes of physics beyond the Standard Model, giving
many interesting constraints on particle properties (an extensive summary is
available, for instance, in \cite{Sarkar:1996}). One uses it to parametrize the energy
density of new relativistic particles
at the time of BBN in
terms of the effective number of additional neutrino species, $\Delta
N_{\nu}$, whose determination involves both a lower limit on
the barion-to-photon ratio ($\eta \equiv n_b/n_{\gamma }$) as well as an upper
bound
on the primordial mass fraction of $^{4}$He, $Y_p $ \cite{Olive:1980bu}. The energy
density of three light right-handed (RH) neutrinos produced by plasmon decay
during BBN is equivalent to the
effective number $\Delta N_{\nu}$ of additional doublet neutrinos
\begin{equation}
\Delta N_{\nu} = 3
\left (\frac{T_{\nu_{R}}}{T_{\nu_{L}}} \right )^4 \;,
\label{Ntemp}
\end{equation}
where $T_{\nu_{L}}$
is the temperature of the SM neutrinos, being the same as
that of photons down to $T \sim 1$ MeV. A better limit on $\Delta
N_{\nu}$ leads to a smaller value of $T_{\nu_R}$ and consequently a higher
decoupling temperature of the RH neutrinos. For $\Delta N_{\nu} = 1$, one finds
$T_{dec} > T_C $, where $T_C
$ is the critical temperature for the deconfinement restoration phase
transition, $T_C \sim 200$ MeV. If $\Delta N_{\nu} \simeq 0.2$, then
$T_{dec}$ would be
close to a critical temperature of the electroweak phase transition,
$T_{dec} \lesssim \; 300 $ GeV. Unfortunately, with the WMAP value
for $\eta $ \cite{Spergel:2003cb}, $Y_p $ was predicted to increase
\cite{Cyburt:2004cq}, having a
tendency to loosen the tight bounds on $\Delta N_{\nu}$ that existed before.
The RH neutrino is commonly considered to decouple at the temperature
$T_{dec}$ when the condition
\begin{equation}
{\Gamma(\gamma_{pl} \to \bar\nu_{R} \nu_{R})}\simeq H(T_{dec})
\label{decouple}
\end{equation}
is satisfied. The plasma frequency in this case is given by
\begin{equation}
\omega_{pl}= \frac{eT_{dec}}{3}\,g^{ch}_{*}\,,
\label{disprel}
\end{equation}
while the Hubble expansion rate satisfies
\begin{equation}
H(T) \simeq 1.66 \,g_{*} \frac{T_{dec}^2}{M_{Pl}}
\label{HT}\;,
\end{equation}
where $g_{*}$ and $g^{ch}_{*}$ are the degrees of freedom specifying the
entropy of the interacting species for all and charged species,
respectively; $M_{Pl}$ is the Planck mass.
Computing the decoupling temperature $T_{dec}$ based on the assumption that
the decay rate (\ref{decouple}) is solely due to noncommutative effects and comparing
with lower bounds on $T_{dec}$ that can be inferred from observational data, we can determine
lower bounds on the scale of noncommutativity $\Lambda_{\rm NC}$.
Proceeding in this spirit, one finds that Big Bang nucleosynthesis provides the following relation between the decoupling
temperature $T_{dec}$ and the noncommutative scale $\Lambda_{\rm NC}$,
assuming that (\ref{decouple}) is fully due to noncommutative contributions:
\begin{equation}
T_{dec} \simeq \frac{M_{Pl}e^3 g^{ch}_*}{39.84\pi
g_*}\left(1-\frac{\sin\xi}{\xi}\right),\quad \xi=\frac{e^2
(g^{ch}_*)^2 T_{dec}^2}{18\Lambda_{\rm NC}^2}\,.
\label{Tdecxi}
\end{equation}
Let us consider the pre-factor
\begin{equation}
\frac{M_{Pl}e^3 g^{ch}_*}{39.84\pi g_*}
=\frac{\pi^{\frac{1}{2}}\alpha^{\frac{3}{2}}g^{ch}_*}{4.98g_*}M_{Pl},
\label{4.98}
\end{equation}
and taking into account $\alpha\simeq 137^{-1}$, $g^{ch}_*/g_*\simeq 1$, we
have
\begin{equation}
\frac{\pi^{\frac{1}{2}}\alpha^{\frac{3}{2}}g^{ch}_*}{4.98g_*}M_{Pl}\simeq
2.22\times 10^{-4}M_{Pl}.
\label{2.22}
\end{equation}
Since this factor is amplified by the Planck mass,
we can test the full interaction (\ref{pf:int}) only for $\Delta N_{\nu} \ll 1$.
Unfortunately, such a precision in observational data is not expected
to be reached anytime soon. Thus, to
account for decoupling temperatures associated with cosmological phase
transitions ($200\;\rm
MeV/300$ GeV), we have to require
$\left(1-\frac{\sin\xi}{\xi}\right)\ll 1$. This only occurs when
$\xi\to 0$, so that within this regime we can use the leading order term
in the Taylor expansion in $\xi$. This gives
\begin{equation}
\begin{split}
T_{dec}\simeq&
\frac{\pi^{\frac{1}{2}}\alpha^{\frac{3}{2}}g^{ch}_*M_{Pl}}{4.98
g_*}\cdot\frac{\xi^2}{6}
\\=&\frac{\pi^{\frac{1}{2}}\alpha^{\frac{3}{2}}g^{ch}_*M_{Pl}}{4.98
g_*}\cdot\frac{4\pi\alpha
(g^{ch}_*)^2}{108}\cdot\left(\frac{T_{dec}}{\Lambda_{\rm NC}}\right)^4\,,
\end{split}
\label{T_dec}
\end{equation}
so that
\begin{equation}
\Lambda_{\rm
NC}\simeq\left(\frac{\pi^{\frac{3}{2}}\alpha^{\frac{5}{2}}
(g^{ch}_*)^3M_{Pl}T_{dec}^3}{4.98\cdot 27 g_*}\right)^{\frac{1}{4}}\,.
\label{LaNC}
\end{equation}
Now setting further $g_*\sim
g^{ch}_*\sim 100$, $M_{Pl}=1.22\times 10^{19}$ GeV and
$T_{dec}>200$ MeV (quark-hadron phase transition), a lower bound on $\Lambda_{\rm
NC}$ can be obtained as
\begin{equation}
\Lambda_{\rm NC}>3.68\; \rm TeV.
\label{8.35}
\end{equation}
For $T_{dec}>300$ GeV (eletro\-weak phase transition), we have
\begin{equation}
\Lambda_{\rm NC}>887 \; \rm TeV,
\label{10^3}
\end{equation}
confirming the previous results obtained by making use of scattering processes
of $\nu_R$s \cite{Horvat:2009cm}.
For the sake of demonstration, we have also studied (\ref{Tdecxi}) numerically, to
investigate how the plasmon rate obtained in the full theory affects determination
of $\Lambda_{\rm NC}$ when $T_{dec} \simeq 10^{-4} M_{Pl}$. The relation
(\ref{Tdecxi}) is depicted in Fig. \ref{fig:LambdaVsTc}, in which the
approximate result (\ref{LaNC}) is also superimposed for comparison.
\begin{figure}[top]
\begin{center}
\includegraphics[width=8.5cm,angle=0]{LambdaVsTdec600.eps}
\end{center}
\caption{The plot of the scale of noncommutativity $\Lambda_{\rm NC}$ versus
decoupling temperature $T_{dec}$. The dashed/full curve corresponds to the
perturbative/full solution, as given by
Eqs.~(\protect\ref{LaNC}) and (\protect\ref{Tdecxi}), respectively. In both curves we
set, for illustration purposes, $g_{*} = g_{*}^{ch} = 100$, and the logarithmic scaling of the fine
structure constant with temperature is ignored. The full curve reveals
$T_{dec}^{max} = 2.7 \times 10^{-4} M_{Pl}$ and $\Lambda_{\rm NC}^{max} =
9.4 \times 10^{-4} M_{Pl}$.}
\label{fig:LambdaVsTc}
\end{figure}
Looking at the asymptotic behavior of the approximate (\ref{LaNC}) and exact (\ref{Tdecxi}) relations
between $\Lambda_{\rm NC}$ and $T_{dec}$, one notes
that Fig.~2 reveals quite a different behavior for them: While
the solution of (\ref{Tdecxi}), obtained by employing the leading order term in
$\xi$, shows no restriction on $T_{dec}$ all the way up till the beginning
of the radiation era, the solution of (\ref{LaNC}) reveals a maximal decoupling
temperature in the said epoch. (Of course, if inflation occurred
well below $T_{dec}^{max} \simeq 3 \times 10^{-4} M_{Pl}$, then
any distinction between the two solutions would practically disappear.) This
feature is accompanied by a maximal scale of noncommutativity, above which
the RH neutrinos can no longer retain thermal contact with the rest of the
universe. This is in contrast with the exact relation (\ref{Tdecxi}), where thermal
equilibrium is maintained for much larger $\Lambda_{\rm NC}$'s (if inflation
occurred well above $T_{dec}^{max}$), and stops when $T_{dec}$
hits the reheating temperature. This means that the effect of the full
interaction (\ref{Feynrule}) is to bring about the maximal upper limit $\Lambda_{\rm
NC}^{max}
\simeq 9 \times 10^{-4} M_{Pl}$, occurring at a decoupling temperature
slightly below $T_{dec}^{max}$, which can be extracted by our
method. Note that for decoupling temperatures lying in a region where the
oscillatory patterns inherent in (\ref{LaNC}) becomes manifest, it may seem
troublesome to infer a bound on the scale of noncommutativity
since several (many) $\Lambda_{\rm NC}$'s correspond to the same $T_{dec}$. In
these cases one chooses the highest $\Lambda_{\rm NC}$, otherwise
$\Lambda_{\rm NC}$'s
obtained with a much smaller decoupling temperature (where the oscillatory
term is shut down) would give much better lower limits. Again, this
characteristic is missing in the perturbative solution, where
better and better limits
on the NC scale are always accompanied with progressively increasing
decoupling temperatures.
Note that these ranges of $T_{dec}$
are of course tremendously above the bounds that can be inferred from
current observational data.
\subsection{Ultrahigh Energy (UHE) Cosmic Rays}
The non-observation of UHE neutrino induced events in neutrino observatories
implies a strong model-independent constraint on the inelastic
neutrino-nucleon cross section \cite{Anchordoqui:2004ma}, which consequently gives a
constraint on the scale of noncommutativity for a NC gauge-field theory in which
neutrinos couple directly to photons \cite{Horvat:2010sr}. It was observed
\cite{Horvat:2010sr} that at energies as high as $10^{11}$ GeV, the usual expansion in $\theta$
is no longer meaningful. In order to fix the breakdown in
the perturbative expansion, a resummation in the neutrino-photon
vertex was undertaken \cite{Horvat:2010sr}. Although
devoid of a firm theoretical background and with only the zeroth order in
the Seiberg-Witten map employed, this {\sl ad hoc} approach did produce the correct $\sin$
term that we have now obtained, redoing the computation using the $\theta$-exact vertex (\ref{Feynrule}).
Curiously enough, (\ref{Feynrule}) can be put in a much simpler form
if the NC vertex connects external (on-shell) neutrino lines. Indeed, with
the aid of the Dirac equation for free fields, the first term in (\ref{Feynrule})
vanishes if one line is on-shell, while the third term in (\ref{Feynrule})
vanishes if both lines are on-shell. Thus, for tree-level processes with
no internal neutrino lines, (\ref{Feynrule}) reduces to
\begin{equation}
\Gamma^{\mu} = 2 i\,e \,\gamma^{\mu} \;{\rm sin} \left (\frac{q \theta k}{2} \right )\,.
\label{adhocFR}
\end{equation}
This exactly coincides with the result \cite{Horvat:2010sr} obtained with
the {\sl ad-hoc} method. This way, the powerful bounds on $\Lambda_{\rm NC}$
obtained there, in the range 200-900 TeV (depending on a model for the
cosmogenic neutrino flux), get further credence as far as the
underlying theoretical background is concerned. Note, though, that
(\ref{adhocFR}) should not be used to calculate
tree-level processes with internal neutrino lines, nor in calculations
involving loops. Hence, if one is, for instance, to
study the UV/IR mixing in the neutrino sector,
then the complete expression as given by (\ref{Feynrule}) should be used.
\section{Conclusion}
In summary, we showed that the tree-level tri-particle
decay $\gamma_{pl}\to\nu\bar\nu$ in the covariant noncommutative quantum gauge
theory based on Seiberg-Witten maps can be computed without an
expansion over the noncommutative parameter $\theta$.
As an application, we focus on plasmon decay into neutrinos,
reconsidering previous computations that were done with less sophisticated tools and deriving new
bounds on the scale of noncommutativity.
Comparing to
previous results, the total decay rate is modified by a factor
which remains finite throughout all energy scales. Thus the new results
behave much better than the $\theta$-expansion method when ultra
high energy processes are considered. We expect that similar
control on the high energy behavior can be extended to $\theta$-exact perturbation theory involving
more than three external fields in the near future. This would provide a considerably
improved theoretical basis for research work in the field of noncommutative
particle phenomenology.
\noindent
\acknowledgments
J.T. would like to acknowledege support of W. Hollik, and MPI Munich for hospitality.
The work of R.H., D.K and J.T. are supported by
the Croatian Ministry of Science, Education and Sports
under Contracts Nos. 0098-0982930-2872 and 0098-0982930-2900, respectively.
The work of J.Y. was supported by the Croatian NSF and the IRB Zagreb,
and by the German Research Foundation (Deutsche
Forschungsgemeinschaft (DFG)) through the Institutional Strategy of the
University of G\"ottingen.
|
\section{Introduction}
In 1950, G. Giuga conjectured (Giuga's conjecture, see \cite{GIU}) that if an integer $n$ satisfies $\displaystyle{\sum_{j=1}^{n-1} j^{n-1} \equiv -1}$ (mod $n$), then $n$ must be a prime. Moreover, Giuga proved that $n$ is a counterexample to his conjecture if and only if for each prime divisor $p$ of $n$, $(p-1)|(n/p-1)$ and $p|(n/p-1)$. Using this characterization, he proved computationally that any counterexample has at least 1000 digits. Equipped with more computing power, E. Bedocchi (see \cite{BEDO}) later raised this bound to 1700 digits. Improving their method, D. Borwein, J. M. Borwein, P. B. Borwein and R. Girgensohn (see \cite{BOR}) determined that any counterexample contains at least 3459 distinct primes and so has at least 13887 digits.
The second of the above conditions ($p|(n/p-1)$) motivated the following definition of Giuga Numbers, introduced by Borwein et al. in the paper \cite{BOR}, where these numbers were studied in detail.
\begin{defi}
A Giuga Number is a composite number $n$ such that $p|(n/p-1)$ for every $p$, prime divisor of $n$.
\end{defi}
There are several characterizations of Giuga Numbers. The most important being the following.
\begin{prop}
Let $n$ be a composite integer. Then, the following are equivalent:
\begin{itemize}
\item [i)] $n$ is a Giuga Number.
\item [ii)] $\displaystyle{\sum_{p|n} \frac{1}{p}-\prod_{p|n} \frac{1}{p} \in \mathbb{N}}$ (see \cite{GIU}).
\item[iii)] $\displaystyle{\sum_{j=1}^{n-1} j^{\phi(n)} \equiv -1}$ (mod $n$), where $\phi$ is Euler's totient function (see \cite{BOR}).
\item [iv)] $nB_{\phi(n)} \equiv -1$ (mod $n$), where $B$ is a Bernoulli number (see \cite{AGO}).
\end{itemize}
\end{prop}
It has been conjectured by Paolo P. Lava (see \cite{LAVA}) that Giuga Numbers are the solutions of the differential equation $n'=n+1$, with $n'$ being the arithmetic derivative of $n$.
Up to date only thirteen Giuga Numbers are known (see A007850 in the \emph{On-Line Encyclopedia of Integer Sequences}):
\begin{itemize}
\item With 3 factors:
$$\textbf{30} = 2 \cdot 3 \cdot 5.$$
\item With 4 factors:
\begin{align*}
\textbf{858}&=2 \cdot 3 \cdot 11 \cdot 13,\\
\textbf{1722}&=2 \cdot 3 \cdot 7 \cdot 41.
\end{align*}
\item With 5 factors:
$$\textbf{66198} = 2 \cdot 3 \cdot 11 \cdot 17 \cdot 59.$$
\item With 6 factors:
\begin{align*}
\textbf{2214408306}&=2 \cdot 3 \cdot 11 \cdot 23 \cdot 31 \cdot 47057,\\
\textbf{24423128562}&=2 \cdot 3 \cdot 7 \cdot 43 \cdot 3041 \cdot 4447.
\end{align*}
\item With 7 factors:
\begin{align*}
\textbf{432749205173838}&=2 \cdot 3 \cdot 7 \cdot 59 \cdot 163 \cdot 1381 \cdot 775807,\\
\textbf{14737133470010574}&=2 \cdot 3 \cdot 7 \cdot 71 \cdot 103 \cdot 67213 \cdot 713863,\\
\textbf{550843391309130318}&=2 \cdot 3 \cdot 7 \cdot 71 \cdot 103 \cdot 61559 \cdot 29133437.
\end{align*}
\item With 8 factors:
\begin{align*}
\textbf{244197000982499715087866346}=\ & 2 \cdot 3 \cdot 11 \cdot 23 \cdot 31 \cdot 47137 \cdot 28282147 \cdot \\ &3892535183,\\
\textbf{554079914617070801288578559178}=\ & 2 \cdot 3 \cdot 11 \cdot 23 \cdot 31 \cdot 47059 \cdot 2259696349 \cdot\\ & 110725121051,\\
\textbf{1910667181420507984555759916338506}=\ & 2 \cdot 3 \cdot 7 \cdot 43\cdots 1831 \cdot 138683 \cdot 2861051 \cdot \\ &1456230512169437.
\end{align*}
\end{itemize}
There are no other Giuga Numbers with less than 8 prime factors. There is another known Giuga Number (found by Frederick Schneider in 2006) which has 10 prime factors, but it is not known if there is any Giuga Number between this and the previous ones. This biggest known Giuga Number is:
\begin{align*}
&\textbf{4200017949707747062038711509670656632404195753751630609228764416}\\ &\textbf{142557211582098432545190323474818}= 2 \cdot 3 \cdot 11 \cdot 23 \cdot 47059 \cdot 2217342227 \cdot\\ &1729101023519 \cdot 8491659218261819498490029296021 \cdot\\
&658254480569119734123541298976556403.
\end{align*}
Observe that all known Giuga Numbers are even. If an odd Giuga Number exists, it must be the product of at least 14 primes. It is not even known if there are infinitely many Giuga Numbers.
The most important result about Giuga's conjectureit is due to Giuga himself (see \cite{GIU} or \cite{BOR}).
\begin{prop}
A composite integer $n$ is a counterexample to Giuga's conjecture if and only if it is both a Carmichael Number and a Giuga Number.
\end{prop}
On the other hand, Luca, Pomerance and Shparlinski (see \cite{LUCA}) have established the following bound for the counter function of the counterexamples to Giuga's conjecture thus improving a result by Tipu in \cite{TIPU}:
$$|\{n<X | n \textrm{ is countarexample of Giuga's conjecture }\}| \ll \frac{X^{\frac{1}{2}}}{(\log(X))^2}.$$
Finally we will mention that Bernd C. Kellner \cite{KELL} has stablished that Giuga's conjecture is equivalent to the following conjecture by Agoh.
\begin{con}[Takashi Agoh, 1990]
Let $B_k$ denote the k-th Bernoulli number. Then, $n$ is a prime if and only if $nB_n \equiv -1 (\textrm{mod. n})$
\end{con}
\section{New characterizations for Giuga Numbers}
The following result will be the core of the paper. It stablishes that in Proposition 1 iii) we can replace Euler's totient function $\phi(n)$ by Carmichael's function $\lambda(n)$ or by any multipli of $\phi(n)$ or $\lambda(n)$.
\begin{prop}
For every natural numbers $A$, $B$ and $N$ we have that:
$$\sum_{j=1}^{N-1} j^{A \lambda(N)} \equiv \sum_{j=1}^{N-1} j^{B \phi(N)}\ \textrm{(mod $N$)}.$$
\end{prop}
\begin{proof}
Put $N=2^ap_1^{r_1}\cdots p_s^{r_s}$ with $p_i$ distinct odd primes. Choose $i\in \{1,\dots, s\}$. We have that:
$$\sum_{j=1}^{N-1} j^{A\lambda(N)} \equiv \frac{N}{p_i^{r_i}}\sum_{j=1}^{p_i^{r_i}-1} j^{A\lambda(N)}\ \textrm{(mod $p_i^{r_i}$)}.$$
$$\sum_{j=1}^{N-1} j^{B\phi(N)} \equiv \frac{N}{p_i^{r_i}}\sum_{j=1}^{p_i^{r_i}-1} j^{B\phi(N)}\ \textrm{(mod $p_i^{r_i}$)}.$$
Now, since both $A\lambda(N),\ B\phi(N)\geq r_i$, we get:
$$\sum_{j=1}^{p_i^{r_i}-1} j^{A\lambda(N)}=\sum_{\substack{1\leq j\leq p_i^{r_i}-1\\ (p_i,j)=1}} j^{A\lambda(N)}+\sum_{\substack{1\leq j\leq p_i^{r_i}-1\\ p_i|j}} j^{A\lambda(N)}\equiv \phi(p_i^{r_i})+0\ \textrm{(mod ($p_i^{r_i}$)}.$$
$$\sum_{j=1}^{p_i^{r_i}-1} j^{B\phi(N)}=\sum_{\substack{1\leq j\leq p_i^{r_i}-1\\ (p_i,j)=1}} j^{B\phi(N)}+\sum_{\substack{1\leq j\leq p_i^{r_i}-1\\ p_i|j}} j^{A\lambda(N)}\equiv \phi(p_i^{r_i})+0\ \textrm{(mod ($p_i^{r_i}$)}.$$
Consequently:
$$\sum_{j=1}^{N-1} j^{A\lambda(N)}\equiv\sum_{j=1}^{N-1} j^{B\phi(N)}\ \textrm{(mod $p_i^{r_i}$)}\ \textrm{for every $i=1,\dots,s$}.$$
Clearly if $N$ is odd the proof is complete. If $n$ is even we have that:
$$\sum_{j=1}^{N-1}j^{A\lambda(N)}\equiv\frac{N}{2^a}\sum_{j=1}^{2^a-1} j^{A\lambda(N)}\equiv\frac{N}{2^a}\left(\sum_{\substack{1\leq j\leq 2^a-1\\ \textrm{$j$ even}}} j^{A\lambda(N)} + 2^{a-1}\right)\ \textrm{(mod $2^a$)}.$$
$$\sum_{j=1}^{N-1}\equiv\frac{N}{2^a}\left(\sum_{\substack{1\leq j\leq 2^a-1\\ \textrm{$j$ even}}} j^{B\phi(N)} + 2^{a-1}\right)\ \textrm{(mod $2^a$)}.$$
Now, if $a=1,2$ or $3$ it can be easily verified that:
$$\sum_{\substack{1\leq j\leq 2^a-1\\ \textrm{$j$ even}}} j^{A\lambda(N)}\equiv\sum_{\substack{1\leq j\leq 2^a-1\\ \textrm{$j$ even}}} j^{B\phi(N)}\ \textrm{(mod $2^a$)}.$$
On the other hand, if $a\geq 4$ we have that $\phi(N)\geq\lambda(N)\geq a$ and, consequently $j^{A\lambda(N)}\equiv j^{B\phi(N)}\equiv 0$ (mod $2^a$) for every $1\leq j\leq 2^{a-1}$ even. Thus:
$$\sum_{j=1}^{N-1} j^{A\lambda(N)}\equiv\sum_{j=1}^{N-1} j^{B\phi(N)}\ \textrm{(mod $2^a$)}$$
and the result follows.
\end{proof}
The proposition above allows us to introduce some new characterizations of Giuga Numbers. Recall that a composite integer $n$ is said to be a Giuga Number if and only if $\displaystyle{\sum_{j=1}^{n-1} j^{\phi(n)}\equiv -1}$ (mod $n$).
\begin{cor}
Let $n$ be any composite integer. Then the following are equivalent:
\begin{itemize}
\item[i)] $n$ is a Giuga Number.
\item[ii)] $\displaystyle{\sum_{j=1}^{n-1} j^{\lambda(n)}} \equiv -1$ (mod $n$).
\item[iii)] For every positive integer $K$, $\displaystyle{\sum_{j=1}^{n-1} j^{K\lambda(n)}} \equiv -1$ (mod $n$).
\item [iv)] There exists a positive integer $K$ such that $\displaystyle{\sum_{j=1}^{n-1} j^{K\lambda(n)}} \equiv -1$ (mod $n$).
\item[v)] There exists a positive integer $K$ such that $\displaystyle{\sum_{j=1}^{n-1} j^{K\phi(n)} \equiv -1}$ (mod $n$).
\item[vi)] For every positive integer $K$, $\displaystyle{\sum_{j=1}^{n-1} j^{K\phi(n)} \equiv -1}$ (mod $n$).
\end{itemize}
\end{cor}
\begin{proof}
It follows readily from the previous proposition and from Proposition 1 iii).
\end{proof}
Proposition 1 also allows us to show in a different and novel way that an integer which is both a Carmichael Number and a Giuga Numberis a counterexample to Giuga's conjecture.
\begin{cor}
If an integer $n$ is both a Carmichael Number and a Giuga Number, then:
$$\sum_{j=1}^{n-1} j^{n-1} \equiv -1\ \textrm{(mod $n$)}.$$
\end{cor}
\begin{proof}
If $n$ is a Carmichael Number, then $\lambda(n) | (n-1)$. If we put $k=\frac{n-1}{\lambda(n)}$, then we have:
$$S:=\sum_{j=1}^{n-1} j^{n-1}=\sum_{j=1}^{n-1} j^{k \lambda(n)}.$$
If, in addition, $n$ is a Giuga Number we can apply Proposition 3 with $A=k$ and $B=1$ to get $S\equiv -1$ (mod $n$) as claimed.
\end{proof}
Although the previous result is not new, it is interesting to observe that the proof avoids the use of Korselt's criterion which is replaced by Carmichael's criterion, i.e., $\lambda(n)|(n-1)$.
The following result, which is also a consequence of Proposition 3, will allow us to generalize Giuga's ideas by considering the congruence $\displaystyle{\sum_{j=1}^{N-1} j^{k(n-1)} \equiv -1}$ (mod $n$) for each positive integer $k$.
\begin{cor}
If an integer $n$ is a counterexample to Giuga's conjecture, then:
$$\sum_{j=1}^{n-1} j^{k(n-1)} \equiv -1\ \textrm{(mod $n$) for every positive integer $k$}.$$
\end{cor}
\begin{proof}
If $n$ is a counterexample to Giuga's conjecture, then it is both a Carmichael and a Giuga Number. Being a Carmichael Number, we have that $\lambda(n) | (n-1)$ so if $\frac{k(n-1)}{\lambda(n)}=\widetilde{k}\in\mathbb{N}$ we get:
$$S:=\sum_{j=1}^{n-1} j^{k(n-1)}=\sum_{j=1}^{n-1} j^{k \lambda(n)\frac{(n-1)}{\lambda(n)}}=\sum_{j=1}^{n-1} j^{\widetilde{k} \lambda(n)},$$
and, since $n$ is a Giuga Number it is enough to apply Corollary 1 and Proposition 3 to get $S\equiv -1$ (mod $n$).
\end{proof}
\section{Generalizing Giuga's conjecture }
In this section we generalize Giuga's ideas in the following way: Do there exist integers $k$ and $n$ such that the congruence $\displaystyle{\sum_{j=1}^{n-1} j^{k(n-1)} \equiv -1}$ (mod $n$) is satisfied by some composite integer $n$?
In what follows we will denote the set of Giuga Numbers by $\mathcal{G}$ and the set of Carmichael Numbers by $\mathcal{C}$. Moreover, for every positive integer $k$ let us define the following sets:
$$\mathcal{G}_k:=\left\{n\in\mathbb{N} \left|\textrm{ $n$ is composite, } \displaystyle{\sum_{j=1}^{n-1} j^{k(n-1)} \equiv -1}\right.\ \textrm{(mod $n$)}\right\},$$
$$\mathcal{K}_n:=\left\{k \in \mathbb{N} \left|\ \sum_{j=1}^{n-1} j^{k(n-1)} \equiv -1\ \right. \textrm{(mod $n$)}\ \right\}.$$
\begin{rem}
With the previous notation, Giuga's conjecture is equivalent to the statement $\mathcal{G}_1=\emptyset$. Also observe that, for every positive integer $k$:
$$ k \in \mathcal{K}_n\ \textrm{if and only if}\ n \in \mathcal{G}_k.$$
\end{rem}
\begin{teor}
Let $n$ be a composite integer. Then the following are equivalent:
\begin{itemize}
\item[i)] $ n \in \mathcal{G}_k $.
\item[ii)] $n \in \mathcal{G}$ and $\displaystyle{\frac{\lambda(n)}{\gcd(\lambda(n),n-1)}}$ divides $k$.
\end{itemize}
\end{teor}
\begin{proof}
Assume that $n\in\mathcal{G}$ and that $\displaystyle{\frac{\lambda(n)}{\gcd(\lambda(n),n-1)}}$ divides $k$. Then we have:
$$\sum_{j=1}^{n-1} j^{k(n-1)} = \sum_{j=1}^{n-1} j^{\frac{\lambda(n)}{gcd(\lambda(n),n-1)} k' (n-1)}=\sum_{j=1}^{n-1} j^{k''\lambda(n)}\equiv -1$$
due to Corollary 1.
Conversely, assume that $n\in\mathcal{G}_k$. This means that $\displaystyle{\sum_{j=1}^{n-1}j^{k(n-1)}\equiv -1}$ (mod $n$). As a consequence (see \cite{WONG}[Theorem 2.3]) we have that $p-1$ divides $k\left(n/p-1\right)$ and that $p$ divides $n/p-1$ for every $p$, prime divisor of $n$. and, moreover, $n$ is square-free. Since $n$ is square-free we have that $\lambda(n)=\textrm{lcm}\{p-1\ |\ \textrm{$p$ odd prime dividing $n$}\}$. Thus, $\lambda(n)$ divides $k(n-1)$ and, consequently, $\displaystyle{\frac{\lambda(n)}{\gcd(\lambda(n),n-1))}}$ divides $k$. It is enough to apply Proposition 3 with $B=1$ and $A=\frac{n-1}{\gcd(\lambda(n),n-1)}$ to finish the proof.
\end{proof}
If we put $k=1$ in the theorem above, we obtain again Corollary 2 and also its reciprocal thus completing the characterization of counterexamples to Giuga's conjecture without the use of Korselt's criterion.
\begin{cor}
An integer $n$ is a counterexample to Giuga's conjecture if and only if it is both a Carmichael and a Giuga Number. In other words: $$\mathcal{G}_1=\mathcal{G}\cap \mathcal{C}.$$
\end{cor}
\begin{proof}
Let $n \in \mathcal{G}_1$. Then, by the previous theorem $n \in \mathcal{G} $ and $\gcd(\lambda(n),n-1)=\lambda(n)$; i.e., $n \in \mathcal{C}$.
Now, let $n \in \mathcal{C}\cap\mathcal{G}$. Then $\frac{\lambda(n)}{\gcd(\lambda(n),n-1)}=1$ and also the previous theorem implies that $n \in \mathcal{G}_1$. This completes the proof.
\end{proof}
By Theorem 1 we can find values of $k$ such that $\mathcal{G}_k$ is non-empty. To do so, we evaluate $\frac{\lambda(n)}{gcd(\lambda(n),n-1)}$ for every known Giuga Number. Thus, we will have thirteen values of $k$ for which $\mathcal{G}_{tk}$ is known to be nonempty for any $t$. In the case $t=1$, they are:
$$\mathcal{G}_4=\{30, ...\};$$
$$\mathcal{G}_{60}=\{30, 858, ...\};$$
$$\mathcal{G}_{120}=\{30, 858, 1772, ...\};$$
$$\mathcal{G}_{2320}=\{30, 66198,...\};$$
$$\mathcal{G}_{1552848}=\{30,2214408306,...\};$$
$$\mathcal{G}_{10080}=\{30,858, 24423128562,...\};$$
$$\mathcal{G}_{139714902540}=\{30,858,432749205173838,...\};$$
$$\mathcal{G}_{93294624780}=\{30,858,14737133470010574,...\};$$
$$\mathcal{G}_{228657996794220}=\{30,858,550843391309130318,...\};$$
$$\mathcal{G}_{4756736241732916394976}=\{30,244197000982499715087866346,...\};$$
$$\mathcal{G}_{20024071474861042488900}=\{30,554079914617070801288578559178,...\};$$
$$\mathcal{G}_{2176937111336664570375832140}=\{30,858,1910667181420507984555759916338506,...\};$$
$$\mathcal{G}_{15366743578393906356665002406454800354974137359272445859047945613961394951904884493965220}$$ $$=\{30,858,42000179497077470620387115096706566324041957537516306092287644$$ $$\hspace{-5cm} 16142557211582098432545190323474818,...\}.$$
Given any positive integer $n$, Theorem 1 gives a complete description of the set $\mathcal{K}_n$ as can be seen in the following corollary.
\begin{cor}
Let $n$ be any positive integer. Then:
$$\mathcal{K}_n =\begin{cases} \left\{\frac{t\lambda(n)}{gcd(\lambda(n),n-1)}\ |\ t\in \mathbb{N} \right\}, & \textrm{if $n \in \mathcal{G} $;}\\ \emptyset, & \textrm{otherwise.}\\ \end{cases}$$
\end{cor}
\begin{rem}
Observe that $\mathcal{K}_n=\mathbb{N}$ if and only if $n$ is a counterexample to Giuga's conjecture.
\end{rem}
\section{Reflecting about Giuga's conjecture }
In recent work by W. D. Banks, C. W. Nevans, and C. Pomerance \cite{POM}, the following bounds were given:
\begin{teor} For any fixed $ \varepsilon> 0$, $\beta=0.3322408$ and all sufficiently large X, we have
$$| \{n<X | n \in \mathcal{C} \} | \geq X^{\beta-\varepsilon} \textrm{ (due to G. Harman \cite{HAR} )}$$
$$| \{n<X | n \in \mathcal{C} \backslash \mathcal{G}_1 \} | \geq X^{\beta-\varepsilon} $$
\end{teor}
The authors of the aforementioned paper consider the above bounds to be consistent with Giuga's conjecture. We believe, however, that the same consideration could be made with respect to conjectures that are actually false, such as $\mathcal{G}_4=\emptyset$ or $\mathcal{G}_{1552848}=\emptyset$. What is clearly consistent is the assumption that the elements of $\mathcal{G}_{k}$, regardless of the value of $k$, are extremely ``scarce''.
Furthermore, the authors of the present paper, in view of the generalization presented here, are convinced that Giuga's conjecture is not based on any sound logical-mathematical consideration and that its strength rests on the extreme rarity of Giuga Numbers, combined with the computational difficulties that the search for new numbers belonging to this family entail, as well as with the null asymptotic density of Carmichael numbers. In fact, if we may be forgiven the joke, we might conjecture -without any fear of our conjecture being refuted in many years- that $\mathcal{G}_2=\mathcal{G}_3=\mathcal{G}_5=\emptyset$ or, to be even more daring, that $\mathcal{G}_p=\emptyset$ for all prime $p$. Of course, Giuga's conjecture has the honour of being the strongest of all of these conjectures. In fact, in virtue of Corollary 3, should it be refuted, all the others would fall with it.
|
\section{Introduction} \label{sec:intro}
The most recent observational evidence points towards a model of structure and galaxy formation that is hierarchical in nature: small fluctuations in the matter density field grow via gravity, with their dynamics being governed in detail by dark matter and dark energy. Baryons trace the dark matter and, in regions of sufficient gravitational depth, they accumulate and form stars and galaxies \citep{WhiteRees78}. As the matter field continues to evolve dynamically - largely oblivious to this process - the stellar content of galaxies grows via a combination of two modes: forming new stars from cold gas, and merging with other galaxies.
LRGs are extensively used as cosmological probes (e.g. \citealt{ReidEtAl10, PercivalEtAl10}), and are also interesting from a galaxy evolution perspective, as they dominate the galaxy mass function at the massive end. For these reasons, the stellar and dynamical evolution of LRGs and early-type galaxies (ETGs) has been studied extensively (see \citealt{TojeiroEtAl10, TojeiroEtAl11} - henceforth T11 - for a summary and list of references). Studies that address one of these growing modes, however, traditionally assume a model for the other. In this paper we propose and apply a new methodology that solves for these two modes independently. In brief, we use the fossil record of local galaxies to predict number and luminosity densities at past redshifts. The differences between the predicted and the observed quantities, under some simple assumptions, can be interpreted as a merger history.
The idea of inferring the properties of galaxies at a different cosmic time from the one they are observed at, and comparing them to the in-situ properties of galaxies at those cosmic times has been proposed before. Most notably \cite{DroryAlvarez08} use the galaxy stellar mass function (GSMF) of galaxies in the FORS Deep Field \citep{DroryEtAl05} and estimates of the instantaneous star-formation rate as a function of observed stellar mass to separate the evolution of the GSMF in terms of its merging and star formation components. Our approach differs from theirs in terms of the observables (we focus on number and luminosity densities, as opposed to the GSMF), but mainly in how we effectively link the galaxies at different cosmic times. We rely on the fossil record of local galaxies to reconstruct their past stellar build-up, which we get from VESPA \citep{TojeiroEtAl07} in a non-parametric way. \cite{DroryAlvarez08} use parametric functions for the evolution of the GSMF and for the instantaneous star-formation rate, which they integrate over past cosmic time to predict a mass build up due to star-formation only. In other words, we use present-day information to predict the past Universe, whilst they use past information to predict the present-day Universe. Our reconstruction of stellar mass build-up is less parametric, but arguably more dependent on the underlying stellar population synthesis (SPS) models. The two results are not comparable due to the sample of galaxies studied, but the approaches are similar in terms of philosophy. An approach more similar to ours was very recently presented in \cite{EskewEtAl11}, where the authors use published estimates of the past star formation histories of three local group galaxies (obtained using resolved stellar populations) to infer where these galaxies would sit in popular diagnostic plots at previous times in their cosmic history. The work we present this paper is similar in terms of concept, but is vastly different in terms of the size and type of galaxies used, as well as overall goal.
This paper is organised as follows: in Section \ref{sec:data} we introduce the data set we use, in Section \ref{sec:method} we introduce our methodology and in Section \ref{sec:observables} we define our observables. We present our results in Section \ref{sec:results}, which we interpret and discuss in Section \ref{sec:discussion}. We summarise and conclude in Section \ref{sec:conclusions}. Throughout this paper we use assume a WMAP7 cosmology \citep{KomatsuEtAl11}.
\section{Data} \label{sec:data}
The SDSS is a photometric and spectroscopic survey, undertaken using a
dedicated 2.5m telescope in Apache Point, New Mexico. For details on
the hardware, software and data-reduction see \citet{YorkEtAl00} and
\citet{StoughtonEtAl02}. In summary, the survey was carried out on a
mosaic CCD camera \citep{GunnEtAl98}, two 3-arcsec fibre-fed
spectrographs, and an auxiliary 0.5m telescope for photometric
calibration. Photometry was taken in five bands: $u, g, r, i$ and $z$,
and Luminous Red Galaxies (LRGs) were selected for spectroscopic
follow-up according to the target algorithm described in
\citet{EisensteinEtAl01}. In this paper we analyse the latest SDSS LRG
sample (data release 7, \citealt{AbazajianEtAl09}), which includes
around 180,000 objects with a spectroscopic footprint of around 8000
sq. degrees and a redshift range $0.15 <z < 0.5$.
The target selection was designed to follow a passive stellar population in colour and apparent magnitude space, and it targets LRGs below and above $z\lesssim 0.4$ with two distinct cuts. Cut I targets low redshift LRGs by using the following cuts:
\begin{equation}
\begin{array}{l}
\displaystyle r_{p} < 13.1 + c_\parallel/0.3 \\
\displaystyle r_{p} < 19.2 \\
\displaystyle c_\perp < 0.2 \\
\displaystyle \mu_{r,p} < 24.2 \text{ mag arcsec}^2 \\
\displaystyle r_{psf} - r_{model} > 0.3 \\
\end{array}
\end{equation}
where the two colours, $c_\parallel$ and $c_\perp$ are defined as
\begin{equation}
\begin{array}{l}
\displaystyle c_\parallel = 0.7(g-r) + 1.2[(r-i) - 0.18] \\
\displaystyle c_\perp = (r-i) - (g-r)/4 - 0.18. \\
\end{array}
\end{equation}
Model magnitudes are used for the colours, and petrosian magnitudes for the apparent magnitude and surface brightness cuts. Cut II targets LRGs at $z\gtrsim 0.4$ following:
\begin{equation}
\begin{array}{l}
\displaystyle r_p < 19.5 \\
\displaystyle c_\perp > 0.45 - (g-r)/6 \\
\displaystyle (g-r) > 1.3 + 0.25(r-i) \\
\displaystyle \mu_{r,p} < 24.2 \text{ mag arcsec}^2 \\
\displaystyle r_{psf} - r_{model} > 0.5 \\
\end{array}
\end{equation}
Two separate cuts are necessary as the passive stellar population turns sharply in a $g-r$ vs $r-i$ colour plane, when the 4000\AA\ break moves through the filters. This bend in the colour selection results in a broader colour range of targets around it.
\section{Method} \label{sec:method}
Our approach consists of tapping into the fossil record of
low-redshift ($z\lesssim0.25$) LRGs to measure their stellar evolution
in terms of their past star-formation and metallicity history out to
$z = 0.45$ (T11). As discussed in T11, using the fossil record allows us to explore
past star-formation histories in a way that is completely decoupled
from the survey's selection function. Although the SDSS selection was
designed to follow a passively evolving stellar population, it is not
true that all galaxies that pass the selection criteria follow this
model.
With the full knowledge of how the selection function of the survey
changes with redshift, we predict the number and luminosity densities
of LRGs at higher redshifts based on low-redshift samples. In the
absence of mergers, the predicted and observed number and luminosity
densities should match at high redshift. In
Section~\ref{sec:discussion} we show how, by making a few simple
assumptions, the difference between the two quantities can be
interpreted as a minimum merger history.
\subsection{Evolving galaxies back in time}
For each galaxy we consider the past history, and at which epochs it
would have been included in the survey. To do this we need to consider
the following:
\begin{enumerate}
\item the stellar evolution of the galaxy; \label{item1}
\item the selection function of the survey; \label{item2}
\item changes in the photometric errors with redshift. \label{item3}
\end{enumerate}
For~\ref{item1} we use the stellar evolution models for the LRGs of
\cite{TojeiroEtAl11}, and refer the reader to that paper for full
details of how these models were obtained. For completeness, we present a summary in Section \ref{sec:kcorrections}.
To account for \ref{item2}, we run the evolution of each individual
galaxy through the selection cuts of the LRG sample (see Section
\ref{sec:data} and \citealt{EisensteinEtAl01}), between
the redshift of the galaxy and $z=0.45$, and we construct a vector
that tells us exactly when any individual galaxy would have been
observed within the survey, were it to exist on our past light cone at
all epochs. We call this vector $V_{obs}(z)$ and we set it to unity if
the galaxy would have been observed at that redshift, and to zero
otherwise.
To incorporate the changing photometric errors (\ref{item3}) into our
colour modelling we take the following steps. First we compute the
average photometric errors in colour as a function of apparent
$r-$band cmodel magnitude, using all LRGs in the sample. Second, for
each galaxy we compute the $g-r$ and $r-i$ offsets with respect to
their assigned model at the redshift of the galaxy - note that these
are not expected to match as each cell used to compute the 124 stacks
has a finite width in colour and redshift. We then add a term to their
predicted colour evolution which keeps the ratio between this offset
and the typical observational error at that redshift constant
throughout its whole evolution. For example, to predict the $g-r$
colour of a galaxy observed with a redshift $z_{obs}$ at any other
redshift $z > z_{obs}$ we compute
\begin{equation}\label{eq:photometric_errors}
[g-r]_{pred}(z) = [g-r]_{model}(z)
+ \Delta[g-r] \frac{\sigma_{g-r}[r(z)]}{\sigma_{g-r}[r(z_{obs})]},
\end{equation}
where $[g-r]_{model}$ is given by the stellar evolution models,
$\Delta[g-r]=[g-r]_{obs}-[g-r]_{model}(z_{obs})$, and $r(z)$ is the
predicted apparent $r-$band cmodel magnitude at that redshift. When
$z=z_{obs}$ this simply returns the observed colour, and the effect at
$z>z_{obs}$ is to correctly match wider areas of colour space at
larger redshift with narrower cells at low redshift. Note that if
there was no change in photometric errors with redshift there would be
no need to widen the cells.
\subsection{The stellar evolution models, K+e corrections and absolute magnitudes}\label{sec:kcorrections}
We follow \cite{TojeiroEtAl10} to compute K+e corrections and obtain
K+e corrected rest-frame absolute magnitudes - see Section 2.1 of that
paper for details - with two differences. First, we do not assume the
passive evolution model of \cite{MarastonEtAl09} but rather the
stellar evolution models of T11, which computed directly from the fossil record of LRGs. This allows us to proceed without making any assumptions about this evolution - thus making the results scientifically very robust. Secondly, a consequence of not wishing to make unsupported assumptions to construct evolutionary models is that we cannot predict how any galaxy would
look like at $z < z_{obs}$. We therefore K+e correct all galaxies to
a redshift that is larger than the high-redshift tail of our
distribution, and we choose $z_c=0.525$ (made to coincide with one of
the redshift nodes at which the k+E corrections are explicitly
computed).
There are many technical differences between the models of T11 and \cite{MarastonEtAl09} (e.g. \citealt{MarastonEtAl09} uses broadband optical colours whereas T11 uses optical spectra; the fitting philosophies are different). For the current work, the primary difference is that T11 uses the {\em fossil record} of galaxies at a given redshift to infer their past star formation history. In contrast, \cite{MarastonEtAl09} fits a model to the colours of galaxies {\em observed over} a range of redshifts. To do so, they have to apply k+E corrections to all galaxies in order to construct a sample that is potentially coeval. The risk of such an approach is that the inferences about the evolution of the sample reflect the assumptions used to define it, entered here via the k+E corrections - they only retained galaxies at a certain redshift that they predict would have been observed at any other redshift given the survey's selection function and the assumed stellar evolution model, thus potentially considering a biased subsample of galaxies. In our approach, however, evolutionary colour paths that stray from the colour selection box {\em are} acceptable and crucially important for the results of our paper.
In T11 we assumed that that LRGs do not necessarily share the same stellar evolution or dust content, and they naturally allow for any intrinsic variation in LRGs, according to their luminosity, observed $r-i$ colour and redshift.
Briefly, the T11 models are computed by running VESPA
\citep{TojeiroEtAl07, TojeiroEtAl09} on 124 high signal-to-noise
stacks of LRG spectra, constructed based on the observed $r-i$ colour,
luminosity and redshift of the galaxies. VESPA returns a star-formation and metallicity history, as well as a present-day dust content, for each of the 124 stacks. These stellar evolutionary
paths are turned into a spectral energy distribution, which can be
computed at any redshift larger than the redshift of the stack and
provide any set of colours, magnitudes, stellar masses or rest-frame
luminosities.
An added subtlety lies in modelling the evolution the surface brightness of the galaxy, which is needed to fully characterise the survey selection function. This
relies on modelling the physical size of a galaxy with redshift. Here we simply
assume the physical size of the galaxy does not change with time, and
we let the apparent size change with the angular diameter distance. In practice we know this is likely to be a simplification, as a growing body of literature discusses a potential size evolution of early-type galaxies (see the Introduction of \citealt{SaraccoEtAl11} and references within for a summary). These studies focus on the detection of a population of compact early-type galaxies at $z > 1$, with effective radii a few times smaller than the effective radii of their local counterparts of the same stellar mass (e.g. \citealt{vanDokkumEtAl08}). This does not seem to necessarily be a problem theoretically, as hydrodynamical simulations show that minor mergers can act to increase the effective radius of a galaxy by a factor of a few between $z=3$ and the present day \citep{NaabEtAl09}. Nonetheless, the difficulty in characterising the (small) samples and making the measurements at high redshift, as well as the detection of a population of compact early-type galaxies in the local Universe (e.g. \citealt{CappellariEtAl09}) makes any realistic modelling of the size of LRGs with redshift unfeasible. Given the small redshift range of our sample and, crucially, the small number of objects that are discarded solely due to failing the surface brightness cut ($< 0.01\%$), assuming no evolution in the physical size of our LRGs has a negligible effect on our results.
\section{Measured Quantities}\label{sec:observables}
The statistics that we use to test the evolution of the galaxies are
the galaxy number and luminosity densities, both as a function of
redshift. As in \cite{TojeiroEtAl10}, we use a proxy for the
luminosity of the galaxy as $L = 10^{-M_i/2.5}$, where $M_i$ is K+e
corrected to $z_c=0.525$, and we work in arbitrary units of luminosity
density throughout.
\subsection{The observed quantities}
We compute
\begin{equation}
n_{obs}(z) = \frac{1}{V_{\Delta z}}\sum_{g \in \Delta z} \frac{1}{c_g},
\end{equation}
\begin{equation}
\ell_{obs}(z) = \frac{1}{V_{\Delta z}}\sum_{g \in \Delta z} \frac{L_g}{c_g},
\end{equation}
where $c_g$ is an estimate of the spectroscopic completeness at the
location of each galaxy, and corrects for the small fraction of
galaxies in the target sample that were not observed \citep{PercivalEtAl07}. $\Delta z$ is the width in redshift in which $n(z)$ and $\ell(z)$ are computed. We sample $n(z)$ and $\ell(z)$ in around 50 bins, between $z_{min} = 0.2$ and $z_{max} = 0.45$.
\subsection{The predicted quantities}
We begin by defining a redshift range, at low redshift, that we use to
predict the number and luminosity densities to higher redshifts. This
redshift bin, which we denote $\Delta z_{\rm base}$, contains galaxies
with redshifts in an interval can be varied in order to learn more
about the evolution of LRGs. For each redshift greater than those in
this bin we compute
\begin{equation}
n_{pred}(z) = \frac{1}{V_{\Delta z_{base}}} \sum_{g\in\Delta z_{base} }
\frac{1}{c_g}V_{obs,g}(z)
\end{equation}
\begin{equation}
\ell_{pred}(z) = \frac{1}{V_{\Delta z_{base}}} \sum_{g\in\Delta z_{base}}
\frac{L_g}{c_g } V_{obs,g}(z).
\end{equation}
\subsection{Rates of change}
We define a number loss rate as
\begin{equation}\label{eq:n_loss_rate}
r_n(z) = \frac{1}{\Delta t}\left(1 - \frac{n_{pred}(z)}{n_{obs}(z)} \right),
\end{equation}
where $\Delta t=t(\overline{\Delta z})-t(\overline{\Delta z_{base}})$, and a luminosity loss rate as
\begin{equation}\label{eq:l_loss_rate}
r_\ell(z) = \frac{1}{\Delta t}\left(1
- \frac{\ell_{pred}(z)}{\ell_{obs}(z)} \right).
\end{equation}
We can compute the average luminosity lost per galaxy as
\begin{equation}\label{eq:merger_rate}
r_g(z) = \frac{1}{\Delta t} \left(1
- \frac{n_{pred}(z) / \ell_{pred}(z)} {n_{obs}(z) / \ell_{obs}(z)} \right).
\end{equation}
This is also the best estimate for the true merger rate of the
galaxies that are the progenitors of the galaxies in $\Delta
z_{base}$, assuming that any contaminants at redshift $z$ have the
same luminosity distribution as the progenitors.
\section{Results} \label{sec:results}
In this Section we show our results for a selection of $\Delta
z_{base}$ and magnitude ranges of local galaxies. If the full sample
was made of a coeval population of galaxies changing these values
would have no effect on the derived merger rates. From previous work
we know this assumption is likely to fail on two accounts: the
targeting is less robust at $z\lesssim 0.2$ as can be seen by the fact
galaxies at these redshifts tend to have small amounts of recent to
intermediate star formation and show an upturn in dust content
\citep{TojeiroEtAl11}; and the fainter objects show stronger evidence
of a non-zero merger history \citep{TojeiroEtAl10}. As our stellar
evolution models are independent of the selection function to larger
redshift, each of these two effects should become apparent as we vary
the magnitude and the redshift of our local sample. Our interpretation
is presented in Section \ref{sec:discussion}.
\begin{figure}
\begin{center}
\includegraphics[width=3.4in]{rates_allM_zA.ps}
\caption{The observed and the predicted number and luminosity
densities (first and second panel from the top,
respectively). Number density in units of $10^{-4}$ Mpc$^{-3} h^3$
and the luminosity density in arbitrary units and Mpc$^{-3}
h^3$. The observed values are in black, and the predicted values are
in red for the number and blue for the luminosity densities. The
shaded area shows the redshift range used to compute the predicted
values. The third panel shows the loss rates in numbers (red) and
luminosity (blue) given by Eqs.~\ref{eq:n_loss_rate} and
\ref{eq:l_loss_rate}. Bottom: the galaxy luminosity loss rate given
by Eq.~\ref{eq:merger_rate}. All rates are in Gyr$^{-1}$.}
\label{fig:rates_allM_zA}
\end{center}
\end{figure}
We begin by making no selection in luminosity, and comparing observed
and predicted densities from galaxies with $\Delta z_{base} = [0.20,
0.22]$ in the top two panels of Fig.~\ref{fig:rates_allM_zA}. On the
data side, we see a roughly constant number density from $ z \approx
0.23$ to 0.33, an excess relative to this plateau at $z \lesssim
0.23$, a bump at $z\approx0.34$, and a steep decline onwards. The
roughly constant densities are a direct result of the targeting
algorithm, which aimed to select LRGs up to a fixed number density by
following a passively evolving stellar population. The fact that these
are roughly constant can in itself put some constraints on the
dynamical evolution of the sample \citep{WakeEtAl06,WakeEtAl08}. The
excess at low redshift is likely to be contamination to the sample,
where the target algorithm has less distinguishing power. The bump at
$z \approx 0.34$ is a result of the widening colour cuts at that
redshift, as we move from cut I to cut II in the target selection. The
steep decline after that is dominated by the apparent magnitude cuts,
with the decline in the observed number and luminosity densities being
steeper than that predicted from the low redshift data.
The loss rates given by Eqs.~\ref{eq:n_loss_rate} and
\ref{eq:l_loss_rate} are shown on the third panel of
Fig.~\ref{fig:rates_allM_zA}, and we combine the information in all
four observables to compute the merger rate, or galaxy luminosity loss
rate, given by Eq.~\ref{eq:merger_rate}, which we show on the bottom
panel of the same Figure. The merger rate is roughly constant and of
the order of 3-4\%, up to the point where the sample is dominated by
the apparent magnitude limit.
In Fig.~\ref{fig:rates_allM}, we show how the predicted quantities
change if we instead compute them from galaxies in different ranges
$\Delta z_{base}$. Note that differences in the predicted quantities
(in the blue and red lines) across the different plots provides
information about the nature and evolution of the galaxies at each of
the redshift intervals - this is discussed further in
Section~\ref{sec:discussion}.
\begin{figure*}
\begin{center}
\includegraphics[width=2.3in]{rates_allM_zB.ps}
\includegraphics[width=2.3in]{rates_allM_zC.ps}
\includegraphics[width=2.3in]{rates_allM_zD.ps}
\caption{The same as Fig.~\ref{fig:rates_allM_zA}, but exploring
different intervals for $\Delta z_{base}$, used to compute the
predicted quantities.}
\label{fig:rates_allM}
\end{center}
\end{figure*}
We can also investigate how the loss rates change with the magnitude
of the galaxies. As all absolute magnitudes are K+e corrected, a flat
cut in $M_i$ with redshift selects the same population of galaxies in
the absence of mergers. In Fig.~\ref{fig:rates_M23}, we show how the
observed and predicted quantities change when we restrict the analysis
to galaxies with $M_i < -23$: this corresponds roughly to the top 30\%
brightest galaxies in the LRG sample. On the data side, we see many
features in the redshift distributions disappear, leading to a flatter
curve. The exception is at low redshift, where the excess in density
is still present. We show the same redshift ranges of $\Delta
z_{base}$ as explored in Figs.~\ref{fig:rates_allM_zA} and
\ref{fig:rates_allM}. Once again the observed curves are the same
across all four panels, with differences in the red and blue curves
holding information on the evolution of bright LRGs. We interpret and
discuss these results in the next Section.
\begin{figure*}
\begin{center}
\includegraphics[width=2.8in]{rates_M23_zA.ps}
\includegraphics[width=2.8in]{rates_M23_zB.ps}
\includegraphics[width=2.8in]{rates_M23_zC.ps}
\includegraphics[width=2.8in]{rates_M23_zD.ps}
\caption{Same as Fig.~\ref{fig:rates_allM_zA}, but using only galaxies
with $M_i < -23$ and exploring different intervals for $\Delta
z_{base}$, used to compute the predicted quantities.}
\label{fig:rates_M23}
\end{center}
\end{figure*}
\subsection{Errors} \label{sec:errors}
The errors quoted throughout this paper are Poisson, and assume
that our ability to compare predicted with observed quantities depends
only how many objects we have available to do so. We account for the
change in photometric errors using Eq.~\ref{eq:photometric_errors},
which is folded into the predicted quantities.
It is particularly hard to estimate the effect of errors from the recovered star-formation histories on the computed merger rates. In T11 we presented an extensive analysis of the errors (see their Section 3.4), which we summarise here for completeness. There are two sources of errors that can affect the recovered star formation histories from VESPA: photon noise, which perturbs the spectra in way that we assume we understand; and systematic errors in models or parametrization, which are the result of our imperfect ability to correctly model certain stages of stellar evolution, chemical enrichment or dust extinction. The former is easily estimated, via a Monte Carlo approach that we explain in \cite{TojeiroEtAl09}. In this case we are in the regime of extremely high signal-to-noise spectra and photon error is negligible. In order to understand systematic errors, in T11 we estimated these using the residuals of the best fit solution to the data (see their Fig. 3) - we found these errors to be at least 10 times larger. We use this error vector to obtain the SFHs used in the current analysis.
The effect of the uncertainties in the SFH on the measured rates comes from how many galaxies leave the survey's selection window as a function of redshift, and how bright they are. The window function of the survey is designed to follow a passive evolution, so only star-formation histories that depart significantly from this model will affect the rates. An episode of recent to intermediate star formation of anything larger than 5\% by mass is enough to take galaxies out of the survey box for up to and around 1Gyr (the exact number depends on the length of episode, the Stellar Population Synthesis (SPS) models used, the metallicity and the exact age of the burst) and, even though none of our formal error estimates allow for a discrepancy of this size, solutions obtained with different SPS models can differ by a few times this amplitude - we explicitly address the effect of using different SPS models in our results in Section~\ref{sec:SPS}.
The effect on the merger rate in turn depends on {\em how many} galaxies are removed from the sample. Our assumption in T11 is that all galaxies within a cell follow the SFH given by the stack that represents that cell. Each of our $\Delta z_{base}$ samples is typically represented by 4 to 12 stacks (depending on the luminosity range of the sample), and if one stack predicts a significant star-formation event ($\gtrsim$ 5\% by mass), then 1/12 to 1/4 of the galaxies will be predicted to leave the sample for a significant period. This means that the effect on the merger rates are somewhat quantised, rather than a smooth function of an error in the SFHs. The effect can obviously be quite dramatic, and at its extreme it can predict that almost all galaxies would leave the sample, suggesting that the observed red sequence is an intricate combination of galaxies becoming red and blue throughout cosmic history (see Section~\ref{sec:SPS} for an example).
\section{Discussion}\label{sec:discussion}
\subsection{Interpreting the observables}\label{sec:interpreting_observables}
If the sample of SDSS-II LRGs was not affected by mergers, and assuming the SFHs are correct (see Section \ref{sec:errors}), then the
observed luminosity and the number densities would obviously match the
extrapolated values. Any deviations from this, implies merger activity and/or a non-zero net flow of galaxies into the sample.
We cannot determine a balance between mergers into and
out of the sample. However, by comparing number and luminosity
evolution we can start to understand the evolution that we do see.
\renewcommand{\theenumi}{(\alph{enumi})}
\begin{figure}
\begin{center}
\includegraphics[width=4.8in]{diagram.eps}
\caption{Diagram of possible interpretations for discrepancies in the observed and predicted quantities. The middle column shows the observed luminosity and number densities. The solid arrows show the prediction given by the stellar evolution from VESPA and the survey's window function, and they lead to the left hand side column that shows the predicted luminosity and number densities. Note that whereas VESPA can only predict changes in the luminosity density, the window function can also affect the number densities. The window function is fully known and introduces no ambiguity when computing the predicted quantities - in here we show the numbers being conserved just for simplicity. The right hand side column shows the observed number and luminosity densities, and the dotted arrows represent our proposed interpretation.
See Section~\ref{sec:interpreting_observables} for more details.}
\label{fig:diagram}
\end{center}
\end{figure}
Fig.~\ref{fig:diagram} shows schematically the simplest interpretation
for three scenarios:
\begin{enumerate}
\item $n_{pred} < n_{obs}$ and $\ell_{pred} = \ell_{obs}$ - the
simplest interpretation is that galaxies within the sample have
merged together to decrease the predicted number density of objects,
whilst retaining the sample luminosity. This assumes no luminosity
loss to the intra-cluster medium.
\item shows how contamination from objects at a given redshift could
also act to give $n_{pred} < n_{obs}$ in the same way as internal
mergers. By contaminants here we refer to objects that would enter
the sample for a short time only, and do not evolve to be present in
the sample at low redshift. However, the comparison between
$\ell_{obs}$ and $\ell_{pred}$ helps to differentiate the two
scenarios, as in this case we expect the excess in the observed number
density due to contamination to be matched by an excess of
luminosity density. The two excesses will be matched in amplitude provided the contaminants have the same
luminosity function as the galaxies that evolve through the sample.
\item $n_{pred} > n_{obs}$ and $\ell_{pred} > \ell_{obs}$. This
scenario requires an enrichment of the sample towards lower redshift
- either by the merging of fainter galaxies becoming brighter and
crossing over the faint magnitude cuts, bluer galaxies turning
redder and crossing over the colour cuts, or a combination of both.
\end{enumerate}
Note that in the case of fainter galaxies merging together to become
brighter and entering the sample by crossing over the magnitude cuts,
scenarios (a) and (c) represent the same physical process - the only
difference between is where the survey's magnitude cut lies. We
explore this fact in a little more detail in
Section~\ref{sec:mergers_LF}.
Obviously, we can expect multiple scenarios such as those in
Fig.~\ref{fig:diagram} to happen simultaneously. If this is the case, under the assumption that the luminosity function of the
contaminants matches that of the objects in the sample, and that
luminosity is conserved in a merger event, Eq.~\ref{eq:merger_rate}
gives the best estimate for the merger rate required for galaxies
within the sample between $\Delta z$ and $\Delta z_{base}$. In fact, even if
these assumptions are broken, Eq.~\ref{eq:merger_rate} still acts as a
{\it lower limit} on the true merger rate. This is mostly easily seen from Equation (\ref{eq:merger_rate}) - we are effectively under-estimating $\ell_{pred}$, as some luminosity has been lost to the intracluster medium, resulting in an under-estimation of $r_g$.
\subsection{Interpreting the results} \label{sec:interpret_results}
Using the results of Section~\ref{sec:results}, we can identify
redshift intervals that are likely to be dominated by each of the
scenarios shown in Fig.~\ref{fig:diagram}. Starting with the results
for all LRGS, shown in Fig.~\ref{fig:rates_allM_zA}, we see the
predicted number and luminosity densities both fall short of the
observed values up to $z\approx0.38$, suggesting a combination of
scenarios (a) and (b). The bump at $z\approx0.34$ is not reproduced by
the stellar evolution of present-day galaxies, strongly suggesting
that is it caused by contamination of galaxies as the colour cuts
widen - in other words, the wider breadth of colour in the sample at
those redshifts is not predicted by the stellar evolution of more
local galaxies. Therefore the signal here is likely dominated by
scenario (b). Finally, the sample of local galaxies predicts a much
shallower slope in the number densities as we go towards higher
redshifts, suggesting that they were still being assembled (in a
dynamical sense) at those epochs - in other words, had the local
galaxies still be in one piece at $z \gtrsim 0.4$ they would have been
bright enough to be in the sample. This puts us in a regime likely to
be dominated by scenario (c) at these high redshifts.
Moreover, at the high-redshift tail where scenario (c) dominates, we
note that the number loss rate is higher (in amplitude) than the
luminosity loss rate. In other words, the typical luminosity per
object at $z\gtrsim 0.38$ is lower than expected. This suggests that
the LRG growth due to external merging is happening predominantly at
the fainter end of the sample. We will investigate this hypothesis
further in Section \ref{sec:changing_mag}. Note that if this growth
was constant with luminosity then both loss rates would have a
negative value, but would be matched in amplitude - resulting in a
zero galaxy luminosity loss rate, interpreted in this case as a zero
internal merger rate.
\subsubsection{Changing $\Delta z_{base}$}
We now consider the effect of changing the low redshift intervals
$\Delta z_{base}$, from which we infer the galaxy evolution from high
redshift as shown in Fig.~\ref{fig:rates_allM}. Changes in the inferred
loss and merger rates as we change our low-redshift sample are
sensitive to differences in the low redshift population of galaxies.
We find no significant difference in the evolution of galaxies
observed between $z = 0.23$ and $z=0.27$ - see the first two panels of
Fig.~\ref{fig:rates_allM} - but comparing the rate of luminosity loss
per galaxy (lower panels in the figures) to those in
Fig.~\ref{fig:rates_allM_zA} suggests that the lower redshift galaxies
experience a more pronounced growth. This is consistent with a
hypothesis in which the low redshift end of the sample is not as pure
as the rest of the sample - the targeting algorithm is less robust at
these redshifts \citep{EisensteinEtAl01, TojeiroEtAl11}. Furthermore,
the sample will naturally be more enriched towards fainter galaxies,
which could have a different growth rate - see Section
\ref{sec:changing_mag} for more details.
Similarly, galaxies observed at redshifts between $0.33$ and $0.35$
are seen to have a different growth compared with galaxies at lower
redshifts. This is not unexpected, as the bump in number and
luminosity density at $z\approx0.34$ is not predicted by galaxies at
low redshifts. These results suggest that the wider colour selection
of cut II selects galaxies not only with a different stellar
evolution, but also a different merger growth rate - essentially a
different population of galaxies. For the same reasons as before, we
argue that this growth must be happening predominantly at the faint
end.
\subsubsection{Changing the luminosity}\label{sec:changing_mag}
In Fig.~\ref{fig:rates_M23} we restrict our analysis to galaxies with
a K+e corrected M$_i < -23$, which corresponds roughly to the
brightest third of the galaxies in the sample. We show results for the
same four ranges of $\Delta z_{base}$ as before. It is immediately
clear that a lot of the features in the observed quantities are less
pronounced - the bump at $z\approx0.34$ almost disappears, and the
high-redshift tail flattens out as we restrict ourselves to what is
close to magnitude-limited sample (colour cuts excluded). The
exception is the bump at low redshifts, which, combined with the
measured loss and merger rates for $\Delta z_{base} = [0.20, 0.22]$,
suggests the luminous galaxies observed at low redshift are still
growing, and this growth is approximately matches that of lower
luminosity objects. Clearly this also fits with the hypothesis that
the lower redshift galaxies form a population with more dynamical
evolution than the higher redshift galaxies within the sample.
For all other cases of $\Delta z_{base}$ we find a picture that is
consistent with passive dynamical growth, with the only - and marginal
- evidence for external growth coming for the galaxies in $\Delta
z_{base} = [0.33, 0.35]$, perhaps suggesting that not even the
brightest galaxies are purely coeval across the two LRG colour cuts
described in Section~\ref{sec:data}.
\subsection{Probing the typical luminosity of mergers}\label{sec:mergers_LF}
We noted in Section \ref{sec:interpreting_observables} that scenarios
(a) and (c) represent the same physical merger process, but can be
distinguished as (c) caused objects to cross the apparent magnitude cut
on the sample. In Section~\ref{sec:interpret_results}, we suggested
than (c) dominates at high redshift, while (a) dominates at low
redshift. We can use the information available to define a typical
luminosity for the dominant merger process happening.
For each case of $\Delta z_{base}$ we take the redshift at which the
number and luminosity loss rates cross the zero line and compute the
absolute magnitude that would correspond to the apparent petrosian
magnitude equal to the survey's limit of $r_p = 19.5$ - in other
words, the faintest absolute magnitude observed at that
redshift. These are the galaxies that are beginning to enter the
sample when dynamical growth transitions from external to internal. Note that the two growth rates change sign at very similar epochs, so we can use either without significant effect on our results. In
Fig.~\ref{fig:Mr_mergers} we show how this transitional absolute
magnitude changes for the four different cases of $\Delta z_{base}$
(those shown in Figs.~\ref{fig:rates_allM_zA} and
\ref{fig:rates_allM}).
\begin{figure}
\begin{center}
\includegraphics[width=3.4in]{Mr_mergers.ps}
\caption{The dots show the transitional absolute magnitude as a
function of $\Delta z_{base}$ (on the right-hand side y-axis) - see
Section \ref{sec:mergers_LF}. This is the absolute $r-band$
magnitude that corresponds to the $r_p = 19.5$ cut on petrosian
magnitude of cut II \citep{EisensteinEtAl01} at the redshifts for
which we infer a transition from external to external growth. For
context we also show the distribution of absolute magnitudes of
galaxies in the sample (on the left-hand side y-axis). }
\label{fig:Mr_mergers}
\end{center}
\end{figure}
Once again our result is that the dynamical growth is happening
predominantly at the faint end, for galaxies observed in all the four
intervals of $\Delta z_{base}$ we studied.
\subsection{Robustness to SPS models}\label{sec:SPS}
In \cite{TojeiroEtAl11} we showed how the measured stellar evolution
of LRGs depends on the SPS models used - see Fig. 6 in that paper for
a summary of the differences. The results presented in this paper thus
far assume the stellar and colour evolution given by the Flexible
Stellar Population Synthesis (FSPS) models of \cite{ConroyEtAl09,
ConroyAndGunn10}, which gave better fits to the spectral data and
resulted in the more passive stellar evolution model in our previous
study. However, as we could not conclusively decide which set of SPS
models gives a more accurate description of the real stellar evolution
of LRGs, in this section we discuss how the stellar evolution given by
other models would change the interpretation given above. The results
on the recovered merger rate and interpretation of the dynamical
evolution can be dramatically affected by events of star formation.
In the case of the Maraston et al. 2011 models (M10, submitted), the
amount of recent to intermediate star formation dominates only in
galaxies at $z\lesssim0.3$. The resulting interpretation is that
contamination of the LRG sample extends to a larger redshift than what
would be assumed using the FSPS models, and that these local galaxies
are not the evolutionary products of the LRGs observed at higher
redshifts. At $z>0.3$, however, the recovered dynamical evolution is
qualitatively similar to that obtained with the FSPS models, although
with slightly larger rates.
The interpretation using the stellar evolution given by the models of
\cite{BruzualEtCharlot03} (BC03), however, is incompatible with the
hypothesis that the LRGs form a coeval population of galaxies since
$z<0.45$. The significant and continuous (with redshift) amount of
recent to intermediate star formation lead to a consistently
under-prediction of the number and luminosity densities at high
redshifts from more local samples, in a way that is only explained by
a continuous amount of contamination in the sample.
\subsection{Comparison with previous results}
The results in this paper are in qualitative agreement with those in
\cite{TojeiroEtAl10}, where we used the luminosity weighted
power-spectrum to argue that the LRG population is growing via
external mergers at the faint end. Quantitatively, however, the rates
we find in this paper are lower by roughly a factor of two. The main
difference between the two approaches lies on the treatment of the
stellar evolution - whilst in \cite{TojeiroEtAl10} we assumed that the
passive stellar evolution model of \cite{MarastonEtAl09} was suitable
for all galaxies in the sample, in here we instead use the stellar
evolution models of \cite{TojeiroEtAl11} which use the fossil record
of LRGs of different colour, luminosity and redshift to infer the
most-likely stellar evolution. Furthermore, in \cite{TojeiroEtAl10},
we assumed that the scatter in colour and magnitude was due to
intrinsic or photometric errors, which is a byproduct of assuming that
all LRGs in the sample share the same stellar evolution. Instead, the
analysis presented in this paper makes no such assumption, and is
robust to an influx and/or an outflux of objects if we interpret the
measured rates as a lower limit. It is the independent treatment of
the stellar evolution, and the detailed comparison of predicted with
observed quantities that allows us to make an assessment on the nature
of the mergers without the need for a clustering analysis.
Comparisons with other works is less trivial, due to the typically
different samples used. We refer the reader to Section 8 of
\cite{TojeiroEtAl10} for a summary of recent results and
comparison. We continue to argue that merging at the faint end of the
LRG population is a more likely explanation published for the
published evidence for both luminosity growth \citep{BrownEtAl07,
WhiteEtAl07, CoolEtAl08, MasjediEtAl08} and loss of number density
\citep{MasjediEtAl06, WakeEtAl06, WhiteEtAl07, WakeEtAl08,
deProprisEtAl10}.
\section{Summary and conclusions} \label{sec:conclusions}
In this paper we have demonstrated how we can use the fossil record of
galaxies to extrapolate from low redshift samples to match samples
selected at higher redshift. By comparing against the observed number
and luminosities, we can measure growth and consider where the
dynamical evolution of galaxies is required. We have applied this
technique to consider the dynamical evolution of galaxies in the
SDSS-II LRG sample, although the technique could be used to link
different surveys at low and high redshift.
For the SDSS-II LRG sample, we have previously predicted the stellar
evolution in \cite{TojeiroEtAl11}, by using the fossil record encoded
in the spectra of LRGs. This allowed us to compute the most likely
stellar evolution of observed LRGs in a way that was decoupled from
the survey's selection function. We used this information to make
predictions on the number and luminosity of LRGs at higher redshift
from a local sample of galaxies, and we showed how differences in
these quantities can be interpreted as a minimum merger rate.
Our main conclusions can be summarised as the two following points:
\begin{itemize}
\item the LRG population is not purely coeval, with some of galaxies
targeted at $z<0.22 $ and at $z>0.34$ showing different dynamical
growth than galaxies targeted throughout the sample;
\item the LRG population is still dynamically growing, and this growth
must be limited to the faint end;
\end{itemize}
Based on these results, we argue that a coeval population of galaxies
that has been dynamically passive to less than 1-2\% can be robustly
selected from the LRG sample, by imposing a cut on the K+e corrected
absolute magnitude, and by selecting galaxies with $z \in [0.23,
0.45]$. This merger rate is lower than those previously measured by roughly a factor of two,
because, by using empirically determined evolutionary tracks we no
longer need to assume passive evolution, and this previously led to incorrect
evolutionary tracks for some galaxies.
We caution against interpreting these results beyond the limits of the stellar population models, and once again stress the important of good models and statistically good fits for analyses of this type. However, given the power inherent in the type of approach, and improvements in SPS modelling due in the future, we consider this analysis to be a forerunner of significant future work.
\section{Acknowledgments}
We thank the anonymous referee for a report that helped us clarify a number of issues. We thank David Wake for helpful discussions and encouragement. RT thanks Alan Heavens and Raul Jimenez for help in the development of VESPA.
RT thanks the Leverhulme trust for financial support. WJP is
grateful for support from the UK Science and Technology Facilities
Council (grant ST/I001204/1), the Leverhulme trust and the European Research Council.
Funding for the SDSS and SDSS-II has been provided by the Alfred
P. Sloan Foundation, the Participating Institutions, the National
Science Foundation, the U.S. Department of Energy, the National
Aeronautics and Space Administration, the Japanese Monbukagakusho, the
Max Planck Society, and the Higher Education Funding Council for
England. The SDSS Web Site is http://www.sdss.org/.
\bibliographystyle{mn2e}
|
\section{Introduction}
An exciton is a bound pair of an electron and a hole in a
semiconductor. At low densities, exciton are Bose quasiparticles
with a small mass, somewhat similar to hydrogen atoms. At low
temperatures, of the order of 1 Kelvin, the excitons are expected to
realize the phenomenon of Bose-Einstein condensation (BEC).
\cite{Keldysh1968} Recently, it has been shown that indirect exciton
(spatially separated electron-hole pairs) in coupled quantum wells
(CQW) can have a long lifetime and high cooling rate. With these
merits, Butov \textit{et al.} have successfully cooled the trapped
excitons to the order of 1 K. \cite{Butov2002a} Although there is not
enough evidence to prove that these excitons are condensed into the
BEC state, it is fascinating enough to observe several novel
features of the photoluminescence (PL) patterns.\cite{Butov2002b,
LaiCW2004}
Major features of the macroscopically ordered exciton states
observed are summarized as follows.\cite{Butov2002b, LaiCW2004} (i)
Two exciton rings are formed. When the focused laser is used to
excite the sample and prompt luminescence is measured in the
vicinity of the laser spot, a ring, called the internal ring, is
formed. A second ring of PL appears away from the source at the
distance about 1 mm, called the external ring. (ii) The intervening
region between the internal and external rings are almost dark
except for some localized bright spots. (iii) Periodic bright spots
appear in the external ring. The bright spots follow the external
ring either when the excitation spot is moved over the sample, or
when the ring radius is varied with the excited power. (iv) PL is
washed out eventually with increasing temperatures. (v) In an
impurity potential well, the PL pattern becomes much more compact
than a Gaussian with a central intensity dip, exhibiting an annular
shape with a darker central region. (vi) With the increase of the
laser power, exciton cloud first contracts then expands. (vii)
Exciting sample by higher-energy lasers, the dip can turn into a tip
at the center of the annular cloud.
The above experimental facts raised several interesting questions
which need to be clarified. (i) What is the reason behind the
two-ring structure? (ii) What is the physical origin of the periodic
periodic bright spots in the external ring? And, (iii) is the a
coherent phenomenon, or a result due to unbalanced transportation?
In Ref.~\onlinecite{butov:117404}, a charge separated
transportation mechanism was proposed. It gave a satisfactory
explanation to the formation of the exciton ring and the dark region
between the internal and external rings. However, the remaining two
questions are still not clarified satisfactorily.
It is commonly believed that the periodic bright spots in the
external ring are formed due to certain kind of instability. It
generates the density modulation by a positive feedback. Regarding
the origin of the instability, Levite \emph{et al.} considered that
exciton states are highly degenerate and the instability comes from
the stimulated scattering.\cite{Levitov2004a} When a local
fluctuation occurs in the exciton density, it will lead to an
increase in the stimulated electron-hole binding rate. The
depletion of local carrier concentration then causes neighboring
carriers to stream towards the point of fluctuation. On the other
hand, Sugakov suggests that the instability is due to an attractive
interaction between the high-density excitons.\cite{Sugakov2004a}
The creation of different structures, for example, the islands or
the rings of the condensed phase, occurs due to the nonequilibrium
state of the system connected with the finite value of the exciton
lifetime and the presence of pumping. Therefore, the appearance of
the patterns is likely a consequence of self-organization processes
in a nonequilibrium system. \cite{sugakov:115303}
To understand the subtle properties of indirect exciton, further
experiments have been done and reported. For example, a novel method
was proposed to demonstrate that cold exciton gases can be trapped
by the laser induced trapping, which is similar to the trapping of
BEC ultracold atoms. \cite{hammack:227402} More recently, an
improved trapping technique was used to lower the effective
temperature of indirect excitons. \cite{high-2008} With the even low
temperature, four narrow PL lines have been observed, which
corresponds to the emission of individual states of indirect
excitons in a disorder potential. The homogeneous line broadening
increases with density and dominates the linewidth at high
densities.
In a recent paper (Ref.~\onlinecite{yang:033311}), for the first
time, Yang \textit{et al.} measured the exciton PL energy along
the circumference of the ring. The most interesting result is that
the average exciton energy depends on temperature {\em
nonmonotonically}. With reducing temperatures, average exciton
energy of the indirect exciton is lowered until the transition
temperature, $T_{tr} \simeq 4$ K, is met. Below $T_{tr}$, the
macroscopically ordered exciton state is formed but the average
exciton energy in the ring actually increases with temperature
being further lowered. In particular, the largest energy of single
exciton is found to locate in the brightest regions. Due to these
observations, Yang \textit{et al.}\cite{yang:033311} argued that
the interaction between exciton is repulsive. A numerical
calculation seems to support this scenario.
\cite{schindler:045313}
Another important experiment is the measurement of the coherence
length. A Mach-Zehnder interferometer with spatial and spectral
resolution was used to probe the spontaneous coherence in cold
exciton gases, which are implemented experimentally in the ring of
indirect exciton in CQW. \cite{yang:187402} A strong enhancement of
the exciton coherence length is observed at temperatures below a few
Kelvin. The increase of the coherence length is correlated with the
macroscopic spatial ordering of excitons. The coherence length at
the lowest temperature corresponds to a very narrow spread of the
exciton momentum distribution, much smaller than that of a classical
exciton gas. It also shows that the apparent coherence length is
well approximated by the quadratic sum of the actual exciton
coherence length and the diffraction correction given by the
conventional Abbe limit divided by $\pi$. \cite{fogler:035411}
Whether the interaction between excitons in the macroscopically
ordered state is attractive or repulsive and to what extent the
interaction affects the formation of the macroscopically ordered
state remain an open question. The current paper attempts to
clarify the above issues. In fact, experimental data had revealed
several important clues. After analyzing the experimental facts,
it is found that attractive interaction is dominant in indirect
excitons and their nonequibirium distribution plays an important
role in the macroscopically ordered state. Moreover, the
uncertainty principle is shown to lead to a good account for the
experimental phenomena and numerical simulations have confirmed
that.
This paper is organized as follows. In Sec.~\ref{Temperature
dependence of exciton average energy in the exciton external ring
-- a qualitative analysis}, based on the uncertainty principle, we
present a qualitative analysis of the temperature dependence of
the average exciton energy in the external exciton ring. No
interaction is considered in this section. In Sec.~\ref{Particle
number density dependence of exciton PL FWHM broadening and energy
shift}, the uncertainty principle is used again to explain the
particle number density dependence of the PL FWHM (full width at
half maximum) broadening and energy shift. The PL spectra, which
are broadened first and then become sharper, provide a strong
evidence of the competition between a two-body attraction and a
three-body repulsion. In Sec.~\ref{Energy spatial distribution in
exciton external ring}, we use a phenomenological nonlinear
Schr\"{o}dinger equation, together with a temperature and energy
dependent exciton distribution, to discuss the exciton energy
spatial distribution in the external ring. The largest energy of
single exciton is found to be in the brightest regions. In
Sec.~\ref{Temperature dependence of exciton average energy in the
exciton external ring}, temperature dependence of the average
exciton energy in the external ring is studied quantitatively
using the approach proposed in Sec.~\ref{Energy spatial
distribution in exciton external ring}. The calculation shows that
average exciton energy first decreases and then increases with
increasing temperatures. In Sec.~\ref{Exciton number distribution
in laser induced trap}, we give a demonstration on the exciton
distribution in the laser induced trap case. Sec.~\ref{Summary} is
devoted to a brief summary.
\section{Temperature dependence of energy: qualitative analysis}
\label{Temperature dependence of exciton average energy in the
exciton external ring -- a qualitative analysis}
To understand the experimental phenomena, we first analyze the
electron and hole creation, transportation, and the exciton
formation qualitatively. As pointed out in
Refs.~[\onlinecite{butov:117404, rapaport:117405}], when electrons
and holes are excited by lasers, they are hot electrons and hot
holes initially. Since the drift speed of hot electrons is larger
than that of hot holes (electron has a smaller effective mass),
electrons and holes are indeed charge-separated at this stage (no
true exciton is formed). Because hot electrons and holes have a
small recombination rate and optical inactive, they can travel a very
long distance from the
laser spot. After a long-distance travel, hot electrons and holes
collide with the lattices and are eventually cooled down. Due to the
neutrality of the CQW, negative charges will slow down and
accumulate far away from the laser spot. As a consequence, spatially
separated electrons and holes form excitons and become optical active at the boundary of
the opposite charges where they recombine and show a sharp luminescence ring.
This kind of charge separated transportation mechanism has been used
to interpret the exciton ring formation.\cite{butov:117404,
rapaport:117405}
It should be emphasized that, however, after the long-distance
transportation and cooling, cooled electrons and holes will meet
in the region of the external ring in the experiment of Butov
\textit{et al.} \cite{Butov2002b}. However, in the experiment of
Lai \textit{et al.} \cite{LaiCW2004}, they meet in the impurity
potential well. It is believed that only cooled electrons and
holes can form excitons. At this stage, charges are not separated,
but coupled or bound together. The long lifetime of the excitons
means the lower electron-hole recombination rate. Therefore the
elemental particles in the external ring \cite{Butov2002b} and in
the impurity potential well \cite{LaiCW2004} are \emph{excitons}.
We thus argue that it is the interaction between excitons, rather
than the unbalanced transportation, which leads to the
nonhomogeneous density distribution associated with the complex PL
patterns.
When the experimental temperature is low ($T\alt 1.4$ K), average
translational kinetic energy and in-plane momentum of excitons are
also low. In such case, excitons are optically active and the
space-dependent PL intensity is proportional to the exciton number
distribution. Since exciton number distribution is equal to the
product of the probability distribution and the total excition
number $N$, if $N$ remains unchanged, PL intensity is directly
proportional to the probability distribution. In the following
discussion, we simply take exciton density distribution as the PL
distribution. Besides, since particles can only move within the
CQW, their motions are essentially two-dimensional (2D).
The periodic array of beads in the external rings is a
low-temperature phenomenon. This means that average exciton
translational kinetic energy is low enough and its wave-particle
duality is important. Therefore exciton energy is governed by the
uncertainty principle.
The uncertainty principle predicts that $\Delta p \sim \hbar/\Delta r$,
where $\Delta p$ is the momentum spread, $\Delta r$ is spatial
uncertainty, and $\hbar$ is the Planck's constant. Since exciton's
momentum can be approximated by $p\sim \Delta p$, the corresponding
energy $E\sim {\hbar^2}/[2m(\Delta r)^2]$. With these in mind, one is able to
consider the temperature dependence of the exciton energy
distribution, such as those reported in Ref.
\onlinecite{yang:033311}. When temperature is lower than
$T_{tr}\simeq 4$K, macroscopic ordered state forms. Exciton are
confined into each bead of the external ring. Their spatial
uncertainty $\Delta r$ can be approximated by the diameter $R$ of
the bead ($\Delta r \simeq R$). Experiments have observed that the
lower the temperature is, the higher contrast the pattern is and the
smaller the $R$ is. \cite{yang:033311} This implies that when
temperature is lower, momentum uncertainty will be larger and hence
the exciton energy will be higher. This gives a qualitative
interpretation why the average exciton energy increases with
decreasing temperatures.
When temperature is increased to be higher than $T_{tr}$,
macroscopically ordered state breaks down and the confined regions
enlarge essentially. The uncertainty related energy will decrease in
principle. However, when the temperature is increased further, the
system undergoes a transition into the classical limit to which
wave character of exciton is no longer important. In this case, the
average exciton translational kinetic energy is directly
proportional to the temperature.
Therefore, non-monotonic temperature dependence of the average
exciton energy is closely related to the pattern formation, and
has little to do with the interaction between excitons. In Fig.~3
of Ref.~\onlinecite{yang:033311}, linear temperature dependence
of the average exciton energy is reported at $T>T_{tr}$. Their
results confirmed that excitons behave classically in this limit
to which $E \sim k_{B}T$ at 2D. Nevertheless, the slope of the
linear temperature dependence was measured in Ref.
\onlinecite{yang:033311} to be about $1.4\times 10^{-4}$ eV
k$^{-1}$, slightly larger than the Boltzman constant,
$k_{B}=8.62\times 10^{-5}$ eV k$^{-1}$. The difference may be due
to some effect of the interactions and extra degrees of freedom in
addition to translational motions.
\section{Exciton PL spectra: qualitative analysis}
\label{Particle number density dependence of exciton PL FWHM
broadening and energy shift}
For a long time, indirect exciton PL FWHM energy and the shift
with increasing density in the presence of a random potential
remain to be fully understood.\cite{high-2008, butov:2004} In a
previous paper \cite{Liu2006}, we have proposed a mechanism to
study the macroscopically ordered exciton states. The most
important point lies in the interaction which involves the
competition between a two-body attraction and a three-body
repulsion. Here we elaborate this point, along with the
uncertainty principle, to illustrate the PL spectra.
It is instructive to emphasize several important features in
connection with our model. The idea first came from the density
dependence of inhomogeneous exciton distribution. It is clear that
the interaction between the indirect exciton is neither purely
attractive, nor purely repulsive. If the interaction between the
indirect exciton is purely repulsive, it will drive the exciton
towards homogeneous distribution and the exciton cloud will expand
with the increase of the exciton number. On the contrary, if the
interaction is purely attractive, the system is expected to collapse
when the exciton density is greater than a critical value to which
there is not enough kinetic energy to stabilize the exciton cloud.
In addition, the case of a purely repulsive or a purely attractive
interaction is incapable to understand the experimental fact that
the exciton cloud contracts first and expands later when the laser
power is increased.\cite{LaiCW2004}
The existence of the attractive interaction does not mean that the
exciton state is unstable against the formation of metallic
electron-hole droplet. The reason is that the repulsive
interaction may dominate over the attraction in that regime and
keep the system stable. Many effects may contribute to the exciton
interaction. The most important one is the exciton dipole-dipole
interaction. Exciton behaves like a dipole, so a strong repulsion
will govern exciton interaction when two dipoles are aligned
parallel. However, if the direction of two dipoles changes from
aligned parallel to inclined, the attraction between the election
of one exciton and the hole of the other exciton will dominate.
This case can happen easily when exciton density is low. In case
of high densities, due to the strong Coulomb interaction, the
dipoles will tend to align parallel and consequently a repulsive
interaction dominates. The other important interaction originates
from the exchange effect. When two indirect exciton approach to
each other, the exchange interaction between two electrons, as
well as the one between two holes become important. This may be
another source of the attractive interaction between excitons.
In fact, the complex exciton interaction which results from the
competition of multi-effects, may be well described by the van der
Waals form. It has pointed out that the effective interaction
between exciton will be attractive when the separation between two
exciton is about 3 to 6 exciton radii.\cite{Sugakov2004a} In the
current experiment, the exciton density is about $10^{10}/$
$\mathrm{cm}^{2}$. For this density, the average distance between
the indirect exciton is about 100 $\mathrm{nm}$. As the exciton
Bohr radius $a_{B}$ is about 10 $\sim$ 50 $\mathrm{nm}$
\cite{LaiCW2004}, the average distance between excitons is about
$2\sim10$ exciton radii. Thus it is reasonable to assume that the
two-body interaction is in the attractive regime. In fact, the
attractive interaction between the exciton has been considered as
a possible candidate to describe the pattern formation observed by
experiments. \cite{Sugakov2004a, Levitov2005}
In a previous paper\cite{Liu2006}, taking into account the
two-body attraction and the three-body repulsion interactions, we
have proposed that in the dilute limit the behavior of the exciton
can be described by a \emph{nonlinear} Schr\"{o}dinger equation
\cite{Liu2006}
\begin{equation}
-\frac{\hbar^2}{2m^{\ast}}\nabla^{2}\psi_{j}+(V_{ex}-g_{1}n+g_{2}n^{2})\psi
_{j}=E_{j}\psi_{j}, \label{the schrodinger equation}
\end{equation}
where $\psi_{j}$ and $E_{j}$ are the \textit{j-th} eigenstate and
eigenvalue, respectively. $V_{ex}$ is an external potential, and
$g_{1}$ and $g_{2}$ are (positive) coupling constants associated
with two-body and three-body interactions. $n=n({\bf r})$ is the
local density of exciton and at the mean field level,
\begin{equation}
n(\mathbf{r}) = \sum_{j=1}^{\mathcal{N}} \eta_{j}(E_{j})
|\psi_{j}(\mathbf{r} )|^{2},\label{eq:density}
\end{equation}
where $\mathcal{N}$ denotes the total number of bound
states and $\eta_{j}$ is the corresponding probability function
associated with the energy level $E_{j}$. $\eta_{j}$
satisfies the normalized condition $\sum_{j=1}^{\mathcal{N
}\eta_{j}(E_{j})=1 $. Readers can refer to
Ref.~\onlinecite{Liu2006} for a detailed description of the
model.
In the present model, when macroscopically ordered state forms in
the ring or in the impurity (disorder) potential well, exciton will
distribute over the discrete energy levels. In fact, four sharp
peaks in PL spectra corresponding to the emissions of indirect
exciton states have been observed recently in a so-called "elevated
trap".\cite{high-2008} We attribute the individual localized states
as the confinement due to both the elevated trap potential and the
self-trapped potential (arising from the attractive interaction).
The present model can give a natural and self-consistent
explanation of the particle density dependence of FWHM broadening
and PL energy shift. For low density exciton, the interaction
between exciton is dominated by the attraction. With the increase
of the particle density (by increasing the laser power), the low
temperature exciton cloud will contract irrespective of localized
or delocalized states. This means that the attractive interaction
hampers the exciton motion. At the mean field level, this
corresponds to the increase of the self-trapped potential. The
larger the particle density is, the larger the attractive
interaction is, and the exciton cloud will contract further. At
low temperature, $T < T_{tr}$, uncertainty principle governs the
exciton motion. Consequently, it results in higher mean exciton
energy and larger energy dispersion. Higher mean exciton energy
indicates blue-shift of PL peak and larger energy dispersion means
that PL peak is more broadened (FWHM becomes large).
When the particle density is further increased, three-body repulsion
becomes more important and self-trapped interaction becomes weaker.
The PL peaks will red shift and become sharp again. In fact, with
the increase of the laser power in the low particle density regime,
the phenomenon that PL spectra broaden first and then become sharper
has been observed (see Fig.~3(c) of Ref.~\onlinecite{high-2008}).
This, in addition to the support given by the experimental data in
Ref. \onlinecite{LaiCW2004}, provides a further evidence that the
interaction between exciton is likely the combination of a two-body
attraction and a three-body repulsion. With increasing the laser
power, blue shift of the PL peaks has been reported in Ref.
\onlinecite{Butov1990}. However, the prediction of blue shift
first and then red shift of the PL peak with the increase of the
laser power has not been observed yet! Further experiments are in
demand.
\section{Spatial energy distribution}
\label{Energy spatial distribution in exciton external ring}
In order to study the spatial dependence of the energy distribution
in the external exciton ring, it is necessary to solve the nonlinear
Schr\"{o}dinger equation (\ref{the schrodinger equation})
numerically. In addition to the assumption of interactions, another
important aspect lying in our model is the disequilibrium energy
disequilibrium energy distributions of indirect excitons. In fact,
the indirect excitons in CQW involves both the (complex) energy
relaxation and recombination processes. These have been studied by
several experimental and theoretical groups. \cite{Benisty1991} It
shows that for a relaxation process, when the exciton density is low
($n\ll a_{B}^{2}$, $a_{B}$ is Bohr radius), the effects due to the
exciton-exciton interaction and the exciton-carrier scattering can
be neglected. In this case, the relaxation time is mainly determined
by the scattering of excitons with acoustic phonons.
\cite{Piermarocchi1996} In particular, at low bath temperatures
$(T_{b}<1$ $\mathrm{K})$, this kind of relaxation rate decreases
dramatically due to the so-called "phonon bottleneck" effects.
\cite{Benisty1991}
For the recombination process, on the other hand, because exciton in
the lowest self-trapped level are quantum degenerate, they are
dominated by the stimulated scattering when the occupation number is
more than a critical value. Strong enhancement of the exciton
scattering rate has been observed in the resonantly excited
time-resolved PL experiment \cite{butov:5608}. Therefore, even
though the phonon scattering rate is still larger than the radiative
recombination rate, thermal equilibrium of the system may not be
reached. Essentially the distribution $\eta_{j}(E_{j})$ may deviate
significantly from the usual Boltzmann or Bose like.
\cite{Ivanov1999}
A recent paper (Ref.~\onlinecite{high-2008}) reported that the
observed narrow PL lines of indirect excitons in a disorder
potential correspond to the emission of individual states. It has
been shown that the intensity of PL spectra is roughly in
proportion to the exciton number. When excitons distribute over
some given energy levels, the PL spectra will exhibit strong peaks
corresponding to these levels. Therefore the relative height of
the PL peaks can be interpreted as the exciton distribution
probability associated with the energy level, $\eta_{j}(E_{j})$.
At the high temperature of 10 K, the PL peak intensity of the high
excited state is higher than that of the ground state (see Fig.~2(c)
of Ref.~\onlinecite{high-2008}). This indicates that exciton have
a larger probability distribution in the high excited state. With
reducing the temperature to 5.1 K, the PL peak intensity becomes
lower for the high excited state while it becomes higher for the
other three peaks associated with the two ground states. It seems
that excitons at higher states will relax to lower states when the
temperature is reduced. At the temperature lower than 5.1 K, the
intensities are found to be almost identical for the four peaks. It
indicates that excitons distribute over the four discrete levels
with almost the same probability. Furthermore, at even lower
temperature of 2.7 K, the low-energy peaks become very dim. This
implies that phonon bottleneck effect is in effect for ground state
and low excited states and in this regime, excitons in high excited
states are having difficulty relaxing to the ground states.
From the above analysis, one can generalize $\eta (E_{j})\rightarrow
\eta (E_{j},T)$ to have a better description for the temperature
dependence of particle number distribution of bound states. As
mentioned above, experiment seems suggesting that the weight of
$\eta (E_{j},T)$ will transfer from high-energy states to low-energy
states when $T$ is decreased. It also suggests that, for a given
$T$, the weight of $\eta (E_{j},T)$ is larger in higher-energy
states than that in lower-energy states. With the above in mind, two
other factors are also important in determining the actual form of
$\eta (E_{j},T)$. Firstly, exciton are Bose quasiparticle and at not
too low temperatures, the distribution can be approximated by the
(classical) Boltzman function, $\exp (-E_{j}/T)$, where $T$ is
considered as an effective temperature related to lattice
temperature. The other important factor is the energy dependence of
exciton luminance efficiency. Low-energy excitons is more optical
active than that of high energy exciton. So the low energy excitons
have a high luminance efficiency which is contrary to the
high-energy excitons. Thus low-energy state will have relatively
smaller exciton distribution taking into account the luminance
efficiency. One can use $\exp (E_{j}/E_{0})$ to describe
qualitatively this effect, where $E_{0}$ is an effective energy
scale related to the exciton life time. As a consequence,
temperature and energy dependent exciton distribution is an outcome
of the competition between the above two effects. We then take the
following {\em phenomenological} form for the weight
\begin{eqnarray}
\eta (E_{j},T) &\equiv&C\exp \left[ -\left( E_{j}-\mu \right)
/T\right]
\exp \left[ \left( E_{j}-\mu \right) /E_{0}\right] \nonumber\\
&=&C\exp \left[ \alpha (E_{j}-\mu)\right], \label{eq:eta}
\end{eqnarray}
where $C$ is the normalization factor, $\mu$ is the chemical
potential, and $\alpha \equiv 1/E_{0}-1/T$. Typically $E_{0}<T$ and
hence $\alpha >0$. When temperature and laser power remain
unchanged, $\mu$ and $\alpha$ will remain constantly. While
increasing the temperature, $\alpha$ will increase correspondingly.
Thus $\alpha$ can be considered as an "effective temperature".
Equation (\ref{eq:eta}) basically gives an appropriate temperature
and energy dependent probability density distribution for indirect
exciton. In Fig.~\ref{fig1}, we compare the theoretical curve
[based on Eq.~(\ref{eq:eta})] with the experimental PL spectra
reported in Ref.~\onlinecite{high-2008}. The experimental data
(circles) in Fig.~\ref{fig1} are in arbitrary unit. For
comparison, we have multiplied a constant to the probability
density. It should be mentioned that we have tried various
distribution functions and found that Eq.~(\ref{eq:eta}) gives the
best fitting. In our earlier study (Ref. \onlinecite{Liu2006}), we
have assumed that exciton are distributed at different energy
levels with almost the same probability. Strictly speaking, it
corresponds only to the low temperature limit in Eq.
(\ref{eq:eta}).
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=8cm]{fig1.eps}
\end{center}
\caption{Energy dependent exciton distribution. Solid line is
obtained from the phenomenological formula (\ref{eq:eta}) with
chemical potential $\mu=1.56$ eV and $\alpha=0.025$ eV$^{-1}$.
Circle line corresponds to the exciton PL spectra (in arbitrary
unit) taken from Fig.~2(a) of [\onlinecite{high-2008}]. For
comparison, a constant is multiplied to the solid line.}
\label{fig1}
\end{figure}
Numerically it is convenient to do the following scaling:
$\psi_j(\mathbf{r})/\sqrt{N}\rightarrow \psi_j(\mathbf{r})$,
$Ng_{1}\rightarrow g_{1}$, and $N^{2}g_{2}\rightarrow g_{2}$, such
that Eq.~(\ref{the schrodinger equation}) remains the same form. In
this case, $n(\mathbf{r})$ becomes the probability
density which satisfies the normalization condition $\int _{S}n(\mathbf{r
)dS=1$. After further rescaling $\psi_j(\mathbf{r})\sigma_{\mathrm{PL
}\rightarrow \psi_j(\mathbf{r})$ and $\mathbf{r}/\sigma_{\mathrm{PL
}\rightarrow \mathbf{r}$, Eq.~(\ref{the schrodinger equation}) is reduced to
\begin{equation}
-\frac{1}{2}\nabla^{2}\psi_{j}+(v_{ex}-a_{1}n+a_{2}n^{2})\psi_{j}
=\varepsilon_{j}\psi_{j}, \label{the reduced schrodinger equation}
\end{equation}
where $v_{ex}\equiv V_{ex}/\epsilon$, $a_{1}\equiv g_{1}/\left( \sigma_
\mathrm{PL}}^{2}\epsilon\right)$, $a_{2}\equiv g_{2}/\left( \sigma_{\mathrm
PL}}^{4}\epsilon\right)$, and $\varepsilon_{j}\equiv E_{j}/\epsilon$. Here
\epsilon\equiv{\hbar^{2}}/{m^{\ast}} \sigma_{\mathrm{PL}}^{2}$ with $\sigma_
\mathrm{PL}}$ being the root-mean-square radius of the exciton
cloud observed by photoluminescence. With the above scaling, it is
found
that $a_{1}^{2}/a_{2}=g_{1}^{2}/g_{2}\epsilon=g_{1}^{2}{m^{\ast}} \sigma_
\mathrm{PL}}^{2}/g_{2}\hbar^{2}$, which is a constant for an explicit sample.
Taking into account the experimental facts,\cite{Butov2002b,
yang:033311} two important points should be clarified. (i) The
exciton patterns are fully determined by its self-trapped
interactions. External potential $V_{ex}$ is not the main cause for
complex exciton patterns. Thus we set $v_{ex}=0$ for simplicity.
(ii) When an electron and a hole form an exciton at low temperature,
it is believed that their kinetic energy is low and all exciton are
and all exciton are self-trapped. Particles with energy greater
potential energy can not be bounded in the self-trapped well and
should be ruled out in the calculations.
The mean kinetic energy of excitons can be given by
\begin{eqnarray}
E_k =\int\int dx dy ~E_{k}(x,y), \label{1}
\end{eqnarray} where
$E_{k}(x,y)$ is the mean local (space-dependent) kinetic energy
density. Considering the probability function of each level,
$\eta_j\equiv \eta(E_j,T)$, $E_k$ can also be given by
\begin{eqnarray}
E_k =\sum_{j=1}^{\mathcal{N}} \eta_j E_{kj}, \label{2}
\end{eqnarray}
where $E_{kj}$ is the kinetic energy associated with level $j$. In
fact,
\begin{eqnarray}
E_{kj} &=&-\int\int dx dy ~\psi_j^*(x,y)\nabla^2\psi_j(x,y) \nonumber\\
&=& \int\int dx dy \left(\left|{\partial \psi_j(x,y)\over\partial
x}\right|^2+\left|{\partial \psi_j(x,y)\over\partial
y}\right|^2\right), \label{3}
\end{eqnarray} with $\psi_j(x,y)$ the wave function of level $j$.
In obtaining the 2nd line of Eq.~(\ref{3}), an integration by
parts and a boundary condition that wave function vanishes at the
infinity (bound states) are applied. By comparing
Eqs.~(\ref{1})--(\ref{3}), the mean spatial energy density is
obtained to be
\begin{eqnarray}
E_k(x,y)= \sum_{j=1}^{\mathcal{N}} \eta_j \left(\left|{\partial
\psi_j(x,y)\over\partial x}\right|^2+\left|{\partial
\psi_j(x,y)\over\partial y}\right|^2\right). \label{4}
\end{eqnarray}
Besides, one can also obtain the angle dependent kinetic energy
density via
\begin{equation}
E_{k}\left( \theta \right) =\int\limits_{\left( x,y\right) \in
\left(\theta,\theta +\Delta \theta \right) }dxdy~ E_{k}(x,y)
\end{equation
with $\tan \theta \equiv y/x$. In Fig.~\ref{fig2}(a), we show both
the particle density distribution, $n(\mathbf{r})$, in the exciton
ring and the angle-dependent kinetic energy distribution,
$E_{k}\left( \theta \right)$. Obviously the maxima of $E_{k}\left(
\theta \right) $ are located at the center of the beads.
Fig.~\ref{fig2}(b) shows the spatial dependence of the kinetic
energy distribution, $E_{k}(x,y)$. One interesting feature is that
$E_{k}(x,y)$ is zero at the center of the beads. This can be
elucidated by future experiments of finer resolution.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=8cm]{fig2.eps}
\end{center}
\caption{(a) Formation of the macroscopically ordered state of the
exciton ring at very low temperatures. The corresponding
angle-dependent energy distribution $E_{k}(\theta)$ (solid line)
is also shown. (b) Spatial distribution of the kinetic energy,
$E_{k}(x,y)$. Parameters used are $a_{1}=250$,
$a_{2}=0.001a_{1}^{2}$, effective temperature $\alpha=0.002$, and
chemical potential $\mu=0$. } \label{fig2}
\end{figure}
\section{Temperature dependence of energy: quantitative results}
\label{Temperature dependence of exciton average energy in the exciton
external ring}
In section \ref{Temperature dependence of exciton average energy
in the exciton external ring -- a qualitative analysis}, with the
uncertainty principle, a qualitative analysis was made on the
temperature dependence of exciton average energy in the external
ring. Here we solve the nonlinear Schr\"{o}dinger equation
(\ref{the schrodinger equation}), together with the disequilibrium
distribution $\eta_j(E_j,T)$, to study the temperature dependence
quantitatively. As mentioned before, $\alpha$ can be treated as an
effective temperature in the present case. Temperature
($\alpha$)-dependent exciton ring patterns are shown in
Fig.~\ref{fig3}(a)-(c), while Fig.~\ref{fig3}(d) shows the
nonmonotonic temperature dependence of the total energy,
$E_{k}=\int E_{k}(\theta) d\theta$.
It should be noted that our theory applies only to low
temperatures when excitons are in or near the condensed state.
When temperature is above the critical point, such as
$T_{tr}\simeq 4$ K presented in Ref.~\onlinecite{high-2008},
excitons will transfer from condensed to non-condensed phase.
Above $T_{tr}$ in the non-condensed phase, the exciton liquid may
be ``boiling" and its spatial distribution could change with the
time. At such high temepratures, our theory will break down to
which numerical calculation turns no convergent solutions. In
fact, the exciton pattern shown in Fig.~\ref{fig3}(c) is just a
snapshot of time-dependent exciton number distribution. When
$\alpha\agt 0.08$ in our numerical simulation, the exciton
behavior can be well described by the classical statistical
physics and a linear formula $\varepsilon=k_B T$ is plotted in
Fig.~\ref{fig3}(d) for illustration purpose.
\begin{figure}[ptb]
\begin{center}
\includegraphics[width=8cm]{fig3.eps}
\end{center}
\caption{Formation of the macroscopically ordered state of the
exciton ring at different effective temperatures: (a)
$\alpha=0.002$, (b) $\alpha=0.004$, and (c) $\alpha=0.008$. (d)
Temperature ($\alpha$) dependence of the total energy $E_{k}$. All
other parameters are taken to be the same as those in
Fig.~\ref{fig2}.} \label{fig3}
\end{figure}
A brief summary is in order here. Initially, if exciton are
uniformly distributed on the external ring, attractive interaction
will drive the exciton to approach each others. When the local
density reaches a critical value, the kinetic energies of the
exciton will drive the high-density excitons to diffuse. At the same
time, the repulsive interaction, which increases with the density,
will hinder further increase of the exciton density. As a result,
an array of clusters on the external ring forms. The size of
these clusters is thus determined by three factors, i.e., the
attractive interaction, the three-body repulsive interactions, and
the kinetic energies of the excitons associated with the
temperatures. This is the basic idea behind our model on the
macroscopically ordered states of excitons.
Finally some remarks on the temperature effect are given. When the
bath temperature is low, cooled exciton have relatively low momenta
and the self-trapped interaction is able to confine most of the
excitons. However, in the low momentum case, cooling efficiency is
low while luminous efficiency is high, excitons can not reach the
thermal equilibrium state. At the meantime, due to the competition
between the self-trapped and the kinetic energies, complex exciton
patterns occur (as discussed above). With increasing the
temperature, the exciton can not be fully cooled and correspondingly
self-trapped interaction confines only part of the excitons. The
attractive interaction also cannot compensate the exciton kinetic
energy and excitons will distribute homogenously in a 2D plane. In
this case, the pattern is washed out. If the temperature is higher
than the indirect exciton binding energy $\sim3.5$ $\mathrm{meV},$
\cite{Snoke2004b,szymanska:193305} most of excitons become ionized
and are in a plasma state. No pattern can be observed in this case.
\section{Number distribution with laser induced trap}
\label{Exciton number distribution in laser induced trap}
Similar to the studies of ultracold alkali atoms and molecules,
the laser-induced trapping was proposed and demonstrated for a
highly degenerate Bose gas of exciton. \cite{hammack:227402} An
important advantage of laser trapping is that it is possible to
control the trap in situ by varying the laser intensity in space
and time. A commonly seen phenomenon for highly degenerate
excitons is its annual particle distribution in the laser-induced
trap. We argue that this kind of distribution is similar to that
found in an impurity potential reported in Ref. \onlinecite{LaiCW2004}.
In our previous
work, a detailed discussions was already given to this kind of
trapped exciton distribution.\cite{Liu2006} Since laser induced
trapping can confine higher density exciton, it is hoped that the
experiment can also be done at very high densities to explore the
many-body physics of excitons.
Using the temperature and energy dependent distribution given in
Eq.~(\ref{eq:eta}) together with a two-dimensional pseudopotential
\begin{equation}
v_{ex}(\mathbf{r}) =\left\{
\begin{array}{cc}
\displaystyle -5 ~~ & \mathrm{for}~~r\leq \sigma_{\mathrm{PL}} \\
0 & \mathrm{otherwise,
\end{array}
\right. \label{V}
\end{equation}
we solve the nonlinear Schr\"{o}dinger equation (\ref{the
schrodinger equation}) for exciton number distribution ranging
from low to high densities. The results are presented in
Fig.~\ref{fig4}. As mentioned before, $\sigma_{\mathrm{PL}}$ is
the root-mean-square radius of the exciton PL pattern. To compare
to experiments, we take $a_{1}$ = $15$, $23$, and $35$ with
$a_{2}=0.004a^{2}_{1}$ accordingly. As shown in Fig.~\ref{fig4},
the excitons distribute annularly at low densities and at high
densities, some fine structures develop at the center of the ring,
which is consistent with the experiments. Readers can refer to
Ref.~\onlinecite{Liu2006} for the reason behind the annular
distribution and some related physics.
\begin{figure}[ptb]
\begin{center}
\includegraphics[width=9cm]{fig4.eps}
\end{center}
\caption{The exciton distribution in laser-induced trap for (a) low
($a_1=15$), (b) intermediate ($a_1=23$), and (c) high ($a_1=35$)
densities. The other parameters are taken to be same as those in
Fig.~\ref{fig2}.} \label{fig4}
\end{figure}
\section{Summary}
\label{Summary}
In summary, the uncertainty principle is used to analyze the spatial
and temperature dependence of the average exciton energy
distribution in a macroscopically ordered state of the exciton. the
competition between a two-body attraction, a three-body repulsion,
and the kinetic energy plays a crucial in determining the behaviors
of exciton distributions. Numerical simulation of the corresponding
nonlinear Schr\"odinger equation seems to confirm the analysis.
Nevertheless, the reason of forming macroscopically ordered exciton
states may be more complex than what is expected. Full understanding
of the exciton interactions requires a reliable many-body
calculation beyond the mean-field approximation, which is obviously
not feasible at the moment.
In order to realize the exciton BEC phenomenon, it is in demand to
obtain the exciton with even lower combination rate and shorter
relaxation time.
\begin{acknowledgments}
This work was supported by the
National Basic Research Program of China (Grant No.
2005CB32170X), and National Science Council of Taiwan (Grant No.
96-2112-M-003-008).
\end{acknowledgments}
|
\section{Introduction}
We consider simple coalitional games called \emph{path coalitional games}, in particular \emph{Edge Path Coalitional Games (EPCGs)} and \emph{Vertex Path Coalitional Games (VPCGs)}.
In these games, the players control the edges and the vertices, respectively, and a coalition of players wins if it enables a path from the source $s$ to the sink $t$ and loses otherwise. Both of these coalitional games are natural representations, for which solution concepts such as the Shapley value or the nucleolus represent the amount of payoff the respective edges or vertices deserve for enabling a path from $s$ to $t$. The payoff can indicate the importance of the players or the proportional resource, profit, maintenance or security allocation required at the respective nodes and vertices. This kind of stability analysis is especially crucial if the underlying graph represents a logistics, communication, military, supply-chain or information network~\citep{BP10a,Nebel10a}.
We study the computational complexity of computing important cooperative game theoretic solutions of path coalitional games.
Path coalitional games also have a natural correspondence with two-person zero-sum noncooperative games. In such games, which we term as \emph{path intercept games}, there are two players, the \emph{interceptor} and the \emph{passer}. The problem is to maximize the probability of intercepting a strategically chosen path in an undirected graph. We refer to the path intercept games as \emph{Edge Path Noncooperative Games (EPNGs)} and \emph{Vertex Path Noncooperative Games (VPNGs)}. The pure strategies of the interceptor are the edges $E$ (or vertices $V$) and the set of pure strategies of the \emph{passer} is the set $\mathcal{P}$ which contains all paths from vertex $s$ to vertex $t$. If the edge (or vertex) used by the interceptor intersects with the chosen path, then the interceptor wins and gets payoff $1$. Otherwise, the interceptor loses and gets payoff $0$. Thus, the {\em value} of the game is the greatest probability that the interceptor can guarantee for successfully intercepting the chosen path.
The area of algorithmic cooperative game theory is beset with negative computational results~\citep[see e.g., ]{DeFa08a,EGGW09a,KBBEGRZ08a}.
In this paper, we present positive algorithmic results for \emph{cost-based generalizations} of path coalitional games and their duals. The cost-based generalization of a simple game is a rich and widely-applicable model. For example in the case of edge path coalitional games, each edge charges a certain cost for its services being utilized. A coalition of edges gets a fixed reward for enabling an $s$-$t$ path. It is then natural to examine payoffs which are fair and stable and also manage to transport goods from $s$ to $t$~\citep{FGM00a}. The cost-based generalizations of path coalitional games have significance in logistics, planning and operations research. Similarly, the cost-based generalizations of duals of path coalitional games have natural importance in proposing stable reward schemes to protect strategic assets or blocking intruders in a network.
\paragraph{Contribution:}
\begin{itemize}
\item We use dualization and cost-based generalization to provide a unifying way to model $s$-$t$ connectivity. In doing so, we also identify some interesting connections between coalitional game theory and network interdiction.
\item For the cost-based generalization of path coalitional games and their duals we present the first polynomial-time algorithms to compute and verify least core payoffs. Interestingly, the problem of computing the least core of the dual of vertex path coalitional game was (wrongly) claimed to be NP-hard~\citep{BP10a} and the problem of computing the least core of cost-based generalization of edge path coalitional game was conjectured to be NP-hard
~\citep[page 65, ]{Nebel10a}.
\item We present an algorithmic technique to compute least core payoffs for cost-based generalizations of simple games in any representation. As a corollary, it is shown that there exist polynomial-time algorithms to compute the least core payoffs of cost-based generalizations of \emph{spanning connectivity games} and \emph{weighted voting games} with bounded weights and costs.
\item The least core payoffs of simple path coalitional games are characterized and purely combinatorial polynomial-time algorithms to compute a least core payoff are presented.
\item The nucleolus is a solution concept which is notoriously hard to compute for most interesting coalitional games. A polynomial-time algorithm to compute the nucleolus of edge path coalitional games for undirected series-parallel graphs is presented.
\end{itemize}
\section{Related work}
Network interdiction is the general framework in which weakening of a network by an adversary or fortification of a network by defenders is considered~\cite{SmithLim08a}. Within this body of literature, shortest path interdiction (in which an adversary wants to maximize the length of the $s$-$t$ shortest path in a directed network) is related to our setting of $s$-$t$ path coalitional games. Whereas all the variants of shortest path interdiction problem are NP-hard~\citep{KBBEGRZ08a}, we present positive computational results.
While \emph{Vertex Path Noncooperative Games (VPNGs)} have not been considered in the interdiction literature (to the best of our knowledge), \emph{Edge Path Noncooperative Games (EPNGs)} is equivalent to the two player zero sum games considered in \citep{WashWood95a}. \citet{WashWood95a} studied maxmin strategies in EPNGs. Our cooperative game formulation helps us in proposing equilibrium refinements such as the nucleolus (which corresponds to a unique refinement of the maxmin strategy of the corresponding path intercept games) and other cooperative-game solution concepts such as the Shapley value.
The coalitional model, especially the cost-based generalizations, helps us reason about more elaborate security settings in which incentives, money, and cooperation of agents is involved.
The comparison between a simple coalitional game and its natural noncooperative version is similar in spirit to \citep{ALPS09a} where spanning trees are considered.
The general definition of cost-based generalization of a simple game is inspired by \citet{FGM00a} and \citet{BP10a} where cost-based versions of specific graph-based simple games are considered.
\citet{FGM00a} examined conditions for the non-emptiness of the core of \emph{`shortest path games'} (which are equivalent to cost-based generalization of EPCGs). However, computational problems such as computing a least core payoff were not considered.
Different variants of EPCGs were considered under the umbrella of `shortest path games' in \citep{Nebel10a} but either the complexity of core-based relaxations is not examined or the complexity of computing least core solutions was left open and in fact conjectured to be computationally hard~\citep[page 65, ]{Nebel10a}.
Similarly, in \citep{BP10a}, it is claimed that the least core of the dual of VPCGs is NP-hard to compute. We disprove the claim in \citep{BP10a} and present a polynomial-time algorithm to solve a generalization of the same problem for this game as well as three other games on \emph{any graph}.
A variant of EPCGs was also considered in \citep{NiRo01a} but the focus was on strategy-proof mechanisms rather than stability issues.
\citet{DIN99a} consider a different type of $s$-$t$ connectivity game which is balanced.
The $s$-$t$ path connectivity setting also has natural links with network reliability where the goal is to compute the probability that there exists a connected path.
However the network reliability literature does not consider strategic settings and certainly has no equivalent concepts such as the least core and the nucleolus etc.
\section{Preliminaries}
In this section, we first define the path coalitional games and path intercept games and then consider suitable game-theoretic solution concepts for these games.
\subsection{Games}
We begin with the formal definition of a \emph{coalitional game}.
\begin{definition}[Coalitional games]
A \emph{coalitional game} is a pair $(N,v)$ where $N=\{1,\ldots, n\}$ is a set of players and $v:2^N \rightarrow \mathbb{R}_+$ is a \emph{characteristic or valuation function} that associates with each coalition $S\subseteq N$ a payoff $v(S)$ where $v(\varnothing)=0$
\footnote{Throughout the paper, we assume $0\in\mathbb R_+$. }
A coalitional game~$(N,v)$ is \emph{monotonic} when it satisfies the property that $v(S)\leq v(T)$ if $S \subseteq T$.
\end{definition}
Throughout the paper, when we refer to a coalitional game, we assume such a coalitional game with transferable utility. For the sake of brevity, we will sometimes refer to the game $(N,v)$ as simply~$v$.
\begin{definition}[Simple game]
A \emph{simple game} is a monotonic coalitional game $(N,v)$ with $v:2^N \rightarrow \{0,1\}$ such that $v(\varnothing)=0$ and $v(N)=1$. A coalition $S \subseteq N$ is \emph{winning} if $v(S)=1$ and \emph{losing} if $v(S)=0$. A \emph{minimal winning coalition} of a simple game $v$ is a winning coalition in which defection of any player makes the coalition losing.
\end{definition}
We now define the following two \emph{path coalitional games}.
\begin{definition}[Path coalitional games]
For an unweighted directed/undirected graph, $G=(V\cup\{s,t\},E)$,
\begin{itemize}
\item the corresponding \emph{Edge Path Coalitional Game (EPCG)} is a simple coalitional game $(N,v)$ such that $N=E$ and for a $S\subseteq N$, $v(S)=1$ if and only if $S$ admits an $s$-$t$ path.
\item the corresponding \emph{Vertex Path Coalitional Game (VPCG)} is a simple coalitional game $(N,v)$ such that $N=V$ and for a $S\subseteq N$, $v(S)=1$ if and only if $S$ admits an $s$-$t$ path.
\end{itemize}
\end{definition}
\begin{definition}[Dual of a game]
For a game $G=(N,v)$, the corresponding \emph{dual} game $G^D=(N,v^D)$ can be defined in the following way:
$v^D(S)=v(N)-v(N\setminus S)$ for all $S\subseteq N$.
\end{definition}
For both EPCG and VPCG, the corresponding duals $EPCG^D$ and $VPCG^D$ can be defined. It will be seen that $VPCG^D$ is equivalent to a well-studied coalitional game.
For a simple game, we can define a game which is the \emph{cost-based generalization}.
\begin{definition}[Cost-based generalization]
For a given simple game $G=(N,v)$ we can define a \emph{cost-based generalization} $C$-$G=(N,v^c)$ based on \emph{cost vector} $c=(c_1,\ldots, c_{|N|})\in {\mathbb{R}_{+}}^{|N|}$ and \emph{reward} $r\in \mathbb{R}_{+}$.
For a coalition $S\subseteq N$, the value of the $v^c(S)=r-\min_{S'\subseteq S, v(S')=1}(\sum_{i\in S'}c_i)$ if $v(S)=1$ and $v^c(S)=0$ if $v(S)=0$.
\end{definition}
The intuition of a cost-based generalization is that each player demands some cost for its services being utilized and a coalition of players $S$ get a reward $r$ only if it is winning and gets the job done. The coalition also incurs a cost of $\min_{S'\subseteq S, v(S')=1}(\sum_{i\in S'}c_i)$ when it pools resources to get the job done. Based on this formulation, we can define Edge Path Coalitional Games with costs C-EPCG and Vertex Path coalitional games with costs, C-VPCG. It is easy to see that for a game C-G, if $r=1$ and the costs are all zero, then C-G is equivalent to G.
\begin{observation}
C-ESPG is equivalent to the \emph{value shortest path game (VSPG)} in \citep{Nebel10a}. ${VPCG}^D$ is equivalent to the \emph{simple path disruption game} in \citep{BP10a}.
\end{observation}
We now define the following two \emph{path intercept games}.
\begin{definition}[Path intercept games]
For an unweighted directed graph, $G=(V\cup\{s,t\},E)$,
the corresponding \emph{Edge Path Noncooperative Game (EPNG)} is a noncooperative game with two players, the \emph{interceptor} and the \emph{passer}. The pure strategies of the interceptor are the edges $E$ and the pure strategies of the \emph{passer} is set $\mathcal{P}$ which contains all paths from vertex $s$ to vertex $t$. If the edge used by the interceptor intersects with the chosen path, then the interceptor wins and gets payoff $1$. Otherwise it loses and gets payoff $0$.
\emph{Vertex Path Noncooperative Games (VPNGs)} have an analogous definition to EPNGs except that the pure strategies of the interceptor are the vertices $V$ and that if the vertex used by the interceptor intersects with the chosen path, then the interceptor wins and gets payoff $1$.
\end{definition}
Both EPCG and VPNG can be generalized to the case with detection probabilities where the probability that the passer moving through edge $e$ (or vertex $v$) will be detected if the interceptor inspects $e$ (or $v$) is $p_e$ (or $p_v$ respectively).
\subsection{Cooperative Solutions}
A cooperative game solution consists of a distribution of the value of the grand coalition over the players. Formally speaking, a solution associates with each cooperative game~$(N,v)$ a set of \emph{payoff vectors $(x_1,\ldots,x_n)\in\mathbb R^N$} such that $\sum_{i\in N}x_i=v(N)$, where $x_i$ denotes player~$i$'s share of~$v(N)$. Such efficient payoff vectors are also called \emph{preimputations}.
As such, solution concepts formalize the notions of fair and stable payoff vectors. In what follows, we use notation similar to that of \citet{EGGW09a}.
Given a cooperative game~$(N,v)$ and payoff vector $x=(x_1,...,x_n)$, the \emph{excess
of a coalition $S$ with respect to $x$} is defined by \[e(x,S)=x(S)-v(S)\text,
\]
where $x(S)=\sum_{i\in S}x_i$.
We are now in a position to define one of the most fundamental solution concepts of cooperative game theory, viz., the core.
\begin{definition}[Core]
A payoff vector $x=(x_1,\ldots,x_n)$ is in the \emph{core} of a cooperative game~$(N,v)$ if and only for all $S\subseteq N$, $e(x,S) \geq 0$.
\end{definition}
A core payoff vector guarantees that each coalition gets at least what it could gain on its own. The core is a desirable solution concept, but, unfortunately it is empty for many games. Games which have a non-empty core are called \emph{balanced}. The possibility of the core being empty led to the development of the \emph{$\epsilon$-core}~\citep{ShSh66a} and the \emph{least core}~\cite{MPS79a}.
\index{$\epsilon$-core}
\begin{definition}[Least core]
For $\epsilon>0$, a payoff vector $x$ is in the \emph{$\epsilon$-core} if for all $S\subseteq N$, $e(x,S)\geq-\epsilon$. The payoff vector~$x$ is in the \emph{least core} if it is in the $\epsilon$-core for the smallest~$\epsilon$ for which the $\epsilon$-core is non-empty. We will denote by $-\epsilon_1(v)$, the minimum excess of any least core payoff vector of $(N,v)$.
\end{definition}
It is easy to see from the definition of the least core, that it is the solution of the following linear program (LP):
\begin{equation}
\label{LC-LP}
\begin{array}{ll}
\min & \epsilon \\
\text{s.t.} & x(S)\geq v(S)-\epsilon\ \ \text{for all}\
S\subseteq N,\,\\
&\epsilon\geq 0, x_i\geq 0\ \text{for all}\
i\in N ,\\
& \sum_{i=1,\ldots,n} x_i = v(N)\ .\\
\end{array}
\end{equation}
The nucleolus is a special payoff vector which is in the core if the core is non-empty and is otherwise a member of the least core. The \emph{excess vector} of a payoff vector $x$, is the vector $(e(x,S_1),...,e(x,S_{2^n}))$ where $e(x,S_1)\leq e(x,S_2)\leq \ldots \leq e(x,S_{2^n})$.
\begin{definition}[Nucleolus]
A payoff vector $x$ such that $x_i\geq v(\{i\})$ for all $i\in N$ and $x$ has lexicographically the largest excess vector is called the \emph{nucleolus}.
\end{definition}
The nucleolus is unique and always exists as long as $v(S)=0$ for all singleton coalitions~\citep{Sc69a}.
\eat{
\begin{definition}[Banzhaf value]
A player $i$ is \emph{critical} in a coalition $S$ when $S \in W$ and $S \setminus \{i\} \notin W$.
For each $i \in N$, we denote the number of \emph{swings} or the number of coalitions in which $i$ is critical in game $v$ by the \emph{Banzhaf value} ${{\eta}_{i}}(v)$.
\end{definition}
Intuitively, the Banzhaf value is the number of coalitions in which a player plays a critical role and the Shapley-Shubik index is the proportion of permutations for which a player is \emph{pivotal}. For a permutation $\pi$ of $N$, the $\pi(i)$th player is pivotal if coalition $\{\pi(1),\ldots, \pi(i-1)\}$ is losing but coalition $\{\pi(1),\ldots, \pi(i)\}$ is winning.
\begin{definition}[Shapley value]
The \emph{Shapley value} of $i$ is the function $\varphi$ defined by
$\varphi_i(v)=\frac{\sum_{X \subseteq N} (|X|-1)!(n-|X|)!(v(X)- v(X-\{i\}))}{n!}.$
\end{definition}
}
\section{Least core of path coalitional game variants}
Before considering other computational issues, we notice that the value of a coalition in EPCGs and VPCGs can be computed in polynomial time. For a coalition $S$ in a EPCG/C-VPCG, use \emph{Depth First Search} to check whether $s$ and $t$ are connected in a graph restricted to $S$. If not, then $v(S)=0$. Otherwise, $v(S)$ is equal to $1$.
Our first observation is that in all games EPCG, VPCG, $EPCG^D$ and $VPCG^D$, the core can be empty. In fact, the following proposition characterizes when the core of these games is non-empty:
\begin{proposition}
The core of
\begin{itemize}
\item EPCG is non-empty if and only if there exists an edge, the removal of which disconnects $s$ and $t$.
\item VPCG is non-empty if and only if there exists a vertex, the removal of which disconnects $s$ and $t$.
\item $EPCG^D$ is non-empty if and only if there exists an $(s,t)$ edge.
\item (\citet{BP10a}) $VPCG^D$ is non-empty if and only if there exists a vertex $x$ such that $(s,x)$ and $(x,t)$ are edges in the graph.
\end{itemize}
\end{proposition}
\begin{proof}
All cases follow directly from the fact that in a simple monotone game, the core is non-empty if and only if there exists a vetoer, i.e., a player $i\in N$ such that $v(N\setminus \{i\})=0$~\citep[see e.g., ][]{EGGW09a}. For the dual games, note the following.
Let $(N,v^d)$ be the dual game and let $x$ be a player such that $v(x)=1$. We want to show that player $x$ is a vetoer in $(N,v^d)$ i.e., $v^d(N\setminus x)=0$.
We know that $v(N)=1$. Then, by definition of dual, $v^d(N\setminus x)=v(N)-v(x)=1-1=0$. Thus x is a vetoer in $(N,v^d)$.
\end{proof}
Since the core can be empty, the least core payoff assumes more importance.
We will first present a general positive result (Theorem~\ref{th:main}) regarding the computation of least core payoff for cost-based generalizations of simple games.
For a simple game $G=(N,v)$ and $(x_1, \ldots, x_{N})\in {\mathbb{R}_{+}}^{|N|}$, denote by $G^x$ game $G$ in which each player $i\in N$ has weight $x_i$. Then, a \emph{minimum weight winning coalition} of $G^x$ a winning coalition $S$ such that $x(S)$ is minimal.
\begin{theorem}\label{th:main}
For a simple game $G=(N,v)$, assume that there exists an algorithm which for a given
weight vector $(x_1, \ldots, x_{N})\in {\mathbb{R}_{+}}^{|N|}$, computes in polynomial time a minimum weight winning coalition of $G^x$. Then, a least core payoff of a cost-based generalization of $G$ can be computed and verified in polynomial time.
\end{theorem}
\begin{proof}
We will denote the algorithm in the statement as Algorithm $A$.
Consider C-G$=(N,v^c)$ be the cost-based generalization of $(N,v)$ with associates cost vector $c=(c_1,\ldots, c_{|N|})\in {\mathbb{R}_{+}}^{|N|}$ and reward $r\in \mathbb{R}_{+}$.
In order to compute a least core payoff of C-G, we consider the least core LP for C-G.
The size of the linear program~\eqref{LC-LP} is exponential in the
size of game C-G, with an inequality for every subset
of players. However, this linear program can be solved using the
ellipsoid method and a separation oracle,
which verifies in polynomial time
whether a solution is feasible or returns a violated
constraint~\cite{Sch04a}. We now demonstrate how algorithm $A$ can be used to construct the separation oracle for the least core LP of C-G.
A candidate solution for the least core payoff is an efficient payoff $x=(x_1,\ldots, x_{|N|})$ such that $x(N)=v^c(N)=r-\min_{S'\subseteq N, v(S')=1}(\sum_{i\in S'}c_i)=r-$ weight of the minimum weight winning coalition of $G^c$. Since $c=(c_1,\ldots, c_{|N|})\in {\mathbb{R}_{+}}^{|N|}$, Algorithm $A$ can be used to compute $\min_{S'\subseteq S, v(S')=1}(\sum_{i\in S'}c_i)$ and therefore $x(N)$. Now that $x(N)$ is known, a separation oracle for the least core LP considers different candidate solutions $x$ such that $x(N)$ is constant.
For a candidate solution $x=(x_1,\ldots, x_{|N|})$, where $x(N)=v(N)=r-$ minimum cost of a winning coalition in C-G, construct the weighted function $x'=(x_1',\ldots, x_{|N|}')$ such that $x_i'=x_i+c_i$ for all $i\in N$. Since $x_i\geq 0$ and $c_i\geq 0$ for all $i\in N$, $x_i'\geq 0$ for all $i\in N$. Therefore, we can use algorithm $A$ to compute a minimum weight winning coalition $S^*$ of $G^{x'}$.
Now the claim is that $S^*$ is a coalition with the minimum excess of C-G with respect to payoff $x$ and that $x'(S^*)-r$ is the minimum of excess of game C-G with respect to payoff $x$. Note that $e(x,S^*)=x(S^*)-v^c(S^*)\leq x(N)-v^c(S^*)=x(N)-(r- \min_{S''\subseteq S^*, v(S'')=1}c(S''))\leq x(N)-(r- \min_{S''\subseteq N, v(S'')=1}c(S''))= x(N)-v^c(N)= 0.$
For the sake of contradiction, assume that there is a coalition $S'$ which has the minimum excess with respect to $x$ in game C-G such that $S'$ is not a minimum weight winning coalition of $G^{x'}$. Then, either $S'$ is not winning or is winning but not a minimum weight winning coalition. If $S'$ is losing, then $e(x,S')=x(S')-v^c(S')=x(S')\geq 0$. Since $e(x,S^*)\leq 0$, $S'$ does not have smaller excess that $S^*$ in game C-G with respect to payoff $x$.
In the second case, assume that $v^c(S')=1$ but $x'(S')>x'(S^*)$. Without loss of generality, $S'$ is a minimal winning coalition. If it were not, then we prove that there exists an $S''\subset S'$ such that $S''$ is a minimal winning coalition and
$e(x,S'')\leq e(x,S')$. If $v(S'')=v(S')$, then we are already done as $x(S'')\leq x(S)'$. Assume that $v(S')<v(S'')$. Then, there exists a minimal winning coalition $S'''\subset S'$ such that $c(S''')<c(S'')$. But then it must be that $x(S''')+c(S''')\geq x(S'')+c(S'')$ because if it were not, then $e(x,S''')\leq e(x,S'')$. Thus, we have established that $S'$ is a minimal winning coalition without loss of generality. Since $x(S')+c(S')=x'(S')> x'(S^*)=x(S^*)+c(S^*)$, therefore $e(x,S')=x(S')-(r-c(S'))=x(S')+c(S')-r=x'(S')-r>x'(S^*)-r=e(x,S^*)$.
We can use the known algorithm $A$ to compute the winning coalition $S^*$ with the smallest total weight $x'(S^*)$.
If we have $x'(S^*)=x(S^*)+c(S^*)-r\geq -\epsilon$, then $x(S)-v^c(S)\geq -\epsilon$ for all $S\subseteq N$. Therefore, $x$ is feasible. Otherwise, the constraint $x(S^*)-v^c(S^*)\geq -\epsilon$ is violated. This completes our argument that a polynomial-time separation oracle for the least core LP of the C-G can be constructed.
A payoff $x=(x_1,\ldots, x_{|N|})$ can be verified if it is in the $\epsilon$-core by using the separation oracle. Since the minimum excess $-\epsilon_1$ of the least core payoff can be computed, therefore the separation oracle can also be used directly to check if the given payoff is in the least core.
\end{proof}
\begin{corollary}\label{cor:scg}
A least core can be computed for cost-based generalizations of the following games: \emph{spanning connectivity games}~\citep{ALPS09a} and \emph{weighted voting games~\citep{EGGW09a} with bounded weights and also bounded costs}.
\end{corollary}
\begin{proof}
For a spanning connectivity game $G$, there exists an algorithm which for a given
weight vector $(x_1, \ldots, x_{N})\in {\mathbb{R}_{+}}^{|N|}$, computes in polynomial time a minimum weight winning coalition of $G^x$~\citep[Proposition 5, ][]{ABH10a}.
For weighted voting games with weights represented in unary, there exists an algorithm which for a given
weight vector $(x_1, \ldots, x_{N})\in {\mathbb{R}_{+}}^{|N|}$, computes in polynomial time a minimum weight winning coalition of $G^x$~\citep[Theorem 5, ][]{EGGW09a}. Since the algorithm works only for weighted voting games with small weights, the algorithm can be used as a separation oracle for the least core of cost-based generalization of weighted voting games only if the associated cost vector is also represented in unary.
\end{proof}
From the proof of Theorem~\ref{th:main} and Corollary~\ref{cor:scg} it is evident that if there the separation oracle to compute the least core LP of a simple game $G$ can also be used as a separation oracle to compute the least core LP of a cost-based generalization of $G$. We say that the representation of a coalitional game $(N,v)$ is \emph{as compact} as the cost function $c=(c_1,\ldots, c_{N})\in {\mathbb{R}_{+}}^{|N|}$, if the following condition holds: if cardinal values used in the representations of $(N,v)$ are in unary, then $c$ is also represented in unary.
\begin{observation}\label{th:main2}
Let $G=(N,v)$ be the underlying simple game and C-G$=(N,v^c)$ be the cost-based generalization of $(N,v)$. Assume that the representation of $(N,v)$ is as compact as the cost function $c$. Then, if there exists a polynomial-time separation oracle for the least core LP of the underlying simple game, then a least core payoff of the cost-based generalization can be computed and verified in polynomial time.
\end{observation}
We now apply Theorem~\ref{th:main} to path coalitional games.
\begin{theorem}\label{th:escg-costs-lc}
There exist polynomial-time algorithms to compute and verify least core payoffs of cost-based generalizations of Edge Path Coalitional Games (C-EPCGs) and Edge Path Coalitional Games (C-EPCGs) for both directed and undirected graphs.
\end{theorem}
\begin{proof}
We use Theorem~\ref{th:main} to prove the statement.\\
\noindent
C-EPCGs: For a C-EPCG $G$, it is sufficient to show that for a weight vector, $x=(x_1,\ldots, x_{|E|})$, we can compute a minimum weight winning coalition of $G^x$. Each player (edge) $i$ has a weight $x_i$ and the minimum weight winning coalition is an $s$-$t$ simple path $P$ with the smallest weight, that is the shortest $s$-$t$ path. Use \emph{Dijkstra's Shortest Path Algorithm} to compute the shortest path $P$ from $s$ to $t$ in graph $G^{x}$ and then the minimum weight winning coalition is $E(P)$, the edges used in path $P$.\\
\noindent
C-VPCGs: For a C-VPCG $G$, it is sufficient to show that for a weight vector, $x=(x_1,\ldots, x_{|V|})$, we can compute a minimum weight winning coalition of $G^x$. Each player (node) $i$ has a weight $x_i$ and the minimum excess coalition is an $s$-$t$ simple path $P$ with the smallest weight, that is the shortest vertex $s$-$t$ path.
Then compute the shortest vertex weighted path $P$ from $s$ to $t$ in graph $G^{x}$ and then the minimum weight winning coalition is $V(P)$. Dijkstra's Shortest Path Algorithm can be used to compute the shortest vertex weighted path as follows.
The problem can be reduced to the classic shortest path problem in the following way:
duplicate each vertex (apart from $s$ and $t$) with one getting all ingoing edges, and the other getting all the outgoing edges, add an internal edge between them with the node weight as the edge weight. Use the algorithm to compute the shortest vertex path $P$ from $s$ to $t$ in graph $G^{x}$.
\end{proof}
A least core payoff of a coalitional game is not necessarily a least core payoff of the dual game.
Therefore, we require new algorithms to compute the least core of dual coalitional path games.
\begin{theorem}\label{th:dual-epcg-costs-lc}
There exist polynomial-time algorithms to compute and verify least core payoffs of C-$EPCG^D$s and C-$VPCG^D$s for both directed and undirected graphs.
\end{theorem}
\begin{proof}
We utilize Theorem~\ref{th:main} to prove the statement.\\
\noindent
C-$EPCG^D$: For a C-$EPCG^D$ $G$, it is sufficient to show that for a weight vector, $x=(x_1,\ldots, x_{|E|})$, we can compute a minimum weight winning coalition of $G^x$. Each player (edge) $i$ has a weight $x_i$ and the minimum weight winning coalition is an $s$-$t$ cut $P$ with the smallest weight. Use the maximum network flow algorithm~\citep[Chapter 27, ]{CLRS01a} to compute the minimum weight edge $s$-$t$ cut $C$ in graph $G^x$. This gives us the minimal winning coalition $C$ with the minimum weight.\\
\noindent
C-$VPCG^D$: For a C-$VPCG^D$ $G$, it is sufficient to show that for a weight vector, $x=(x_1,\ldots, x_{|V|})$, we can compute a minimum weight winning coalition of of $G^x$. Each player (node) $i$ has a weight $x_i$ and the minimum weight winning coalition is a minimum weight $s$-$t$ vertex cut. Then compute the minimum weight $s$-$t$ vertex cut in graph $G^{x}$ and then the minimum weight winning coalition is $V(P)$. It is known that
the minimum weight vertex $s$-$t$ cut can be computed in polynomial time for directed graphs by standard network-flow methods. The network flow method to compute the minimum edge $s$-$t$ cut can be used to compute the minimum vertex $s$-$t$ cut as following.
The problem can be reduced to the problem of min weight $s$-$t$ edge cut of an edge weight directed graph in the following way:
duplicate each vertex (apart from $s$ and $t$) with one getting all incoming edges, and the other getting all the outgoing edges, add an internal edge between them with the node weight as the edge weight. Set the weight of all original edges as infinite (sufficiently large). We use existing algorithms to compute the minimum weight vertex $s$-$t$ cut to construct the separation oracle for the
C-${VPCG}^D$ least core LP.\end{proof}
We note that if instead of using $s$-$t$ connectivity settings, we consider more than two terminals then some problems such as {\sc In-$\epsilon$-Core} become NP-hard. This follows from the fact that computing a min cut for more than two terminals is NP-hard.
\eat{
A natural solution when the core is non-empty is to increase the reward to the grand coalition to ensure that all players work together and the grand coalition is stable. An external agent who increases the reward would rather increase the reward minimally to ensure that the core of the resultant game is non-empty. This concept is termed as the \emph{cost of stability} and was introduced in \citep{BMZRR09a}. From our results it follows that the cost of stability can also be computed in polynomial time.
\begin{theorem}\label{th:cos-easy}
The cost of stability of C-EPCG, C-VPCG, C-${EPCG}^D$ and C-${VPCG}^D$ can be computed in polynomial time.
\end{theorem}
\begin{proof}
We already saw that separation oracles for the least core LPs for EPCG, VPCG, ${EPCG}^D$ and ${VPCG}^D$ can be solved in polynomial time. Since all these algorithms rely on the construction of a polynomial-time separation oracle, by Observation~1~\citep{ABH10a}, the cost of stability of all these games can be computed in polynomial time.\end{proof}
}
\section{A closer look at path coalitional games without costs}
In this section, we take a closer look at simple path coalitional games without costs.
We will refer to the minimum size of an $s$-$t$ cut of a unweighted graph as $c_E$ if we refer to edge cuts and as $c_V$ if we refer to vertex cuts. Then we have the following theorem:
\begin{theorem}[Characterization of path coalitional games without costs]\label{th:lc-charac}
Consider an EPCG $G_{EPCG}$ and a VPCG $G_{{VPCG}}$. Then $\epsilon_1(G_{EPCG})=1-1/c_E$ and $\epsilon_1(G_{{VPCG}})=1-1/c_V$. Moreover, there are combinatorial polynomial-time algorithms to compute and verify a least core payoff of EPCGs and VPCGs.
\end{theorem}
\begin{proof}
Consider an EPNG based on graph $G$ with detection probabilities $(p_1,\ldots, p_{|E|})$.
Let $\Delta(A)$ denote set of mixed strategies
(probability distributions) on a finite set $A$.
The equilibrium or {\em maxmin} strategies of the interceptor
are
the solutions
$\{ x\in \Delta(E)\
|\ \sum_{v\in P} x_e\cdot p_e\ge val(G)\ \text{for all}\ P\in \mathcal{P}
\}$
to the following linear program,
which has the optimal value $val(G)$.
\begin{equation}
\label{e-LPmaxmin}
\begin{array}{ll}
\max & \alpha \\
\text{s.t.} & \sum_{e\in P} x_e\cdot p_e \geq \alpha\
\text{for all}\ P\in \mathcal{P}\ ,\\
& x \in \Delta(E)\ .\\
\end{array}
\end{equation}
We notice that if $p_e=1$ for $e\in E$, then LP~\eqref{e-LPmaxmin} is equivalent to the least core LP for EPCGs. It is clear that maxmin strategy $x$ of EPNG where $p_e=1$ for all $e\in E$ is equivalent to the least core payoff of EPCG corresponding to $G$.
LP~\eqref{e-LPmaxmin} is equivalent to LP 1 in \citep{WashWood95a} if $p_e$ is set to $1$ for each edge in both LPs. This demonstrates that computing maxmin strategies of EPNG is equivalent to computing least core payoffs of EPCGs.
\citet{WashWood95a} conclude that maxmin strategy is obtained by constructing a graph $G^{c'}$ where $c'_e=1/p_e$ and then computing the minimum weight $s$-$t$ cut $S$. Each edge $e\in S$ is then given interdiction probability proportional to $c_e=1/p_e$.
It follows that if $p_e=1$ for all $e\in E$, then $c'_e=1/p_e$, and the minimum weight $s$-$t$ cut $S$ of $G^{c'}$ is simply the min cardinality $s$-$t$ cut $S$ of $G$.
A maxmin strategy $x$ of the interceptor, each edge in $S$ is inspected with probability $1/|S|=1/c_E$.
Therefore for EPCG corresponding to $G$, the payoff of each simple $s$-$t$ path or equivalently minimum winning coalition has payoff $1/c_E$ and the minimum excess $-\epsilon_1$ of the EPCG is $1/c_E-1$.
We note that a similar analysis holds for VPCGs.
\end{proof}
Theorem~\ref{th:lc-charac} helps to give a correspondence between EPCG and EPNG and also between VPCG and VPNG. We note that there is no such correspondence between, for example EPNG with detection probabilities and C-EPCG.
Theorem~\ref{th:lc-charac} helps us formulate combinatorial algorithms to compute the least core of EPCGs and VPCGs (without costs).
The problem of computing a least core payoff reduces to computing a minimum cardinality edge cut (or vertex cut) of the graph and uniformly distributing the probability over the minimum cut. \emph{Such least core payoffs are the extreme points of the least core convex polytope and in fact any other least core payoff is a convex combination of the extreme points.}
\eat{
\begin{table}[t]
\footnotesize
\centering
\begin{tabular}{lll}
\toprule
Graph&EPCG/VPCG&EPNG/VPNG\\ \midrule
E/V&Players&pure strategy of interceptor\\
$s$-$t$ paths&MWCs&pure strategies of passer\\
graph weights&payoff&interceptor's strategy\\
$x$&LC payoff&maxmin strategy\\
\bottomrule
\end{tabular}
\caption{Path coalitional games \& Path intercept games}
\label{SCG-wiretap}
\end{table}
}
Our demonstrated connection of EPNGs to the corresponding coalitional EPCG in the proof of Theorem~\ref{th:lc-charac} helps examine refinements of the maxmin strategies such as the nucleolus.
The nucleolus of a coalitional game is the unique and arguably the fairest solution concept which is guaranteed to lie in the core if the core is non-empty. The interpretation of the nucleolus in the non-cooperative setting is the maxmin strategy of the passer which not only minimizes the number of pure best responses of the interceptor but also maximizes the potential extra payoff if the interceptor does not choose the optimal strategy.
Computing the nucleolus is a notoriously hard problem and only a handful of non-trivial coalitional games are known for which the nucleolus can be computed efficiently~\citep[see \eg][]{ALPS09a}.
We will show that for certain graph classes like series-parallel graphs, the nucleolus strategy can be computed in polynomial time (Theorem~\ref{th:epcg-nucleolus-sp}). Series-parallel graphs are an especially useful class of graphs because they are present in many settings such as electrical networks, urban grid lay-outs etc.
\begin{definition}[Series-parallel graph]
Let $G=(V,E)$ be a graph with source $s$ and sink $t$. Then $G$ is a \emph{series-parallel graph} if it may be reduced to $K_2$(a two vertex clique) by a sequence of the following operations:
\begin{enumerate}
\item replacement of a pair of parallel edges by a single edge that connects their common endpoints;
\item replacement of a pair of edges incident to a vertex of degree 2 other than $s$ or $t$ by a single edge so that $2$ degree vertices get removed.
\end{enumerate}
\end{definition}
Denote the set of edge min cuts of a graph $G$ by $\mathcal{C}(G)$. Denote by $C_e(G)$ the set $\{S\in \mathcal{C}(G) \mid e'\in S \}$.
\begin{lemma}\label{lemma:epcg}
For an undirected series-parallel graph $G=(V\cup\{s,t\},E)$, let $x$ be a least core payoff of the corresponding EPCG and let $e\in E$ be such that $C_e(G)=\emptyset$. Then $x_e=0$.
\end{lemma}
\begin{proof}
Let $e=(a,b)\in E$ be such that $C_e(G)=\emptyset$ and assume for contradiction that there is a least core payoff of EPCG for $G$ such that $x_e>0$.
Let the graph component in series with and left of $e$ be $G_1$, the graph component in series with and right of $e$ be $G_2$, the graph component in parallel and above $e$ be $G_3$ and the graph component in parallel below $e$ be $G_4$. Since $C_e(G)=\emptyset$ there exists no edge $e'\in G_3 \cup G_4$ such that $C_e'(G)>0$. Now assume that the mincut value of $G$ is $c^*$. The mincut $C$ with size $c^*$ must either be in $G_1$ or $G_2$. We also know that since $x$ is a least core payoff, the length of the shortest $s$-$t$ path in $G$ is $1/c^*$. We show that if $x_e>0$, then a transfer of payoff from certain edge in $G_3\cup G_4 \cup \{e\}$ increases the minimum excess, thereby showing that $x$ is not a least core payoff.
If there exists no shortest $a$-$b$ path which includes $e$, then we know that $e$ is present in no coalition which gets the minimum excess. Therefore $e$ can donate its payoff uniformly to $C$ and increase the minimum excess by $x_e/c^*$. Now assume that $e$ is in one of the shortest $a$-$b$ paths. Clearly, this is not the only simple $a$-$b$ paths because if this were the case then $e$ would be a bridge be one of the mincuts. We know that mincut value of $G_3\cup G_4 \cup \{e\}$ is more than $c^*$. Let $S$ be the minimum cut of $G_3\cup G_4 \cup \{e\}$. We know that $|S|>|C|=c^*$. Then, we can show that the minimum excess of $x$ increases if $x(S)$ is distributed uniformly over $C$. Each shortest $s$-$t$ path if $G^x$ has to pass one edge in $C$ and one edge in $S$. The the weight of each edge in $C$ has increased by $x(S)/|S|$ and the length of the shortest $a$-$b$ has decreased by $x(S)/|S|$, the excess increases exactly by positive value $x(s)/|C|-x(S)/|S|$ without decreasing any other excesses.
\end{proof}
\begin{theorem}\label{th:epcg-nucleolus-sp}
The nucleolus of EPCGs for undirected series-parallel graphs can be computed in polynomial time.
\end{theorem}
\begin{proof}
We show that the problem of computing the nucleolus of EPCGs of undirected series-parallel graphs reduces to computing the parallel-series decomposition of the graph.
There are known standard algorithms to identify and decompose series-parallel graphs~\citep[see e.g., ]{HeYe87a}.
The reduction is based on an inductive argument in which if we know the nucleolus of two graphs $G'$ and $G''$, then we can also compute in polynomial time the nucleolus of the graph made by connecting $G'$ and $G''$ in series or parallel.
The proof by induction is as follows:
\paragraph{Base case:} The base case is trivial. In any graph $G$ with a single edge $e$ connecting $s$ and $t$, the (pre)nucleolus $x$ gives payoff $1$ to $e$.
\paragraph{Induction:} Our induction involves two cases: attaching two graph components in series and parallel. Consider two series-parallel undirected graphs $G'$ and $G''$ and assume we already know that their nucleoli are $x'$ and $x''$ respectively. We will show that computing the nucleolus of $G$ formed by joining $G'$ and $G''$ in series and parallel is polynomial-time easy.
\begin{enumerate}
\item
Assume we attach $G'$ and $G''$ in series to obtain $G$. Let the size of any edge mincut be $c'$ and any edge mincut be $c''$. If $c'< c''$, then by Lemma~\ref{lemma:epcg}, there is no advantage of giving payoff to any edges in $G''$.
Therefore, the nucleolus of $G'$ is equal to the nucleolus of $G$ and we are done.
Assume that $c'=c''$.
We recall that the nucleolus satisfies \emph{anonymity} and \emph{covariance}~\citep{PeSu03a,Sn95a}.
Then due to Lemma~\ref{lemma:epcg} and covariance and anonymity property of the nucleolus, we have $x=(\alpha{x'},(1-\alpha)x'')$ where $0<\alpha<1$.
Let $m'$ and $m''$ be the smallest non-zero payoff of a player in $x'$ and $x''$ respectively.
We then show that $x=(\alpha{x'},(1-\alpha)x'')$ is the nucleolus if $\alpha$ has the unique value for which $m'(\alpha)=m''(1-\alpha)$, i.e., $\alpha=m''/(m'+m'')$. If this were true, then $x=(m''/(m'+m''){x'},(1-m''/(m'+m''))x'')$. In this case, the minimum excess for $x$ is $1/c'-1$ and the number of coalitions achieving this in $G$ is $|A|\times 2^{|B|}$ where $A$ is the set number of simple paths in $G^x$ and $B=\{e\in E(G)\mid C_e(G)=\emptyset\}$. We also know that the value of the second minimum excess is $1/c'-1+ (m'\cdot m'')/(m'+m'')$.
Now assume that there exists another payoff $y=(\alpha{x'},(1-\alpha)x'')$ for some $\alpha \neq m''/(m'+m'')$ such that $y$ has a lexicographically greater excess vector than $x$. Clearly $y$ is a least core payoff of $G$.
Then the minimum excess for $x$ is $1/c'-1$ and the number of coalition achieving this in $G$ is still $|A|\times 2^{|B|}$.
However the second minimum excess for $y$ is less than $1/c'-1+ (m'\cdot m'')/(m'+m'')$. Therefore, $y$ has a smaller lexicographical excess vector than $x$ which is a contradiction.
\item Consider two series-parallel undirected graphs $G'$ and $G''$ and assume we attach them in parallel to obtain $G$. Let the size of any edge mincut of $G'$ be $c'$ and the size of any edge mincut of $G''$ be $c''$. Both the mincut values can be computed in polynomial time for a graph.
We know that the size of mincut of $G$ is $c'+c''$.
Then due to Lemma~\ref{lemma:epcg} and covariance and anonymity properties of the nucleolus, we know that $x=(\alpha{x'},(1-\alpha)x'')$ where $0<\alpha<1$. We then show that $x=(\alpha{x'},(1-\alpha)x'')$ is the nucleolus if alpha has the unique value $c'/(c'+c'')$. Since the size of a mincut of $G$ is $c'+c''$, every least core payoff $y$ of is such that $G^y$ has shortest path $1/(c'+c'')$. We want that every shortest $s$-$t$ path which passes from $G$ to have length $1/(c'+c'')$. This is only possible if $\alpha=c'/(c'+c'')$.
\end{enumerate}
\end{proof}
We conjecture that a similar approach may help construct a polynomial-time algorithm to compute the nucleolus of VPCGs for series-parallel graphs.
\eat{
\begin{observation}\label{tree-path-games}
The Banzhaf indices, Banzhaf values and Shapley-Shubik indices and nucleolus of EPCG, $EPCG^D$, VPCG and $VPCG^D$ can be computed trivially. We note that in a tree, there is a unique $s$-$t$ path $p$. All the edges/vertices in $p$ are equi-important and the rest of the edges/vertices are dummies. Therefore the payoff is distributed uniformly over the non-dummy players.
\end{observation}
}
\eat{
\section{Power indices of coalitional path games}
Power indices are popular solution concepts of cooperative games. Even in the noncooperative settings where instead of one interceptor, there are multiple interceptors, power indices measure which nodes and edges play a more pivotal role in connecting $s$-$t$ and therefore suggest in what proportion, safety levels must be established on those nodes and edges. Computing Banzhaf values and Shapley values is hard in most settings. \citet{BP10a} proved that computing Banzhaf values of ${VPCG}^D$ is \#P-complete. The result was obtained by a reduction from a counting version of {\sc NAE-SAT}. Similarly, \citet{Nebel10a} proved that computing the Banzhaf values and Shapley values of EPCGs is \#P-complete by a reduction from {\sc $s$-$t$ Node Connectedness}~\citep{Val79a}.
We show that computing the Banzhaf values and Shapley values of all variants is \#P-complete.
\begin{definition}[Banzhaf value]
A player $i$ is \emph{critical} in a coalition $S$ when $S \in W$ and $S \setminus \{i\} \notin W$.
For each $i \in N$, we denote the number of \emph{swings} or the number of coalitions in which $i$ is critical in game $v$ by the \emph{Banzhaf value} ${{\eta}_{i}}(v)$.
\end{definition}
Intuitively, the Banzhaf value is the number of coalitions in which a player plays a critical role and the Shapley-Shubik index is the proportion of permutations for which a player is \emph{pivotal}. For a permutation $\pi$ of $N$, the $\pi(i)$th player is pivotal if coalition $\{\pi(1),\ldots, \pi(i-1)\}$ is losing but coalition $\{\pi(1),\ldots, \pi(i)\}$ is winning.
\begin{definition}[Shapley value]
The \emph{Shapley value} of $i$ is the function $\varphi$ defined by
$\varphi_i(v)=\frac{\sum_{X \subseteq N} (|X|-1)!(n-|X|)!(v(X)- v(X-\{i\}))}{n!}.$
\end{definition}
\begin{proposition}\label{th:powerindices-edge-pathgames}
Computing the Shapley values and Banzhaf values of the VPCG, ${VPCG}^D$, and ${EPCG}^D$ is \#P-complete.
\end{proposition}
\begin{proof}
\citet{BP10a} proved that computing the Banzhaf value of ${VPCG}^D$ is \#P-complete.
We now prove that computing the Shapley values of the ${VPCG}^D$ $(N,v)$ is \#P-complete.
A representation of a simple game is considered \emph{reasonable} if, for a simple game $(N,v)$, the new game $(N\cup\{x\},v')$ where $v(S)=1$ if and only if $v'(S\cup\{x\})=1$, can also be represented with only a polynomial blowup. \citet{AzizThesis09} proved in Theorem 3.29 that for a simple game with a reasonable representation, if computing the Banzhaf values is \#P-complete, then computing the Shapley values is \#P-complete. It is known that computing the Banzhaf value of the ${VPCG}^D$ is \#P-complete. We use the theorem of \citep{AzizThesis09} to prove that computing the Shapley values of the ${VPCG}^D$ is \#P-complete. Since computing the Banzhaf value of the ${VPCG}^D$ is \#P-complete, it is sufficient to prove that ${VPCG}^D$ has a reasonable representation. For a graph $G=(V,E)$ construct a graph $G'=(V',E')$ such that $V'=V\cup\{x\}$ and $E'=E\cup \{(s,x), (x,t)\}$. Then, we know that $v(S)=1$ if and only if $v'(S\cup\{x\})=1$.
We prove that computing the Shapley values and Banzhaf values of the VPCG $(N,v)$ is \#P-complete. It is known that for a simple game $(N,v)$ if the Banzhaf value of player $i$ is $\eta_i(N,v)$ then the Banzhaf value of $i$ in the dual of $(N,v)$ is $\eta_i(N,v^d)=\eta_i(N,v)$ (Theorem 5, \citep{DuSh79a}). Therefore, it immediately follows that computing the Banzhaf values of the VPCG is \#P-complete. Similarly, \citet{Fun94a} showed in Lemma~2.7 that for a cooperative game $(N,v)$ and player $i$, $\varphi_i((N,v))=\varphi_i((N,v^d))$. Therefore, computing the Shapley value of the VPCG is \#P-complete.
We know that the Shapley value and Banzhaf value of a player is invariant in the dual of the simple game\citep{DuSh79a,Fun94a}.
Since computing the Banzhaf values and Shapley values of EPCG is \#P-complete, it follows that computing the corresponding values for the $EPCG^D$ is also \#P-complete.
\end{proof}
We note that EPCG, VPCG, ${EPCG}^D$ and ${VPCG}^D$ can be generalized to games in which instead of being concerned about the connectivity of two terminals, there are more than two terminals. Therefore, our hardness results for the two terminal case are stronger than the hardness results for more elaborate models (with possibly multiple terminals). A natural question is to identify classes of graphs for which power indices can be computed efficiently. We have some good news in this regard:
\begin{theorem}\label{th:sp-epcg-easy}
Banzhaf values of the Edge Path and Dual Edge Path Coalitional Game can be computed in polynomial time for series-parallel undirected graphs and acyclic directed graphs.
\end{theorem}
\begin{proof}
Consider the problem {\sc 2-Terminal Reliability} where for a graph $G$ and specified vertices $s$ and $t$, each edge has a certain probability of being operational. The question is to compute the overall probability that $s$ and $t$ are connected. The problem {\sc 2-Terminal Reliability} is hard in general but can be computed in polynomial time for the following graph classes: series-parallel graphs;.
We now show that existing algorithms to solve {\sc 2-Terminal Reliability} for restricted graph classes can be used in a black-box fashion to compute Banzhaf values of the EPCG.
Set the probability that each edge is operational to 0.5. In that case the {\sc 2-Terminal Reliability} of $G$ is equal to the number of winning of winning coalitions $\omega(N_G,v_G)$. Consider the graph $G$ where the probability of
edge $e$ being operational is set to $1$ whereas the
probability of other edges being operational is set to
$0.5$. Then the 2-terminal reliability of the graph is equal to $\omega_e(N_G,v_G)$ the number of winning coalitions which includes edge $e$. We know that $\eta_i(G)=2\omega_i(G)-\omega(G)$ where $\omega_i(G)$ denotes the number of winning coalitions in $G$ which contain player $i$ ~\citep{DuSh79a}. Therefore the Banzhaf values of EPCG and thereby the $EPCG^D$ can be computed in polynomial time.
\end{proof}
Observe that the Banzhaf indices, Banzhaf values and Shapley-Shubik indices and nucleolus of EPCG, $EPCG^D$, VPCG and $VPCG^D$ can be computed trivially. We note that in a tree, there is a unique $s$-$t$ path $p$. All the edges/vertices in $p$ are equi-important and the rest of the edges/vertices are dummies. Therefore the payoff is distributed uniformly over the non-dummy players.
}
\eat{
\begin{table}[t]
\small
\centering
\begin{tabular}{llcccccc}
\toprule
Graph&Game&LC&nucleolus\\
general&C-EPCG&P~(Th.~\ref{th:escg-costs-lc})&?\\
general&C-VPCG&P~(Th.~\ref{th:escg-costs-lc})&?&\\
general&C-${EPCG}^D$&P~(Th.~\ref{th:dual-epcg-costs-lc})&?\\
general&C-${VPCG}^D$&P~(Th.~\ref{th:dual-epcg-costs-lc})&?\\
\midrule
series-parallel&EPCG&P~(Th.~\ref{th:escg-costs-lc})&P~(Th.~\ref{th:epcg-nucleolus-sp})\\
\bottomrule
\end{tabular}
\caption{Summary of results}
\label{table:path-games-summary}
\end{table}
}
\section{Conclusion}
Path coalitional games provide a simple yet rich framework to model strategic settings in the area of network security and logistics.
In this paper we analyzed different generalizations and variants of path coalitional games and classified the computational complexity of computing different cooperative and noncooperative game solutions.\footnote{This material is based upon work supported by the Deutsche Forschungsgemeinschaft under grants BR-2312/6-1 (within the European Science Foundation's EUROCORES program LogICCC) and BR~2312/7-1. The research of Troels Bjerre S{\o}rensen was supported by EPSRC award EP/G064679/1.
We thank Mingyu Xiao, Evangelia Pyrga, and Markus Brill for useful comments.}
One key result was a general method to compute least core payoffs of cost-based generalizations of simple games.
Many of our positive results are based on separation oracles and linear programs. It will be interesting to see if there are purely combinatorial algorithms for the same problems. Apart from the EPCGs on series-parallel graphs, the complexity of computing the nucleolus is open for all other games. For all variants of path coalitional games, we assumed that each edge/vertex is owned by a separate player. It will be interesting to see if our positive results can be extended to the more general scenario where a single player may own more than one edge or vertex.
\eat{
\section*{Acknowledgements}
This material is based upon work supported by the Deutsche Forschungsgemeinschaft under grants BR-2312/6-1 (within the European Science Foundation's EUROCORES program LogICCC) and BR~2312/7-1. The research of Troels Bjerre S{\o}rensen was supported by EPSRC award EP/G064679/1.
We thank Mingyu Xiao and Evangelia Pyrga for useful comments.
}
\def\bibfont{\small}
|
\section{Introduction}
\setcounter{equation}{0}
Studying the new particles that will be produced at the LHC will likely be a non-trivial task. In many scenarios the particles of the new sector are charged under a new parity that makes the lightest such particle a dark matter candidate. The new particles will then predominantly be produced in pairs, giving multiple decay products some of which will be undetectable (including at least two dark matter particles). The reconstruction of such events with many objects and missing energy in the final states will be complicated and often ambiguous. It may therefore be useful to also look at signals arising from the near-threshold formation and annihilation of QCD bound states of the new particles. The bound states behave as resonances which annihilate primarily into just two particles (or jets) and no missing energy, so the analysis and the interpretation of the signal are much more straightforward than in the more conventional channels. In particular, detecting the peak in the invariant mass distribution of any of the possible annihilation channels provides a direct measurement of the particle mass. The disadvantages of the bound state signals are their relatively small cross sections and the fact that the dominant annihilation mode is into dijets (which have large QCD background), with only small branching ratios for annihilation into cleaner signals such as $\gamma\gamma$. However, whenever the bound state signals are observable, they can provide an entirely independent method for characterizing or verifying the properties of the new particles.
In this paper we study how the spins of the new particles are reflected in the properties of their bound states and the resulting signals at the LHC. As an example, we compare the bound states of the level-$1$ Kaluza-Klein (KK) modes in the universal extra dimensions model (UED) with bound states of the superpartners in the minimal supersymmetric extension of the standard model (MSSM). In both scenarios, the new particles are charged in the same way under the standard model gauge groups, and the masses of the particles are to a large extent free parameters. The fundamental difference between the two scenarios is the spin of the new particles. The KK modes in UED have the same spins as the standard model particles (which are the zero modes), while the spins of the superpartners in MSSM are different. However, the collider signatures of the decay products of a pair of level-1 KK modes of UED and a pair of analogous MSSM superpartners are unfortunately very similar and it is often challenging to determine the spin~\cite{Smillie:2005ar,Datta:2005zs,Barr:2005dz,Alves:2006df,Athanasiou:2006ef,Athanasiou:2006hv,Wang:2006hk,Kilic:2007zk,Csaki:2007xm,Burns:2008cp,Gedalia:2009ym,Cheng:2010yy,MoortgatPick:2011ix}. It is therefore interesting to study whether and when the bound state annihilation signals can give information about the spin.
While much is known about bound states of the MSSM particles (see~\cite{Kats:2009bv} and references therein), the bound states in UED models have not been explored, with the exception of bound states of KK quarks in the context of a lepton collider~\cite{Carone:2003ms,Fabiano:2008xk}. Here we will study the various possible bound states of colored particles of UED in the context of the LHC and compare them to the corresponding bound states in the MSSM. We will see as we go along that once the spin of the particle is specified many of the results are largely model-independent since they are determined by gauge interactions. Thus our comparisons between particles of different spins (in the same color representation) will hold even more generally.
We start in section~\ref{sec-setup} by discussing the UED model in the context of our study and reviewing the general formalism for bound state computations. In sections~\ref{sec-KKgluonium}--\ref{sec-diKKquarks-KKqKKg} we present the production cross sections, branching ratios and angular distributions for the various bound states and compare signals obtained from UED and MSSM. In section~\ref{sec-simulation} we simulate the bound state signals and the relevant backgrounds in the dijet, $b\bar b$, $t\bar t$, $\gamma\gamma$ and $e^+e^-$ channels and estimate the LHC reach for the various cases. We summarize our conclusions in section~\ref{sec-conclusions}.
\section{Setup\label{sec-setup}}
\setcounter{equation}{0}
\subsection{Particle spectrum of UED}
In the simplest UED scenario~\cite{Appelquist:2000nn} (for a review, see~\cite{Hooper:2007qk}), all the standard model fields are propagating in a single extra dimension of size $R \sim \mbox{TeV}^{-1}$. The right- and left-handed standard model fermions are represented by separate fields in 5 dimensions, and each has two chiralities when reduced to 4 dimensions. To restrict each fermion zero-mode to a single chirality, the extra dimension is assumed to be an $S^1/\mathbb{Z}_2$ orbifold. The zero modes of the unwanted chiralities of the fermions and the 5th components of the gauge fields are then projected out by declaring them to be odd under the $\mathbb{Z}_2$ orbifold symmetry. The KK modes of the right- and left-handed quarks will be denoted by $\kkqR$ and $\kkqL$, or collectively by $q^\star$. We will refer to $\kkqR$ and $\kkqL$ as right- and left-handed KK quarks, even though each one of them is a full Dirac fermion. (Similarly, we will refer to the MSSM partners of the right- and left-handed quarks, as right- and left-handed squarks.) All the standard model interactions are contained in the bulk Lagrangian. Since in 5 dimensions the gauge, Yukawa and quartic-Higgs couplings have negative mass dimension, this is an effective theory with a UV cutoff $\Lambda$ which is assumed to be above the compactification scale: $\Lambda > 1/R$.
In the 4-dimensional description, the theory contains towers of KK modes for each standard model particle, with the mass of the $n$-th mode given (at tree level) by
\be
m_n^2 = \frac{n^2}{R^2} + m_0^2
\ee
where $m_0$ is the mass of the zero mode (that is the standard model particle itself). As a result, all the KK modes at a particular level $n$ are approximately degenerate, with masses ordered like in the standard model. Loop corrections shift the masses~\cite{Cheng:2002ab,Cheng:2002iz}, and the resulting spectrum of the first KK level has the KK gluon as the heaviest particle, followed by the KK quarks, and then the KK excitations of the electroweak gauge bosons, the Higgs, and the leptons. The lightest KK particle (LKP), which is typically one of the neutral KK gauge bosons, is a dark matter candidate~\cite{Servant:2002aq,Cheng:2002ej}. At tree level, the LKP is stable because of momentum conservation in the extra dimension. When loops are taken into account it remains stable because the theory retains KK parity under which odd-$n$ modes are charged. The KK parity also prevents tree-level contributions to the electroweak observables, thus allowing a compactification scale $1/R$ as low as a few hundred GeV. For our purposes it is important that due to KK parity the level-1 KK particles will be produced in pairs, and thus they can form bound states.
In general, the theory also includes boundary terms (on the orbifold fixed points) whose coefficients are free parameters. The boundary terms can be assumed to be symmetric under the interchange of the two orbifold fixed points, and then the theory still preserves KK parity. However, the mass spectrum of the KK modes can be affected dramatically. Even if boundary terms are absent at a particular scale, they are generated by the renormalization group running~\cite{Georgi:2000ks,Cheng:2002ab,Cheng:2002iz}, and the resulting corrections to the masses are of the order of the loop corrections. The spectra of~\cite{Cheng:2002ab,Cheng:2002iz} described above were obtained with the simplifying assumption that the coefficients of the boundary terms vanish at a specific cutoff scale. However, as has been shown in~\cite{Carena:2002me} for the case of the gauge kinetic term, even when the coefficient of the boundary term is not much larger than the expectation from na\"{\i}ve dimensional analysis, there is a large effect on the spectrum of the KK modes. Similarly, the effects of boundary kinetic and mass terms for a massive scalar field have been analyzed in~\cite{Flacke:2008ne}, where the results were applied to the electroweak sector and it was found that the identity of the LKP was sensitive to the boundary terms.
In both UED and MSSM, annihilation may or may not be the dominant decay mode of the various bound states, depending on the intrinsic decay rates of the constituent particles, which in turn depend on the mass spectrum of the model. Since the theoretical and experimental constraints on the possible mass spectra are a question that is decoupled from the bound state analysis, we will leave it out of the scope of this paper. We will assume that the bound states decay predominantly by annihilation, but the reader should remember that if the constituent particles are not sufficiently long-lived, the cross sections will need to be multiplied by appropriate branching ratios based on the annihilation rates that we compute here and the model-dependent single-particle decay rates. Note, however, that annihilation branching ratios close to $1$ are not implausible. In particular, in the MSSM there exist various motivated scenarios in which this happens to be true for the gluinonium~\cite{Kats:2009bv} or the stoponium~\cite{Martin:2008sv}. Or looking at this from a different perspective, the observation or non-observation of bound state signals can give us certain information about the spectrum of the model.
In UED, the spectrum of the level-1 KK modes plays a crucial role in determining which of them are sufficiently long-lived for their bound states to decay by annihilation rather than by the decays of the constituent particles. If the assumption of~\cite{Cheng:2002ab,Cheng:2002iz} that the boundary terms vanish at the UV cutoff were true, the KK gluon would have strong two-body decays into a KK quark and an antiquark, and the KK quarks could decay electroweakly into a KK electroweak gauge boson and a quark. In this case, similarly to what happens in much of the parameter space of the MSSM~\cite{Kats:2009bv}, the branching ratios for annihilation will be small. However, we believe (although without constructing explicit examples) that the presence of boundary terms can change the spectrum in ways that would make these decays kinematically forbidden for some of the particles.
For example, if the KK gluon $\kkg$ becomes lighter than the KK quarks $\kkq$, its 2-body decays will be replaced by 3-body decays into a KK electroweak gauge boson, a quark and an antiquark through diagrams involving an off-shell KK quark, a process suppressed by the electroweak coupling and the KK quark mass. This is analogous to the decay of the gluino into a neutralino, quark and antiquark in MSSM scenarios with squarks heavier than the gluino, which easily makes the gluino sufficiently stable for our purposes~\cite{Kats:2009bv}. Similarly to the case of the gluino, there is no need for an unnaturally large gap between the KK gluon and KK quark masses in order for the annihilation to dominate. For example, suppose that the dominant decay process is $\kkg \to q \bar q \kkB$, where $\kkB$ is the KK hypercharge gauge boson. Then the rate is
\be
\Gamma_\kkg \simeq \frac{11}{45\pi}\,\alpha_s\,\frac{\alpha}{\cos^2\theta_W} \l(1 - \frac{m_\kkB}{m_\kkg}\r)^5 \l(\frac{m_\kkg}{m_\kkq}\r)^4 m_\kkg
\sim 10^{-4}\, \l(1 - \frac{m_\kkB}{m_\kkg}\r)^5 \l(\frac{m_\kkg}{m_\kkq}\r)^4 m_\kkg
\label{KKg-decay}
\ee
where for simplicity we assumed $m_\kkq \gg m_\kkg \approx m_\kkB$.\footnote{The assumption $m_\kkq \gg m_\kkg$ is not essential for the validity of this estimate. We have checked that for $m_\kkq = 2m_\kkg$, for example, the exact result differs from (\ref{KKg-decay}) by less than a factor of $2$.} On the other hand, the bound state annihilation rates are
\be
\Gamma_{\rm annih} \sim \alpha_s^5\, m_\kkg \sim 10^{-5}\, m_\kkg
\ee
which can easily dominate over $2\Gamma_{\kkg}$, especially when we include the numerical prefactors that appear in the exact expressions for $\Gamma_{\rm annih}$ (see appendix~\ref{app-rates-KKgKKg}), which are as large as ${\cal O}\,(100)$ for color-singlet bound states, primarily due to multiple powers of color factors.
The collider signatures of the 3-body decays of the KK gluon or gluino have been studied in~\cite{Csaki:2007xm} in an attempt to distinguish between UED and MSSM. One of the goals of the present paper is to find out to what extent detecting the annihilation decays of KK gluonia (which are bound states of two KK gluons) compared to gluinonia (bound states of two gluinos) can provide an additional method for distinguishing between UED and MSSM and determining the properties of the underlying particles. While gluinonia have been studied extensively in the literature, our paper is the first study of KK gluonia.
Another object of our study is KK-quarkonia, which are KK quark-KK antiquark bound states. These have been studied in~\cite{Carone:2003ms,Fabiano:2008xk} in the context of their production and detection at a lepton collider. It was found that the spectrum of~\cite{Cheng:2002ab,Cheng:2002iz} does allow the KK quarks to be sufficiently stable for forming KK-quarkonia, but not for the annihilation decays to dominate over the single KK quark decays.\footnote{It was claimed in~\cite{Carone:2003ms} that KK top quarks can be very stable so that their annihilation decays dominate, but this was incorrect, as explained in~\cite{DePree:2005yv}.} Including boundary terms can probably change these results in either direction, and we are particularly interested in the situation in which some of the KK quarks are more long-lived so that the branching ratio for annihilation is enhanced.
Other than that, it is useful to note that many of our results are largely determined by gauge interactions alone and depend only on the properties of the binding particles rather than the full spectrum of the model and are therefore valid much more generally than just for MSSM or UED. It is therefore useful to study all the possible bound states in these models, even if some of the binding particles are not sufficiently stable in generic MSSM or UED scenarios. With this motivation in mind, our study covers all the possible bound states of pair-produced colored particles in MSSM and UED.
\subsection{Bound state formalism}
\begin{table}[t]
$$\begin{array}{|c|c|c|c|c|}\hline
\, \mbox{UED} \,&\, \mbox{MSSM} \,&\, \mbox{binding} \,&\, \mbox{non-binding} \,\\\hline
\, gg \to \kkg \kkg \,&\, gg \to \go\go \,&\, \vv{1}, \vv{8_S}, \vv{8_A} \,&\, \vv{10}, \vv{\bar{10}}, \vv{27} \,\\
\, q\bar q \to \kkg \kkg \,&\, q\bar q \to \go\go \,&\, \vv{1}, \vv{8_S}, \vv{8_A} \,&\, \,\\\hline
\, gg \to \kkq \bar \kkq \,&\, gg \to \sq\sq^\ast \,&\, \vv{1} \,&\, \vv{8} \,\\
\, q\bar q \to \kkq \bar \kkq \,&\, q\bar q \to \sq\sq^\ast \,&\, \vv{1} \,&\, \vv{8} \,\\\hline
\, qq \to \kkq \kkq \,&\, q q \to \sq\sq \,&\, \vv{\bar 3} \,&\, \vv{6} \,\\\hline
\, qg \to \kkq \kkg \,&\, q g \to \sq\go \,&\, \vv{3}, \vv{\bar 6} \,&\, \vv{15} \,\\\hline
\end{array}$$
\caption{Pair production processes and the color representations of the pair. Two color-octets can form an octet with either a symmetric ($\propto d^{abc}$, denoted $\vv{8_S}$) or antisymmetric ($\propto f^{abc}$, $\vv{8_A}$) wave function. The Clebsch-Gordan coefficients for all the color decompositions can be found in~\cite{Beneke:2009rj}. We have not explicitly listed processes obtained by replacing particles by antiparticles.}
\label{tab-processes}
$$\begin{array}{|c|c|c|c|}\hline
\,\mbox{UED} \,&\,\mbox{MSSM} \,&\, \mbox{SU(3)} \,&\, C \,\\\hline
(\kkg\kkg) \,&\, (\go\go) \,&\, \vv{1} \,&\, 3 \,\\
\,&\, \,&\, \vv{8} \,&\, 3/2 \,\\\hline
(\kkq\bar\kkq) \,&\, (\sq\sq^\ast) \,&\, \vv{1} \,&\, 4/3 \,\\\hline
(\kkq\kkq), (\bar\kkq\,\bar\kkq)\,&\, (\sq\sq), (\sq^\ast\sq^\ast) \,&\, \vv{\bar 3}, \vv{3} \,&\, 2/3 \,\\\hline
(\kkq\kkg), (\bar\kkq\kkg) \,&\, (\sq\go), (\sq^\ast\go) \,&\, \vv{3}, \vv{\bar 3} \,&\, 3/2 \,\\
\,&\, \,&\, \vv{\bar 6}, \vv{6} \,&\, 1/2 \,\\\hline
\end{array}$$
\caption{Bound states in UED and MSSM, their color representations and the strength of their potential~(\ref{V(r)}). The possible spins of these bound states will be discussed later.}
\label{tab-C}
\end{table}
Assuming that the masses of the pair of particles satisfy $m_1, m_2 \gg \Lambda_{\rm QCD}$, their dynamics can be described by the single-gluon exchange potential with the Coulomb-like form
\be
V(r) = -C\frac{\bar\alpha_s}{r}
\label{V(r)}
\ee
where $\bar\alpha_s$ denotes the strong coupling constant evaluated self-consistently at the scale of the inverse Bohr radius $a_0^{-1} = C\bar\alpha_s \mu$, where $\mu \equiv m_1m_2/(m_1+m_2)$ (while the plain $\alpha_s$ will be reserved for its value at the scale of $m_1$ and $m_2$) and
\be
C = \frac{1}{2}\l(C_1 + C_2 - C_{(12)}\r)
\label{color-factor}
\ee
where $C_1$ and $C_2$ are the quadratic Casimirs of the color representations of the two particles and $C_{(12)}$ of the bound state. For bound states considered in this paper, the values of $\bar\alpha_s$ are between $0.11$ and $0.15$, so using (\ref{V(r)}) should be a good approximation, even though subleading corrections may have sizeable effects and it would be desirable to compute them, along with higher-order corrections to the production and annihilation processes, in a future work. Table~\ref{tab-processes} lists all the possible pair production processes in UED and MSSM and specifies in each case which of the color representations of the pair have an attractive potential and which do not, based on the sign of $C$. The values of $C$ for all the attractive configurations are given in table~\ref{tab-C}. In cases where the bound states are colored, they will further hadronize with ordinary quarks or gluons to become color-neutral (if they are sufficiently long-lived for this to happen). However, these processes will be happening at much larger distance scales than both the production and annihilation of the bound states and we therefore ignore their effects.
One can get the matrix elements for bound state production and annihilation by representing the bound state as a superposition of states of two free particles with momenta distributed according to the bound state wavefunction $\psi(\vv{r})$. For an $S$-wave bound state, neglecting the dependence of the short-distance process on the momenta, it can be easily shown that the matrix element between the bound state and the particles from which the pair is produced is given by~\cite{Peskin-Schroeder,Novikov:1977dq}
\be
\cM_{\rm bound} = \frac{\psi(\vv{0})}{\sqrt{2\mu}}\,\cM_0
\label{times-wavefunction}
\ee
where $\cM_0$ is the matrix element describing the production of the free constituent particles at threshold and $\mu$ is their reduced mass. The cross section will then be proportional to
\be
\l|\psi(\vv{0})\r|^2 = \frac{C^3 \bar\alpha_s^3 (2\mu)^3}{8\pi}
\label{wavefunction}
.\ee
More specifically, the bound state production cross section can be written in terms of the near-threshold production cross section of the pair of particles $\hat\sigma_0(\hat s)$ as\footnote{While (\ref{times-wavefunction})--(\ref{wavefunction}) assume the binding particles to be distinct, the relation (\ref{sigma-bound}), with $|\psi(\vv{0})|^2$ given by (\ref{wavefunction}), is valid also if they are identical. Later, in (\ref{ann-rate}), $|\psi(\vv{0})|^2$ again refers to the expression (\ref{wavefunction}) even in case of identical particles.}
\be
\hat\sigma_{\rm bound}(\hat s) = \frac{8\pi}{2\mu}\,\frac{\hat\sigma_0(\hat s)}{\beta(\hat s)}\, |\psi(\vv{0})|^2\,2\pi\,\delta(\hat s-M^2)
\label{sigma-bound}
\ee
where $M \simeq m_1 + m_2$ is the mass of the bound state and
\be
\beta(\hat s) \equiv \sqrt{\frac{2\mu}{(M/2)^2} \l(\sqrt{\hat s} - m_1 - m_2\r)}
\ee
is the factor that makes the continuum production cross section $\hat\sigma_0(\hat s)$ vanish at threshold. For $m_1 = m_2$, $\beta$ is the velocity of each particle in the center-of-mass frame.
The annihilation rate of the bound state into two mass-$m_0$ particles is
\bea
\Gamma &=& \frac{|\psi(\vv{0})|^2}{64\pi\, m_1 m_2} \sqrt{1 - \frac{m_0^2}{(M/2)^2}} \int_0^\pi d\theta\,\sin\theta\,\sum|\cM_0(\theta)|^2\nn\\
&&\q\l(\times\frac{1}{2}\;\begin{array}{c}\mbox{ for identical}\\ \mbox{ bound particles}\end{array}\r)\l(\times\frac{1}{2}\;\begin{array}{c}\mbox{ for identical}\\ \mbox{ final particles}\end{array}\r)
\label{ann-rate}
\eea
where $\cM_0(\theta)$ is the matrix element between the pair of particles that form the bound state (with a particular spin and color representation) and the annihilation products, for any polarization and color state of the bound state, and the sum is over the colors and polarizations of the products.
\section{KK gluonia vs. gluinonia\label{sec-KKgluonium}}
\setcounter{equation}{0}
In this section we will study bound states of pairs of color octets: KK gluons $\kkg$ (spin $1$) in UED and gluinos $\go$ (spin $1/2$) in MSSM. Later in this section we will also look at bound states of spin $0$ color octets.
It is useful to classify the various possible bound states in the model according to their spin $J$, color representation, parity $P$, charge conjugation $C$, and the transformation under the chiral $U(1)$ symmetry of the quarks. This allows one to immediately determine the possible production and annihilation channels and describe the corresponding processes using effective interaction vertices involving the bound state and Standard Model fields. These effective vertices can then be used to simulate the bound state processes in an event generator. The coefficients of the vertices can be determined by matching to the relevant Feynman diagrams of the free pair of particles (times the wavefunction at the origin and the other factors).
\begin{table}[t]
$$\begin{array}{c|c|c|c}
\mbox{} & \,\mbox{color} \,&\, J^{PC} \,&\, \mbox{can couple to} \,\\\hline
(\kkg\kkg) & \vv{1}, \vv{8_S} & 0^{++} & G^{\rho\sigma}G_{\rho\sigma},\, \bar t t \\
& \vv{8_A} & 1^{+-} & \epsilon^{\mu\nu\rho\sigma}i\bar t \l[\gamma_\rho,\gamma_\sigma\r] t \\
& \vv{1}, \vv{8_S} & 2^{++} & \; G^{\rho\mu}G_{\rho\nu},\, G^{\rho\sigma}D_\mu D_\nu G_{\rho\sigma},\, i\bar q \gamma^\mu D_\nu q \\\hline
(\go\go) & \vv{1}, \vv{8_S} & 0^{-+} & G^{\rho\sigma}\tilde G_{\rho\sigma},\, i\bar t \gamma^5 t \\
& \vv{8_A} & 1^{--} & \bar q\gamma^\mu q,\, i\bar t \l[\gamma^\mu, \gamma^\nu\r] t \\
\end{array}$$
\caption{KK gluonia $(\kkg\kkg)$ (UED) and gluinonia $(\go\go)$ (MSSM) and their possible couplings to gluons and quarks. Here $G^{\mu\nu}$ is the gluon field strength and $\tilde G^{\mu\nu} = \frac{1}{2}\epsilon^{\mu\nu\rho\sigma}G_{\rho\sigma}$; $q$ denotes any quark flavor, while $t$ refers to the top quark whose mass we do not neglect. In part of the spin-$1$ cases, the coupling is not to the bound state operator $V_\mu$ but to its derivative $\pd_\mu V_\nu$. }
\label{tab-KKgKKg-gogo-couplings}
\end{table}
In table~\ref{tab-KKgKKg-gogo-couplings} we list the allowed KK gluonia and gluinonia (taking the spin-statistics theorem into account) and the effective vertices through which each of them can couple to gluon or quark bilinears. The possible couplings to a pair of photons (for color-singlets) are like the couplings to a pair of gluons.
To determine the charge conjugation properties of the color-octet bound states, it is useful to know that the action of charge conjugation $C$ on a non-abelian gauge field $V_\mu^a$ is~\cite{Tyutin:1982fx,Smolyakov:1980wq,Burgess-Moore}
\be
T^a V_\mu^a \to -\l(T^a\r)^{\rm T}V_\mu^a
\label{C-nonAbelian}
\ee
where $T^a$ are the generators of gauge transformations. The minus sign here can be described by saying that gluons have $C = -1$ (this is similar to the photon which transforms as $A_\mu \to -A_\mu$, thus having $C = -1$ in a more straightforward sense). Analogous transformation rules apply to KK gluons and gluinos. In the same sense, we find that $C = +1$ for $\vv{8_S}$ KK gluonia and gluinonia and $C = -1$ for $\vv{8_A}$ KK gluonia and gluinonia.\footnote{Our result for the charge conjugation of the $\vv{8_A}$ gluinonium ($(\go\go),\,\vv{8_A},\,J^{PC} = 1^{--}$), or perhaps just the convention for defining the sign of $C$ that would describe (\ref{C-nonAbelian}), differs from that of~\cite{Goldman:1984mj,Keung:1983wz,Kauth:2009ud}. Note that our result (but not theirs) allows the coupling of this gluinonium to massless $q\bar q$ pairs (via the vector current $\bar q\gamma^\mu q$), which must be possible as we know explicitly from the diagrams.}
From table~\ref{tab-KKgKKg-gogo-couplings} we see that the scalar KK gluonia and pseudoscalar gluinonia can be produced only by gluon fusion (but can annihilate also to $t\bar t$). The spin-$1$ KK gluonium cannot be produced at all from gluons or massless quarks while the vector gluinonium can be produced from quarks. The tensor KK gluonium couples to both gluons and quarks. We have checked that the explicit results that we will now present, based on diagrams in figures~\ref{fig-diagrams-KKgKKg} and~\ref{fig-diagrams-gogo}, indeed match these effective vertices. The tensor KK gluonium happens to couple to gluons only through $G^{\rho\mu}G_{\rho\nu}$ but not $G^{\rho\sigma}D_\mu D_\nu G_{\rho\sigma}$.
\begin{figure}[t]
$$
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\startrotation by 0.6 -0.6 about -100 35
\plot -84 30 -84 45 /
\plot -69 30 -69 45 /
\plot -54 30 -54 45 /
\ellipticalarc axes ratio 2:1 150 degrees from -92 31 center at -99 35
\ellipticalarc axes ratio 2:1 280 degrees from -77 31 center at -84 35
\ellipticalarc axes ratio 2:1 280 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\stoprotation
\startrotation by 0.6 0.6 about -100 -35
\plot -84 -30 -84 -45 /
\plot -69 -30 -69 -45 /
\plot -54 -30 -54 -45 /
\ellipticalarc axes ratio 2:1 -150 degrees from -92 -31 center at -99 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -77 -31 center at -84 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\stoprotation
\startrotation by 0.6 0.6 about -30 35
\ellipticalarc axes ratio 2:1 -150 degrees from -38 31 center at -31 35
\ellipticalarc axes ratio 2:1 -280 degrees from -53 31 center at -46 35
\ellipticalarc axes ratio 2:1 -280 degrees from -68 31 center at -61 35
\ellipticalarc axes ratio 2:1 -280 degrees from -83 31 center at -76 35
\stoprotation
\startrotation by 0.6 -0.6 about -30 -35
\ellipticalarc axes ratio 2:1 150 degrees from -38 -31 center at -31 -35
\ellipticalarc axes ratio 2:1 280 degrees from -53 -31 center at -46 -35
\ellipticalarc axes ratio 2:1 280 degrees from -68 -31 center at -61 -35
\ellipticalarc axes ratio 2:1 280 degrees from -83 -31 center at -76 -35
\stoprotation
\put {$\kkg$} at -115 30
\put {$\kkg$} at -115 -30
\put {$g$} at -20 30
\put {$g$} at -20 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -120 0 to 0 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -5 26 10 26 /
\plot -5 11 10 11 /
\plot -5 -4 10 -4 /
\plot -5 -19 10 -19 /
\plot -5 -34 10 -34 /
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -150 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 150 degrees from 62 -31 center at 69 -35
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\kkg$} at -55 15
\put {$\kkg$} at -55 -15
\put {$\kkg$} at -18 0
\put {$g$} at 60 15
\put {$g$} at 60 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -5 26 10 26 /
\plot -5 11 10 11 /
\plot -5 -4 10 -4 /
\plot -5 -19 10 -19 /
\plot -5 -34 10 -34 /
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -150 degrees from 57 25 center at 64 29
\stoprotation
\startrotation by 0.7 -0.75 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -280 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -280 degrees from 77 31 center at 84 35
\ellipticalarc axes ratio 2:1 -150 degrees from 92 31 center at 99 35
\stoprotation
\startrotation by 0.7 0.75 about 0 -30
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 280 degrees from 62 -31 center at 69 -35
\ellipticalarc axes ratio 2:1 280 degrees from 77 -31 center at 84 -35
\ellipticalarc axes ratio 2:1 150 degrees from 92 -31 center at 99 -35
\stoprotation
\put {$\kkg$} at -55 15
\put {$\kkg$} at -55 -15
\put {$\kkg$} at -18 0
\put {$g$} at 75 15
\put {$g$} at 75 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\startrotation by 0.6 -0.6 about -100 35
\plot -84 30 -84 45 /
\plot -69 30 -69 45 /
\plot -54 30 -54 45 /
\ellipticalarc axes ratio 2:1 150 degrees from -92 31 center at -99 35
\ellipticalarc axes ratio 2:1 280 degrees from -77 31 center at -84 35
\ellipticalarc axes ratio 2:1 280 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\stoprotation
\startrotation by 0.6 0.6 about -100 -35
\plot -84 -30 -84 -45 /
\plot -69 -30 -69 -45 /
\plot -54 -30 -54 -45 /
\ellipticalarc axes ratio 2:1 -150 degrees from -92 -31 center at -99 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -77 -31 center at -84 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\stoprotation
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\startrotation by 0.6 0.6 about 37 35
\ellipticalarc axes ratio 2:1 -150 degrees from 29 31 center at 36 35
\ellipticalarc axes ratio 2:1 -280 degrees from 14 31 center at 21 35
\ellipticalarc axes ratio 2:1 -280 degrees from -1 31 center at 6 35
\ellipticalarc axes ratio 2:1 -280 degrees from -16 31 center at -9 35
\stoprotation
\startrotation by 0.6 -0.6 about 37 -35
\ellipticalarc axes ratio 2:1 150 degrees from 29 -31 center at 36 -35
\ellipticalarc axes ratio 2:1 280 degrees from 14 -31 center at 21 -35
\ellipticalarc axes ratio 2:1 280 degrees from -1 -31 center at 6 -35
\ellipticalarc axes ratio 2:1 280 degrees from -16 -31 center at -9 -35
\stoprotation
\put {$\kkg$} at -70 35
\put {$\kkg$} at -70 -30
\put {$g$} at -30 20
\put {$g$} at 5 30
\put {$g$} at 5 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 70 0
\endpicture
$$
$$
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot 0 35 70 35 /
\plot 0 -35 70 -35 /
\barrow from 5 35 to 40 35
\barrow from 70 -35 to 25 -35
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\plot 0 35 0 -35 /
\plot -6 25 6 25 /
\plot -6 10 6 10 /
\plot -6 -10 6 -10 /
\plot -6 -25 6 -25 /
\setsolid
\barrow from 0 -8 to 0 8
\put {$\kkg$} at -55 15
\put {$\kkg$} at -55 -15
\put {$\kkq$} at 15 0
\put {$q$} at 60 20
\put {$\bar q$} at 60 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot 0 35 70 -35 /
\plot 0 -35 30 -5 /
\plot 40 5 70 35 /
\barrow from 70 -35 to 45 -10
\barrow from 50 15 to 60 25
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\plot 0 35 0 -35 /
\plot -6 25 6 25 /
\plot -6 10 6 10 /
\plot -6 -10 6 -10 /
\plot -6 -25 6 -25 /
\barrow from 0 8 to 0 -8
\put {$\kkg$} at -55 15
\put {$\kkg$} at -55 -15
\put {$\kkq$} at -15 0
\put {$q$} at 70 15
\put {$\bar q$} at 70 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\startrotation by 0.6 -0.6 about -100 35
\plot -84 30 -84 45 /
\plot -69 30 -69 45 /
\plot -54 30 -54 45 /
\ellipticalarc axes ratio 2:1 150 degrees from -92 31 center at -99 35
\ellipticalarc axes ratio 2:1 280 degrees from -77 31 center at -84 35
\ellipticalarc axes ratio 2:1 280 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\stoprotation
\startrotation by 0.6 0.6 about -100 -35
\plot -84 -30 -84 -45 /
\plot -69 -30 -69 -45 /
\plot -54 -30 -54 -45 /
\ellipticalarc axes ratio 2:1 -150 degrees from -92 -31 center at -99 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -77 -31 center at -84 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\stoprotation
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\plot 37 37 0 0 37 -37 /
\barrow from 0 0 to 25 25
\barrow from 35 -35 to 15 -15
\put {$\kkg$} at -70 35
\put {$\kkg$} at -70 -30
\put {$g$} at -30 20
\put {$q$} at 10 30
\put {$\bar q$} at 10 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -120 0 to 60 0
\endpicture$$
\caption{Diagrams for production or annihilation of a pair of KK gluons. For bound states, the diagrams with $s$ channel gluon do not happen to contribute.}
\label{fig-diagrams-KKgKKg}
$$\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 35 0 35 0 -35 -70 -35 /
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -150 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 150 degrees from 62 -31 center at 69 -35
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\go$} at -55 15
\put {$\go$} at -55 -15
\put {$\go$} at -18 0
\put {$g$} at 60 15
\put {$g$} at 60 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 35 0 35 0 -35 -70 -35 /
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -150 degrees from 57 25 center at 64 29
\stoprotation
\startrotation by 0.7 -0.75 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -280 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -280 degrees from 77 31 center at 84 35
\ellipticalarc axes ratio 2:1 -150 degrees from 92 31 center at 99 35
\stoprotation
\startrotation by 0.7 0.75 about 0 -30
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 280 degrees from 62 -31 center at 69 -35
\ellipticalarc axes ratio 2:1 280 degrees from 77 -31 center at 84 -35
\ellipticalarc axes ratio 2:1 150 degrees from 92 -31 center at 99 -35
\stoprotation
\put {$\go$} at -55 15
\put {$\go$} at -55 -15
\put {$\go$} at -18 0
\put {$g$} at 75 15
\put {$g$} at 75 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -100 35 -65 0 /
\startrotation by 0.6 -0.6 about -100 35
\ellipticalarc axes ratio 2:1 150 degrees from -92 31 center at -99 35
\ellipticalarc axes ratio 2:1 280 degrees from -77 31 center at -84 35
\ellipticalarc axes ratio 2:1 280 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\stoprotation
\plot -100 -35 -65 0 /
\startrotation by 0.6 0.6 about -100 -35
\ellipticalarc axes ratio 2:1 -150 degrees from -92 -31 center at -99 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -77 -31 center at -84 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\stoprotation
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\startrotation by 0.6 0.6 about 37 35
\ellipticalarc axes ratio 2:1 -150 degrees from 29 31 center at 36 35
\ellipticalarc axes ratio 2:1 -280 degrees from 14 31 center at 21 35
\ellipticalarc axes ratio 2:1 -280 degrees from -1 31 center at 6 35
\ellipticalarc axes ratio 2:1 -280 degrees from -16 31 center at -9 35
\stoprotation
\startrotation by 0.6 -0.6 about 37 -35
\ellipticalarc axes ratio 2:1 150 degrees from 29 -31 center at 36 -35
\ellipticalarc axes ratio 2:1 280 degrees from 14 -31 center at 21 -35
\ellipticalarc axes ratio 2:1 280 degrees from -1 -31 center at 6 -35
\ellipticalarc axes ratio 2:1 280 degrees from -16 -31 center at -9 -35
\stoprotation
\put {$\go$} at -70 35
\put {$\go$} at -70 -30
\put {$g$} at -30 20
\put {$g$} at 5 30
\put {$g$} at 5 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 90 0
\endpicture$$
$$
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 35 70 35 /
\plot -70 -35 70 -35 /
\barrow from 5 35 to 40 35
\barrow from 70 -35 to 25 -35
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\setdashes
\plot 0 35 0 -35 /
\setsolid
\barrow from 0 -8 to 0 8
\put {$\go$} at -55 15
\put {$\go$} at -55 -15
\put {$\sq$} at 15 0
\put {$q$} at 60 20
\put {$\bar q$} at 60 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -60 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 35 0 35 70 -35 /
\plot -70 -35 0 -35 30 -5 /
\plot 40 5 70 35 /
\barrow from 70 -35 to 45 -10
\barrow from 50 15 to 60 25
\ellipticalarc axes ratio 2:1 150 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\ellipticalarc axes ratio 2:1 280 degrees from -32 31 center at -39 35
\ellipticalarc axes ratio 2:1 280 degrees from -17 31 center at -24 35
\ellipticalarc axes ratio 2:1 220 degrees from -0 35 center at -10 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\setdashes
\plot 0 35 0 -35 /
\setsolid
\barrow from 0 8 to 0 -8
\put {$\go$} at -55 15
\put {$\go$} at -55 -15
\put {$\sq$} at -15 0
\put {$q$} at 70 15
\put {$\bar q$} at 70 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -100 35 -65 0 /
\startrotation by 0.6 -0.6 about -100 35
\ellipticalarc axes ratio 2:1 150 degrees from -92 31 center at -99 35
\ellipticalarc axes ratio 2:1 280 degrees from -77 31 center at -84 35
\ellipticalarc axes ratio 2:1 280 degrees from -62 31 center at -69 35
\ellipticalarc axes ratio 2:1 280 degrees from -47 31 center at -54 35
\stoprotation
\plot -100 -35 -65 0 /
\startrotation by 0.6 0.6 about -100 -35
\ellipticalarc axes ratio 2:1 -150 degrees from -92 -31 center at -99 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -77 -31 center at -84 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\stoprotation
\setsolid
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\plot 37 37 0 0 37 -37 /
\barrow from 0 0 to 25 25
\barrow from 35 -35 to 15 -15
\put {$\go$} at -70 35
\put {$\go$} at -70 -30
\put {$g$} at -30 20
\put {$q$} at 10 30
\put {$\bar q$} at 10 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -120 0 to 60 0
\endpicture
$$
\caption{Diagrams for production or annihilation of a pair of gluinos.}
\label{fig-diagrams-gogo}
\end{figure}
\begin{table}[t]
$$\begin{array}{|c|c|c|c|c|}\hline
\,\mbox{process} \,&\, J = 0\,, J_z = 0 \,&\, J = 1\,, |J_z| = 1 \,&\, J = 2\,, |J_z| = 2 \,&\, J = 2\,, |J_z| = 1 \,\\\hline\hline
\,gg\to(\kkg\kkg)_\vv{1} \,&\,729/128\,&\, \,&\,243/8 \,& \\
\,gg\to(\kkg\kkg)_\vv{8_S} \,&\,729/512\,&\, \,&\,243/32\,& \\\hline
\,gg \to(\go\go)_\vv{1} \,&\,243/64 \,&\, \,&\, \,& \,\\
\,gg \to(\go\go)_\vv{8_S} \,&\,243/256\,&\, \,&\, \,& \,\\\hline
\,gg \to(\phi\phi)_\vv{1} \,&\,243/128\,&\, \,&\, \,& \,\\
\,gg \to(\phi\phi)_\vv{8_S} \,&\,243/512\,&\, \,&\, \,& \,\\\hline\hline
\,q\bar q\to(\kkg \kkg)_\vv{1} \,&\, \,&\, \,&\, \,&\, \ds 4\l(\frac{2m_\kkg^2}{m_\kkg^2 + m_\kkq^2}\r)^2 \,\\
\,q\bar q\to(\kkg \kkg)_\vv{8_S}\,&\, \,&\, \,&\, \,&\, \ds\frac{5}{4}\l(\frac{2m_\kkg^2}{m_\kkg^2 + m_\kkq^2}\r)^2 \,\\\hline
\,q\bar q\to(\go\go)_\vv{8_A} \,&\, \,&\,\ds\frac{9}{4}\l(\frac{m_\sq^2 - m_\go^2}{m_\sq^2 + m_\go^2}\r)^2 \,&\, \,&\, \,\\\hline
\end{array}$$
\caption{KK gluonium $(\kkg\kkg)$ (UED), gluinonium $(\go\go)$ (MSSM), and octetonium $(\phi\phi)$ production cross section prefactors $P_{ij}$ of (\ref{parton-cs}). The different columns correspond to different values of the spin $J$ of the bound state and its polarization (the projection $J_z$ on the beam axis).}
\label{tab-KKgluonium}
\end{table}
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.49\textwidth]{cs-jj-M.eps}
\includegraphics[width=0.50\textwidth]{cs-bb-M.eps}\\\vspace{3mm}
\includegraphics[width=0.50\textwidth]{cs-tt-M.eps}
\caption{Dijet (top left), $b\bar b$ (top right) and $t\bar t$ (bottom) signals from KK gluonium annihilation (solid lines) vs. gluinonium annihilation (dashed lines) as a function of the resonance mass ($M = 2m_\kkg$ or $2m_\go$) at the $14$ TeV LHC. The different curves correspond to different ratios of $m_\kkq/m_\kkg$ or $m_\sq/m_\go$ that are indicated next to them. The $b\bar b$ signal of $(\go\go)$ vanishes for $m_\sq = m_\go$, and the $b\bar b$ and $t\bar t$ signals of $(\kkg\kkg)$ vanish for $m_\kkq/m_\kkg \to \infty$.}
\label{fig-cs-M}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.43\textwidth]{dijet-angular.eps}
\includegraphics[width=0.43\textwidth]{bbbar-angular.eps}\\\vspace{3mm}
\includegraphics[width=0.43\textwidth]{ttbar-angular.eps}
\caption{Angular distributions in annihilation of KK gluonia and gluinonia into dijets (top left), $b\bar b$ (top right) and $t\bar t$ (bottom) at the LHC for $M = 800$~GeV (i.e., $m_\kkg = m_\go = 400$~GeV). Here $\theta$ is the angle between the beam axis and the direction of motion of the annihilation products in the center-of-mass frame. The different curves correspond to different ratios of $m_\kkq/m_\kkg$ or $m_\sq/m_\go$ that are indicated next to them (except for the $b\bar b$ signal of $(\go\go)$ where there is no such dependence).}
\label{fig-angular}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.43\textwidth]{bbbar-angular-pythia.eps}
\caption{Simulated angular distributions in annihilation of KK gluonia and gluinonia into $b\bar b$ at the LHC.}
\label{fig-angular-Pythia}
\end{center}
\end{figure}
The parton-level bound state production cross sections can be written as
\be
\hat\sigma_{{\rm bound},\,ij}(\hat s) =
P_{ij}\, \zeta(3)\,\pi^2\bar\alpha_s^3\alpha_s^2\,\delta(\hat s-M^2)
\label{parton-cs}
\ee
where $\zeta(3) = \sum_{n=1}^\infty 1/n^3 \simeq 1.2$ takes into account the contributions of the radial excitations and the prefactors $P_{ij}$ for producing the various bound states from partons $i,j$ are presented in table~\ref{tab-KKgluonium} for KK gluonium and gluinonium. These results are based on the diagrams in figures~\ref{fig-diagrams-KKgKKg} and~\ref{fig-diagrams-gogo} and factors such as (\ref{wavefunction}). More details are given in appendix~\ref{app-rates}. After convoluting (\ref{parton-cs}) with the parton distribution functions as\footnote{We are using NLO MSTW 2008 PDFs~\cite{Martin:2009iq} evaluated at the scale $M/2$, and the center-of-mass energy $\sqrt s = 14$ TeV for the LHC.}
\be
\sigma_{\rm bound} = \frac{\zeta(3)\,\pi^2\bar\alpha_s^3\alpha_s^2}{s} \sum_{i,j}P_{ij}\int_{M^2/s}^1 \frac{dx}{x}\, f_{i/p}(x)\, f_{j/p}\l(\frac{M^2}{xs}\r)
\ee
and multiplying by the appropriate branching ratios, we obtain the cross sections for dijet, $b\bar b$ and $t\bar t$ final states as shown in figure~\ref{fig-cs-M}. We also show to what extent the results depend on the mass ratios $m_\kkq/m_\kkg$ or $m_\sq/m_\go$ by varying them between $1$ and $\infty$.\footnote{In practice, the ratio must be somewhat larger than $1$ in order for the annihilation rates to dominate, and the extreme limit of $m_\kkq/m_\kkg \to \infty$ is unphysical for UED (although it may be relevant for some other theory that does not contain KK quarks).} We see that the signals will typically be an order of magnitude larger for KK gluonia than for gluinonia of the same mass (except if the KK quarks are very heavy, in which case the $b\bar b$ and $t\bar t$ signals of KK gluonia disappear).
It is interesting to ask what creates this large difference in the cross sections. Unless the KK gluons (or gluinos) are heavier than about a TeV, production via the $gg$ channel dominates due to its higher luminosity, so let us discuss it first and understand the order-of-magnitude difference seen already in the parton level expressions of table~\ref{tab-KKgluonium}. The differences should be attributed to a large extent to the more numerous spin possibilities for the KK gluonia. According to table~\ref{tab-KKgluonium}, in the $gg$ channel (in both $\vv{1}$ and $\vv{8_S}$ representations), the parton-level production cross section for $J=0$ KK gluonium is only a factor of $3/2$ larger than that for $J=0$ gluinonium. However, the KK gluonium can also be produced in $J=2$ state, whose cross section (summed over the spin projections) is larger than that of the ($J=0$) gluinonium by a factor of $8$ (or larger than that of $J=0$ KK gluonium by a factor of $16/3 \sim 5$). Overall, the $gg$ channel produces $9.5$ times more KK gluonium than gluinonium.\footnote{In this case, this happens to be exactly the ratio of the near-threshold pair production cross sections (despite the fact that for the purposes of bound states we multiply this quantity by a different $\l|\psi(\vv{0})\r|^2$ for each of the attractive color representations and exclude the repulsive ones). In general, this does not need to be the case. For example, in the $q\bar q$ channel the ratio between bound state and near-threshold pair production cross section is $3$ times bigger for the KK gluonia than for the gluinonia because the color representations in which KK gluon and gluino pairs are produced at threshold are different.} Note that these results are determined by gauge interactions alone, and therefore apply much more generally than to just UED and MSSM. In particular, the production diagrams do not involve any other new particles besides the KK gluons or gluinos.
In the $q\bar q$ channel, the comparison between KK gluonia and gluinonia is sensitive to the masses of the KK quarks (relative to the KK gluon) or squarks (relative to the gluino). If the KK quarks are very heavy the KK gluonium cross section goes to zero, while there is no such effect for the gluinonium. This is because the gluinonium can be produced via a diagram with an $s$ channel gluon that does not involve squarks, while a similar diagram for KK gluonium vanishes at threshold. On the other hand, the same $s$ channel diagram of gluinonium interferes destructively with diagrams with $t$ and $u$ channel squarks, and this makes the gluinonium cross section vanish if the squarks are degenerate with the gluino, while there is no such effect for the KK gluonium. Note however that the presence of this effect for the gluinonium depends on the existence of squarks in MSSM, and thus will not be a general feature of new physics models with color-octet spin-$1/2$ particles. Similarly, the production of KK gluonia through the $q\bar q$ channel depends on the existence of KK quarks, and thus may not be present in some other model with color-octet spin-$1$ particles. Another difference between gluinonia and KK gluonia is that even though in both cases the bound states are produced with spin component $\pm 1$ along the beam axis, it is an $\vv{8_A}$ ($J=1$) state for gluinonium and $\vv{1}$ or $\vv{8_S}$ ($J=2$) state for KK gluonium. The production of the $\vv{1}$ state (with its large color factor $(C_\vv{1}/C_\vv{8})^3 = 8$) in UED makes the cross section larger.
We also see from table~\ref{tab-KKgluonium} that KK gluonia will be produced predominantly in spin-$2$ states. If the angular distributions of the annihilation products can be measured, they can be used to distinguish the KK gluonium from the gluinonium (which is produced predominantly in spin-$0$ states, and also in the spin-$1$ state which is important for some of the annihilation channels). The angular distributions in the dijet, $b\bar b$ and $t\bar t$ channels are plotted in figure~\ref{fig-angular}. The curves measured in experiment will be further affected by detector acceptances, effects of QCD radiation, mistakes in reconstruction, cuts, and uncertainties in modeling the QCD background which needs to be subtracted from the data. However, the remarkable differences between KK gluonium and gluinonium that we see in figure~\ref{fig-angular} will hopefully remain.
To simulate the effects of some of the experimental factors on the angular distributions, we generated events with \textsc{Pythia}, and the result (for example, for the $b\bar b$ channel) is shown in figure~\ref{fig-angular-Pythia}. The details of the simulation and the imposed cuts are the same as will be described in section~\ref{sec-simulation} (except that the cut on $\cos\theta$ is not included and the $p_T$ cut is relaxed to $p_T > 0.3 M$).
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.55\textwidth]{cs-gammagamma-M.eps}
\caption{Diphoton annihilation signal of gluinonium (thick dashed line), KK gluonium (estimate; thick solid line), squarkonium (thin dashed line) and KK quarkonium (thin solid line) at the $14$ TeV LHC as a function of the resonance mass. For gluinonium we assumed $m_\sq = m_\go$ (and similarly for KK gluonium). For squarkonium and KK quarkonium we present the contribution of a single flavor and chirality, and we assumed the charges of the constituent particles to be $|Q| = 2/3$ (for $|Q| = 1/3$, the cross section is $16$ times smaller).}
\label{fig-cs-diphoton-M}
\end{center}
\end{figure}
Another channel we can look at is the annihilation of color-singlet KK gluonia or gluinonia into a pair of photons. Even though these processes can only proceed through loop diagrams because the KK gluons and gluinos are neutral, it is still interesting to consider this signal because the background in the diphoton channel is much more favorable than in the dijet, $b\bar b$ or $t\bar t$ channels. As long as the squarks are not much heavier than the gluino, the annihilation rate of the color-singlet gluinonium into $\gamma\gamma$ is $\sim 10^{-5}$ of its annihilation rate into $gg$ (the exact expression from~\cite{Kauth:2009ud} is given in~(\ref{gogo-diphoton})). The relevant loop diagrams for KK gluonium have not been computed, but for our estimates we will assume that the diphoton branching ratio for spin-$0$ and spin-$2$ color-singlet KK gluonia is the same as for gluinonium. We expect this to be correct up to an ${\cal O}\,(1)$ factor since the relevant coupling constants and the gauge quantum numbers of the relevant particles are the same in both cases. The signal cross sections are shown in figure~\ref{fig-cs-diphoton-M}. The angular distributions in this channel may also be useful for discrimination since for gluinonia only the $J=0$ state contributes, while for KK gluonia the contribution may be coming from both $J=0$ and $J=2$.
We can also compare the KK gluonia and gluinonia (which are bound states of spin-$1$ and spin-$1/2$ color octets) with bound states of scalar color-octets. Pairs of scalar color-octets $\phi$ will be produced by gluon fusion via the same diagrams as the KK gluons in figure~\ref{fig-diagrams-KKgKKg}, although in practice only the quartic vertex is relevant to bound states -- ``octetonia'' $(\phi\phi)$ in $\vv{1}$ and $\vv{8_S}$ color representations. Their production cross sections are included in table~\ref{tab-KKgluonium}.\footnote{The color-singlet result can be found in~\cite{Kim:2008bx}, and it agrees with ours.} They are twice as small as the gluinonium cross sections. The octetonia decay rates are given in appendix~\ref{app-rates-octetonia}. Since the octetonia are scalars they will annihilate isotropically, and unless $\phi$ is charged under the electroweak group or interacts with some other particles, the only significant annihilation signal of octetonia will be $gg$ dijets. It will be easy to distinguish them from bound states of spin-$1$ color-octets (like KK gluonia) because of the factor of $\sim 20$ difference in the cross section and the very different angular distributions. It will be more difficult to distinguish them from bound states of spin-$1/2$ color-octets (like gluinonia) because the angular distributions are also almost isotropic in the spin-$1/2$ case and the difference in the cross section is only a factor of $2$. If the difference in the cross section is to be used, potential multiplicity of degenerate color-octets and higher order QCD corrections will be important. Part of the QCD corrections for gluinonium have been computed in~\cite{Hagiwara:2009hq,Kauth:2009ud}.
\section{KK quarkonia vs. squarkonia\label{sec-KKquarkonium}}
\setcounter{equation}{0}
\begin{table}[t]
$$\begin{array}{c|c|c|c}
\mbox{} & \,\mbox{color} \,&\, J^{PC} \,&\, \mbox{can couple to} \,\\\hline\hline
(\kkqR\bar \kkqR) + (\kkqL\bar \kkqL) & \vv{1} & 0^{-+} & G^{\rho\sigma}\tilde G_{\rho\sigma},\, i\bar t \gamma^5 t \\\hline
(\kkqR\bar \kkqR) - (\kkqL\bar \kkqL) & \vv{1} & 0^{+-} & - \\\hline
(\kkqL\bar \kkqR) & \vv{1} & 0^{-+} & i\bar q P_R q,\, i\bar t\gamma^5 t \\\hline
(\kkqR\bar \kkqR) + (\kkqL\bar \kkqL) & \vv{1} & 1^{--} &\; \bar q \gamma^\mu q,\, i\bar t\l[\gamma^\mu,\gamma^\nu\r]t,\, \bar\ell \gamma^\mu \ell,\, \bar\ell\gamma^\mu\gamma^5\ell \\\hline
(\kkqR\bar \kkqR) - (\kkqL\bar \kkqL) & \vv{1} & 1^{++} & \bar q\gamma^\mu\gamma^5q,\, \bar\ell \gamma^\mu \ell,\, \bar\ell\gamma^\mu\gamma^5\ell \\\hline
(\kkqL\bar \kkqR) & \vv{1} & 1^{--} & i\bar q\l[\gamma^\mu,\gamma^\nu\r]P_R q,\, \bar t \gamma^\mu t \\\hline\hline
(\sq_R\sq_R^\ast) + (\sq_L\sq_L^\ast) & \vv{1} & 0^{++} & G^{\rho\sigma}G_{\rho\sigma},\, \bar t t \\\hline
(\sq_R\sq_R^\ast) - (\sq_L\sq_L^\ast) & \vv{1} & 0^{--} & - \\\hline
(\sq_L\sq_R^\ast) & \vv{1} & 0^{++} & \bar q P_R q,\, \bar tt\\
\end{array}$$
\caption{KK quarkonia $(\kkq\bar\kkq)$ (UED) and squarkonia $(\sq\sq^\ast)$ (MSSM) and their possible couplings to gluons, quarks and leptons. $P_{R,L} = (1\pm\gamma^5)/2$. Antiparticle bound states, where relevant, are also present even if not listed in the table. In part of the spin-$1$ cases, the coupling is not to the bound state operator $V_\mu$ but to its derivative $\pd_\mu V_\nu$.}
\label{tab-KKqKKaq-sqasq-couplings}
\end{table}
In this section we study bound states of fundamental-antifundamental pairs of KK quarks (spin $1/2$) in UED and squarks (spin $0$) in MSSM. Later in this section we will also look at bound states of spin $1$ particles in the fundamental representation.
Table~\ref{tab-KKqKKaq-sqasq-couplings} classifies the KK quarkonia and squarkonia and their possible couplings to gluons and quarks (via the strong interactions) and leptons (via the electroweak interactions). The possible couplings to a pair of photons are like to a pair of gluons. The diagrams through which the various processes can be realized are shown in figures~\ref{fig-diagrams-KKqKKqbar} and~\ref{fig-diagrams-sqasq}.
For both KK quarkonia and squarkonia, the dominant products of the gluon fusion channel are spin-$0$ bound states in which the KK quarks or squarks have same flavors and chiralities. These are also the only bound states that can annihilate into a pair of photons, which is the most promising detection channel as we discuss in the following. The angular distributions in the $\gamma\gamma$ channel will be isotropic in both UED and MSSM but the size of the cross section can still be used for discrimination.
\begin{figure}[t]
$$
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 0 37 0 -37 -70 -37 /
\barrow from -37 37 to -25 37
\barrow from -25 -37 to -37 -37
\barrow from 0 5 to 0 -5
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -6 25 6 25 /
\plot -6 10 6 10 /
\plot -6 -10 6 -10 /
\plot -6 -25 6 -25 /
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -150 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 150 degrees from 62 -31 center at 69 -35
\put {$\kkq$} at -55 20
\put {$\bar\kkq$} at -55 -20
\put {$\kkq$} at 15 0
\put {$g$} at 50 20
\put {$g$} at 50 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 3 37 3 -37 -70 -37 /
\barrow from -37 37 to -25 37
\barrow from -25 -37 to -37 -37
\barrow from 3 5 to 3 -5
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -3 25 9 25 /
\plot -3 10 9 10 /
\plot -3 -10 9 -10 /
\plot -3 -25 9 -25 /
\startrotation by 0.7 -0.75 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -280 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -280 degrees from 77 31 center at 84 35
\ellipticalarc axes ratio 2:1 -150 degrees from 92 31 center at 99 35
\stoprotation
\startrotation by 0.7 0.75 about 0 -30
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 280 degrees from 62 -31 center at 69 -35
\ellipticalarc axes ratio 2:1 280 degrees from 77 -31 center at 84 -35
\ellipticalarc axes ratio 2:1 150 degrees from 92 -31 center at 99 -35
\stoprotation
\put {$\kkq$} at -55 20
\put {$\bar\kkq$} at -55 -20
\put {$\kkq$} at -15 0
\put {$g$} at 75 15
\put {$g$} at 75 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -100 35 -65 0 -100 -35 /
\barrow from -90 25 to -80 15
\barrow from -80 -15 to -90 -25
\startrotation by 0.6 -0.6 about -100 35
\plot -84 30 -84 45 /
\plot -69 30 -69 45 /
\plot -54 30 -54 45 /
\stoprotation
\startrotation by 0.6 0.6 about -100 -35
\plot -84 -30 -84 -45 /
\plot -69 -30 -69 -45 /
\plot -54 -30 -54 -45 /
\stoprotation
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\startrotation by 0.6 0.6 about 37 35
\ellipticalarc axes ratio 2:1 -150 degrees from 29 31 center at 36 35
\ellipticalarc axes ratio 2:1 -280 degrees from 14 31 center at 21 35
\ellipticalarc axes ratio 2:1 -280 degrees from -1 31 center at 6 35
\ellipticalarc axes ratio 2:1 -280 degrees from -16 31 center at -9 35
\stoprotation
\startrotation by 0.6 -0.6 about 37 -35
\ellipticalarc axes ratio 2:1 150 degrees from 29 -31 center at 36 -35
\ellipticalarc axes ratio 2:1 280 degrees from 14 -31 center at 21 -35
\ellipticalarc axes ratio 2:1 280 degrees from -1 -31 center at 6 -35
\ellipticalarc axes ratio 2:1 280 degrees from -16 -31 center at -9 -35
\stoprotation
\put {$\kkq$} at -70 35
\put {$\bar\kkq$} at -65 -30
\put {$g$} at -30 20
\put {$g$} at 5 30
\put {$g$} at 5 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 60 0
\endpicture
$$
$$\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 70 37 /
\plot -70 -39 70 -39 /
\barrow from -37 37 to -25 37
\barrow from -25 -39 to -37 -39
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -32 -54 -47 /
\plot -39 -32 -39 -47 /
\plot -24 -32 -24 -47 /
\plot -9 -32 -9 -47 /
\plot -5 26 10 26 /
\plot -5 11 10 11 /
\plot -5 -4 10 -4 /
\plot -5 -19 10 -19 /
\plot -5 -34 10 -34 /
\barrow from 35 37 to 45 37
\barrow from 45 -39 to 35 -39
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -160 degrees from 57 25 center at 64 29
\stoprotation
\put {$\kkq$} at -50 20
\put {$\bar\kkq$} at -50 -20
\put {$\kkg$} at -18 0
\put {$q$} at 60 20
\put {$\bar q$} at 60 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 100 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -100 35 -65 0 -100 -35 /
\barrow from -90 25 to -80 15
\barrow from -80 -15 to -90 -25
\startrotation by 0.6 -0.6 about -100 35
\plot -84 30 -84 45 /
\plot -69 30 -69 45 /
\plot -54 30 -54 45 /
\stoprotation
\startrotation by 0.6 0.6 about -100 -35
\plot -84 -30 -84 -45 /
\plot -69 -30 -69 -45 /
\plot -54 -30 -54 -45 /
\stoprotation
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\plot 37 37 0 0 37 -37 /
\barrow from 0 0 to 25 25
\barrow from 35 -35 to 15 -15
\put {$\kkq$} at -70 35
\put {$\bar\kkq$} at -65 -30
\put {$g$} at -30 20
\put {$q$} at 10 30
\put {$\bar q$} at 10 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 60 0
\endpicture
$$
$$
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 0 37 0 -37 -70 -37 /
\barrow from -37 37 to -25 37
\barrow from -25 -37 to -37 -37
\barrow from 0 5 to 0 -5
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -6 25 6 25 /
\plot -6 10 6 10 /
\plot -6 -10 6 -10 /
\plot -6 -25 6 -25 /
\ellipticalarc axes ratio 2:1 180 degrees from 12 35 center at 6 35
\ellipticalarc axes ratio 2:1 180 degrees from 12 35 center at 18 35
\ellipticalarc axes ratio 2:1 180 degrees from 36 35 center at 30 35
\ellipticalarc axes ratio 2:1 180 degrees from 36 35 center at 42 35
\ellipticalarc axes ratio 2:1 180 degrees from 60 35 center at 54 35
\ellipticalarc axes ratio 2:1 180 degrees from 12 -35 center at 6 -35
\ellipticalarc axes ratio 2:1 180 degrees from 12 -35 center at 18 -35
\ellipticalarc axes ratio 2:1 180 degrees from 36 -35 center at 30 -35
\ellipticalarc axes ratio 2:1 180 degrees from 36 -35 center at 42 -35
\ellipticalarc axes ratio 2:1 180 degrees from 60 -35 center at 54 -35
\put {$\kkq$} at -55 20
\put {$\bar\kkq$} at -55 -20
\put {$\kkq$} at 15 0
\put {$\gamma$} at 50 18
\put {$\gamma$} at 50 -18
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 0 37 0 -37 -70 -37 /
\barrow from -37 37 to -25 37
\barrow from -25 -37 to -37 -37
\barrow from 0 5 to 0 -5
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -6 25 6 25 /
\plot -6 10 6 10 /
\plot -6 -10 6 -10 /
\plot -6 -25 6 -25 /
\startrotation by 0.8 -0.85 about 0 37
\ellipticalarc axes ratio 2:1 180 degrees from 12 37 center at 6 37
\ellipticalarc axes ratio 2:1 180 degrees from 12 37 center at 18 37
\ellipticalarc axes ratio 2:1 180 degrees from 36 37 center at 30 37
\ellipticalarc axes ratio 2:1 180 degrees from 36 37 center at 42 37
\ellipticalarc axes ratio 2:1 180 degrees from 60 37 center at 54 37
\ellipticalarc axes ratio 2:1 180 degrees from 60 37 center at 66 37
\ellipticalarc axes ratio 2:1 180 degrees from 84 37 center at 78 37
\stoprotation
\startrotation by 0.8 0.9 about 0 -37
\ellipticalarc axes ratio 2:1 180 degrees from 12 -37 center at 6 -37
\ellipticalarc axes ratio 2:1 180 degrees from 12 -37 center at 18 -37
\ellipticalarc axes ratio 2:1 180 degrees from 36 -37 center at 30 -37
\ellipticalarc axes ratio 2:1 180 degrees from 60 -37 center at 54 -37
\ellipticalarc axes ratio 2:1 180 degrees from 60 -37 center at 66 -37
\ellipticalarc axes ratio 2:1 180 degrees from 84 -37 center at 78 -37
\stoprotation
\put {$\kkq$} at -55 20
\put {$\bar\kkq$} at -55 -20
\put {$\kkq$} at -15 0
\put {$\gamma$} at 75 20
\put {$\gamma$} at 75 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\qqq
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -100 35 -65 0 -100 -35 /
\barrow from -90 25 to -80 15
\barrow from -80 -15 to -90 -25
\startrotation by 0.6 -0.6 about -100 35
\plot -84 30 -84 45 /
\plot -69 30 -69 45 /
\plot -54 30 -54 45 /
\stoprotation
\startrotation by 0.6 0.6 about -100 -35
\plot -84 -30 -84 -45 /
\plot -69 -30 -69 -45 /
\plot -54 -30 -54 -45 /
\stoprotation
\ellipticalarc axes ratio 2:1 180 degrees from -48 0 center at -56 0
\ellipticalarc axes ratio 2:1 180 degrees from -48 0 center at -42 0
\ellipticalarc axes ratio 2:1 180 degrees from -24 0 center at -30 0
\ellipticalarc axes ratio 2:1 180 degrees from -24 0 center at -18 0
\ellipticalarc axes ratio 2:1 180 degrees from 0 0 center at -6 0
\plot 37 37 0 0 37 -37 /
\barrow from 0 0 to 25 25
\barrow from 35 -35 to 15 -15
\put {$\kkq$} at -70 35
\put {$\bar\kkq$} at -65 -30
\put {$\gamma,Z$} at -30 20
\put {$\ell^-$} at 55 33
\put {$\ell^+$} at 55 -33
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 60 0
\endpicture
$$
\caption{Diagrams for production or annihilation of a KK quark-KK antiquark pair. For bound states, the diagrams with $s$-channel gluon do not actually contribute.}
\label{fig-diagrams-KKqKKqbar}
$$
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -37 37 0 0 -37 -37 /
\setsolid
\barrow from -25 25 to -15 15
\barrow from -15 -15 to -25 -25
\startrotation by 0.8 0.5 about 0 0
\ellipticalarc axes ratio 2:1 -220 degrees from 0 0 center at 10 0
\ellipticalarc axes ratio 2:1 -280 degrees from 17 -4 center at 24 0
\ellipticalarc axes ratio 2:1 -280 degrees from 32 -4 center at 39 0
\ellipticalarc axes ratio 2:1 -150 degrees from 47 -4 center at 54 0
\stoprotation
\startrotation by 0.8 -0.5 about 0 0
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at 10 0
\ellipticalarc axes ratio 2:1 280 degrees from 17 4 center at 24 0
\ellipticalarc axes ratio 2:1 280 degrees from 32 4 center at 39 0
\ellipticalarc axes ratio 2:1 150 degrees from 47 4 center at 54 0
\stoprotation
\put {$\sq$} at -10 32
\put {$\sq^\ast$} at -10 -32
\put {$g$} at 20 31
\put {$g$} at 20 -35
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -50 0 to 60 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -70 37 0 37 0 -37 -70 -37 /
\setsolid
\barrow from -45 37 to -35 37
\barrow from -35 -37 to -45 -37
\barrow from 0 5 to 0 -5
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -150 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 150 degrees from 62 -31 center at 69 -35
\put {$\sq$} at -45 20
\put {$\sq^\ast$} at -45 -20
\put {$\sq$} at 15 0
\put {$g$} at 50 20
\put {$g$} at 50 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -70 37 0 37 0 -37 -70 -37 /
\setsolid
\barrow from -45 37 to -35 37
\barrow from -35 -37 to -45 -37
\barrow from 0 5 to 0 -5
\setsolid
\startrotation by 0.7 -0.75 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -280 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -280 degrees from 77 31 center at 84 35
\ellipticalarc axes ratio 2:1 -150 degrees from 92 31 center at 99 35
\stoprotation
\startrotation by 0.7 0.75 about 0 -30
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 280 degrees from 62 -31 center at 69 -35
\ellipticalarc axes ratio 2:1 280 degrees from 77 -31 center at 84 -35
\ellipticalarc axes ratio 2:1 150 degrees from 92 -31 center at 99 -35
\stoprotation
\put {$\sq$} at -45 20
\put {$\sq^\ast$} at -45 -20
\put {$\sq$} at -15 0
\put {$g$} at 75 15
\put {$g$} at 75 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -100 35 -65 0 -100 -35 /
\setsolid
\barrow from -90 25 to -80 15
\barrow from -80 -15 to -90 -25
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\startrotation by 0.6 0.6 about 37 35
\ellipticalarc axes ratio 2:1 -150 degrees from 29 31 center at 36 35
\ellipticalarc axes ratio 2:1 -280 degrees from 14 31 center at 21 35
\ellipticalarc axes ratio 2:1 -280 degrees from -1 31 center at 6 35
\ellipticalarc axes ratio 2:1 -280 degrees from -16 31 center at -9 35
\stoprotation
\startrotation by 0.6 -0.6 about 37 -35
\ellipticalarc axes ratio 2:1 150 degrees from 29 -31 center at 36 -35
\ellipticalarc axes ratio 2:1 280 degrees from 14 -31 center at 21 -35
\ellipticalarc axes ratio 2:1 280 degrees from -1 -31 center at 6 -35
\ellipticalarc axes ratio 2:1 280 degrees from -16 -31 center at -9 -35
\stoprotation
\put {$\sq$} at -75 35
\put {$\sq^\ast$} at -70 -30
\put {$g$} at -30 20
\put {$g$} at 5 30
\put {$g$} at 5 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 60 0
\endpicture
$$
$$\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot 70 37 0 37 0 -39 70 -39 /
\setdashes
\plot -70 37 0 37 /
\plot -70 -39 0 -39 /
\setsolid
\barrow from -45 37 to -35 37
\barrow from 35 37 to 45 37
\barrow from -35 -39 to -45 -39
\barrow from 45 -39 to 35 -39
\setsolid
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\sq$} at -50 20
\put {$\sq^\ast$} at -50 -20
\put {$\go$} at -18 0
\put {$q$} at 60 20
\put {$\bar q$} at 60 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 100 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -100 35 -65 0 -100 -35 /
\setsolid
\barrow from -90 25 to -80 15
\barrow from -80 -15 to -90 -25
\setsolid
\ellipticalarc axes ratio 2:1 220 degrees from 0 0 center at -10 0
\ellipticalarc axes ratio 2:1 280 degrees from -17 -4 center at -24 0
\ellipticalarc axes ratio 2:1 280 degrees from -32 -4 center at -39 0
\ellipticalarc axes ratio 2:1 260 degrees from -47 -4 center at -54 0
\plot 37 37 0 0 37 -37 /
\barrow from 0 0 to 25 25
\barrow from 35 -35 to 15 -15
\put {$\sq$} at -70 35
\put {$\sq^\ast$} at -70 -30
\put {$g$} at -30 20
\put {$q$} at 10 30
\put {$\bar q$} at 10 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 60 0
\endpicture
$$
$$
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -37 37 0 0 -37 -37 /
\setsolid
\barrow from -25 25 to -15 15
\barrow from -15 -15 to -25 -25
\startrotation by 0.8 0.6 about 0 0
\ellipticalarc axes ratio 2:1 180 degrees from 12 0 center at 6 0
\ellipticalarc axes ratio 2:1 180 degrees from 12 0 center at 18 0
\ellipticalarc axes ratio 2:1 180 degrees from 36 0 center at 30 0
\ellipticalarc axes ratio 2:1 180 degrees from 36 0 center at 42 0
\ellipticalarc axes ratio 2:1 180 degrees from 60 0 center at 54 0
\stoprotation
\startrotation by 0.8 -0.6 about 0 0
\ellipticalarc axes ratio 2:1 180 degrees from 12 0 center at 6 0
\ellipticalarc axes ratio 2:1 180 degrees from 12 0 center at 18 0
\ellipticalarc axes ratio 2:1 180 degrees from 36 0 center at 30 0
\ellipticalarc axes ratio 2:1 180 degrees from 36 0 center at 42 0
\ellipticalarc axes ratio 2:1 180 degrees from 60 0 center at 54 0
\stoprotation
\put {$\sq$} at -10 32
\put {$\sq^\ast$} at -10 -32
\put {$\gamma$} at 20 31
\put {$\gamma$} at 20 -35
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -50 0 to 70 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -70 37 0 37 0 -37 -70 -37 /
\setsolid
\barrow from -45 37 to -35 37
\barrow from -35 -37 to -45 -37
\barrow from 0 5 to 0 -5
\ellipticalarc axes ratio 2:1 180 degrees from 12 35 center at 6 35
\ellipticalarc axes ratio 2:1 180 degrees from 12 35 center at 18 35
\ellipticalarc axes ratio 2:1 180 degrees from 36 35 center at 30 35
\ellipticalarc axes ratio 2:1 180 degrees from 36 35 center at 42 35
\ellipticalarc axes ratio 2:1 180 degrees from 60 35 center at 54 35
\ellipticalarc axes ratio 2:1 180 degrees from 12 -35 center at 6 -35
\ellipticalarc axes ratio 2:1 180 degrees from 12 -35 center at 18 -35
\ellipticalarc axes ratio 2:1 180 degrees from 36 -35 center at 30 -35
\ellipticalarc axes ratio 2:1 180 degrees from 36 -35 center at 42 -35
\ellipticalarc axes ratio 2:1 180 degrees from 60 -35 center at 54 -35
\put {$\sq$} at -45 20
\put {$\sq^\ast$} at -45 -20
\put {$\sq$} at 15 0
\put {$\gamma$} at 50 18
\put {$\gamma$} at 50 -18
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -70 37 0 37 0 -37 -70 -37 /
\setsolid
\barrow from -45 37 to -35 37
\barrow from -35 -37 to -45 -37
\barrow from 0 5 to 0 -5
\setsolid
\startrotation by 0.8 -0.85 about 0 37
\ellipticalarc axes ratio 2:1 180 degrees from 12 37 center at 6 37
\ellipticalarc axes ratio 2:1 180 degrees from 12 37 center at 18 37
\ellipticalarc axes ratio 2:1 180 degrees from 36 37 center at 30 37
\ellipticalarc axes ratio 2:1 180 degrees from 36 37 center at 42 37
\ellipticalarc axes ratio 2:1 180 degrees from 60 37 center at 54 37
\ellipticalarc axes ratio 2:1 180 degrees from 60 37 center at 66 37
\ellipticalarc axes ratio 2:1 180 degrees from 84 37 center at 78 37
\stoprotation
\startrotation by 0.8 0.9 about 0 -37
\ellipticalarc axes ratio 2:1 180 degrees from 12 -37 center at 6 -37
\ellipticalarc axes ratio 2:1 180 degrees from 12 -37 center at 18 -37
\ellipticalarc axes ratio 2:1 180 degrees from 36 -37 center at 30 -37
\ellipticalarc axes ratio 2:1 180 degrees from 60 -37 center at 54 -37
\ellipticalarc axes ratio 2:1 180 degrees from 60 -37 center at 66 -37
\ellipticalarc axes ratio 2:1 180 degrees from 84 -37 center at 78 -37
\stoprotation
\put {$\sq$} at -45 20
\put {$\sq^\ast$} at -45 -20
\put {$\sq$} at -15 0
\put {$\gamma$} at 75 20
\put {$\gamma$} at 75 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
$$
\caption{Diagrams for production or annihilation of a squark-antisquark pair. For bound states, only the first diagram in each row actually contributes.}
\label{fig-diagrams-sqasq}
\end{figure}
\begin{table}[t]
$$\begin{array}{|c|c|c|c|}\hline
\,\mbox{process} &\,J = 0\,, J_z = 0 \,&\, J = 1\,,|J_z| = 1 \,&\, J = 1\,,J_z = 0 \,\\\hline\hline
\begin{array}{c}gg\to(\kkqL\bar \kkqL)\\\q\mbox{ or }(\kkqR\bar \kkqR)\end{array} &\,\ds\frac{4}{81}\,&\, \,&\, \\ \hline
\begin{array}{c}gg\to(\sq_L\sq_L^\ast)\\\q\mbox{ or }(\sq_R\sq_R^\ast)\end{array} &\,\ds\frac{2}{81}\,&\, \,&\, \,\\ \hline\hline
\begin{array}{c}q\bar q^{(\prime)}\to(\kkqL \bar \kkqL^{(\prime)})\\\qq\mbox{ or }(\kkqR \bar \kkqR^{(\prime)})\end{array} &\, \,&\frac{128}{2187}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2\l(2 + \frac{m_\kkq^2}{m_\kkg^2}\r)^2 &\,\\
\begin{array}{c}q\bar q^{(\prime)}\to(\kkqL\bar \kkqR^{(\prime)})\\\qq\mbox{ or }(\kkqR\bar \kkqL^{(\prime)})\end{array} &\frac{64}{2187}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2 \l(4 + \frac{m_\kkq^2}{m_\kkg^2}\r)^2 &\, \,&\frac{64}{2187}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2 \l(\frac{m_\kkq^2}{m_\kkg^2}\r)^2 \\\hline
\begin{array}{c}q\bar q^{(\prime)}\to(\sq_L\sq^{(\prime)\ast}_R)\\\qq\mbox{ or }(\sq_R\sq^{(\prime)\ast}_L)\end{array} &\ds\frac{512}{2187}\l(\frac{2m_\go m_\sq}{m_\sq^2+m_\go^2}\r)^2 &\, \,&\, \\\hline
\end{array}$$
\caption{KK quarkonium $(\kkq\bar\kkq)$ (UED) and squarkonium $(\sq\sq^\ast)$ (MSSM) production cross section prefactors $P_{ij}$ of (\ref{parton-cs}). All the bound states are color-singlets. We use $q$ and $q'$ to refer to two different flavors of quarks (and similarly for KK quarks and squarks), while the notation $q^{(\prime)}$ means that the flavor can be either the same or different from that of $q$. All the numbers refer to a single choice of flavors and chiralities. For simplicity, we assumed the masses of the different flavors and chiralities of the KK quarks or squarks to be the same.}
\label{tab-KKquarkonium}
\end{table}
In the $q\bar q$ fusion channel, the most interesting products are spin-$1$ bound states of KK quarks with same flavors and chiralities since they can annihilate into a pair of leptons, while there is no such process for squarkonia. A disadvantage of this production mechanism (which goes through the diagram with a $t$ channel KK gluon from figure~\ref{fig-diagrams-KKqKKqbar}) is that it only has access to KK quarks of the flavors that are present in the colliding protons, so it will not be useful in the likely case that the lightest KK quark is a KK top. This leads us to consider also subleading production mechanisms which do not suffer from this problem. In particular, while spin-$1$ KK quarkonia cannot couple to a pair of gluons, they can be produced, similarly to $J/\psi$ and $\Upsilon$~\cite{Barger:1984qg}, in the process
\be
gg \to g(q^\star_\chi \bar {q^\star_\chi})_{\vv{1},J=1}\,,\qq \chi = L \mbox{ or } R
\label{bleaching}
\ee
which is flavor-universal.
The $gg$ and $\gamma\gamma$ KK quarkonium diagrams in figure~\ref{fig-diagrams-KKqKKqbar} describe the same processes as one would have for heavy quarkonia (see~\cite{Kats:2009bv} and references therein for the discussion of toponium and~\cite{Barger:1987xg,Arik:2002nd} for quarkonia of new heavy quarks). The production mechanism~(\ref{bleaching}) is also model-independent (it is determined by interactions with gluons). On the other hand, the $q\bar q$ channel is dominated by a diagram involving the KK gluon which does not exist for quarkonia. Another important difference is that the dominant annihilation process of spin-$1$ heavy quarkonia will be into pairs of longitudinal $W$ bosons because of the quarks' large coupling to the Higgs which is responsible for their large masses~\cite{Barger:1987xg}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.49\textwidth]{cs-jj-M-1f.eps}
\includegraphics[width=0.49\textwidth]{cs-jj-M-allf.eps}
\caption{Dijet signal of the various bound states at the $14$ TeV LHC as a function of the bound state mass ($M = 2m$), where all the particles are assumed to have the same mass $m$. The contributions of antiparticle bound states are included wherever relevant. On the left, we assume that the signal comes from particular flavors and chiralities of KK quarks or squarks and present several examples, while on the right we sum over all the possible combinations.}
\label{fig-dijet-all}
\end{center}
\begin{center}
\includegraphics[width=0.49\textwidth]{cs-tt-M-1f.eps}
\includegraphics[width=0.49\textwidth]{cs-tt-M-allf.eps}
\caption{$t\bar t$ signal of the various bound states at the $14$ TeV LHC as a function of the bound state mass ($M = 2m$), where all the particles are assumed to have the same mass $m$. For KK quarkonia and squarkonia, on the left we present examples of the contribution from KK tops and stops of a single chirality (the results for the other chirality are the same), while on the right we assume that both chiralities are close in mass and contribute.}
\label{fig-ttbar-all}
\end{center}
\end{figure}
The production cross sections for KK quarkonia and squarkonia are given by the expressions in appendices~\ref{app-rates-KKqKKq} and~\ref{app-rates-sqsq} and shown in table~\ref{tab-KKquarkonium}. Overall, the cross sections are about $2$ orders of magnitude below those of gluinonia of the same mass. This is primarily due to the smaller color factors both in $\l|\psi(\vv{0})\r|^2 \propto C^3$ and in the short-distance matrix element. As a result, the dijet and $t\bar t$ annihilation channels, whose cross sections are included in figures~\ref{fig-dijet-all} and~\ref{fig-ttbar-all}, are not promising.
However, the possibility of tree-level annihilation into $\gamma\gamma$ (which has low standard model background) is much more attractive despite the $(\alpha^2/\alpha_s^2)$-suppressed branching ratio of this mode. The cross section for the $\gamma\gamma$ signal is shown in figure~\ref{fig-cs-diphoton-M} as a function of the bound state mass. The angular distributions in this channel are isotropic for both KK quarkonium and squarkonium but the signal will be twice larger in UED (when comparing equal KK quark and squark masses) due to the twice larger production cross section. This property can be used as a discriminator, although it may be important to take higher-order QCD corrections into account. Such corrections for the squarkonium (stoponium) have been studied in~\cite{Martin:2009dj,Younkin:2009zn}. Also, we have assumed the total annihilation rate of squarkonium to be dominated by $gg$, while in some cases annihilation into pairs of $W$, $Z$ or Higgs bosons cannot be neglected~\cite{Herrero:1987df,Barger:1988sp,Drees:1993uw,Martin:2008sv}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.48\textwidth]{cs-dilepton-M.eps}\;\;
\includegraphics[width=0.48\textwidth]{cs-dileptonj-M.eps}
\caption{Dilepton annihilation signal of KK quarkonium at the $14$ TeV LHC as a function of the resonance mass. The cross sections apply for any single flavor of leptons. Left: production via $q\bar q\to\kkq\bar\kkq$ (result independent of $m_\kkg$ unless $m_\kkg \gg m_\kkq$). Right: production via $gg\to\kkq\bar\kkq g$ for $m_\kkg = 2m_\kkq$ (for $m_\kkg = m_\kkq$ the branching ratios would be an order of magnitude smaller).}
\label{fig-cs-dilepton-M}
\end{center}
\end{figure}
The cross sections for the dilepton annihilation channel of spin-$1$ KK quarkonia are shown in figure~\ref{fig-cs-dilepton-M}. The right plot refers to KK quarkonia produced through the subleading process~(\ref{bleaching}) which has
\be
\hat\sigma_{\rm bound}\l(\hat s\r)
= \frac{5\,\zeta(3)\,\pi}{243\,m_\kkq^2}\,\alpha_s^3\bar\alpha_s^3\, I\l(\frac{\hat s}{M^2}\r)
\ee
where as in~\cite{Barger:1984qg}
\be
I(\gamma) = \theta(\gamma-1)\l[\frac{2}{\gamma^2}\l(\frac{\gamma+1}{\gamma-1} - \frac{2\gamma\ln \gamma}{(\gamma-1)^2}\r) + \frac{2(\gamma-1)}{\gamma(\gamma+1)^2} + \frac{4\ln \gamma}{(\gamma+1)^3}\r]
\ee
On the other hand, squarkonia do not give rise to a dilepton signal.
\begin{table}[t]
$$\begin{array}{|c|c|c|c|c|}\hline
\,\mbox{process} &\,J = 0,\, J_z = 0 \,&\, J = 2,\, |J_z| = 2 \,\\\hline\hline
\; gg\to (WW^\ast) \;&\,2/27\,&\,32/81\,\\ \hline
\end{array}$$
\caption{Tripletonium production cross section prefactors $P_{ij}$ of (\ref{parton-cs}).}
\label{tab-tripletonium}
\end{table}
While squarks and KK quarks are examples of spin-$0$ and spin-$1/2$ particles in the $\vv{3}$ representation, it may be interesting to consider also bound states of spin-$1$ particles in the same representation. Such vector color-triplets $W^\mu$ couple to gluons via
\be
\cL = -\frac12W_{\mu\nu}^\ast W^{\mu\nu} - ig_s W_\mu^\ast T^a W_\nu G^{\mu\nu a} + m_W^2 W_\mu^\ast W^\mu
\ee
where $W_{\mu\nu} = D_\mu W_\nu - D_\nu W_\mu$ with $D_\mu = \pd_\mu - ig_s A_\mu^a T^a$ and $G^a_{\mu\nu}$ is the gluon field strength. The second term here is chosen such that tree-level unitarity in the production of vector pairs from $gg$ is preserved at high energies, like in the situation when $W^\mu$ is a gauge boson of an extended gauge group~\cite{Borisov:1986ev}. The $W^\mu$ particles may also couple to photons, which can be described by including the $-ieQ A_\mu$ term in $D_\mu$ and adding an electromagnetic term analogous to the second term above. The diagrams coupling these vector particles to gluons or photons are the same as for squark-antisquark pairs, figure~\ref{fig-diagrams-sqasq}. The $s$-channel diagram does not actually contribute for the bound states. We do not consider production from $q\bar q$ since it would depend on the (model-dependent) couplings of the quarks to the vector bosons. We present the cross sections for the resulting bound states, ``tripletonia'', in table~\ref{tab-tripletonium}. The production cross section from gluons is an order of magnitude bigger than that of squarkonia and KK quarkonia. Furthermore, most of the tripletonia are produced with $J=2$ rather than $J=0$, which leads to different angular distributions of the annihilation products. More details are given in appendix~\ref{app-rates-tripletonia}.
\section{Di-KK quarks vs. di-squarks and KK quark-KK gluon vs. squark-gluino bound states\label{sec-diKKquarks-KKqKKg}}
\setcounter{equation}{0}
\begin{figure}[t]
$$\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 70 37 /
\plot -70 -39 70 -39 /
\barrow from 35 37 to 45 37
\barrow from 35 -39 to 45 -39
\barrow from -37 37 to -25 37
\barrow from -37 -39 to -25 -39
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -32 -54 -47 /
\plot -39 -32 -39 -47 /
\plot -24 -32 -24 -47 /
\plot -9 -32 -9 -47 /
\plot -5 26 10 26 /
\plot -5 11 10 11 /
\plot -5 -4 10 -4 /
\plot -5 -19 10 -19 /
\plot -5 -34 10 -34 /
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\kkq$} at -50 20
\put {$\kkq$} at -50 -20
\put {$\kkg$} at -18 0
\put {$q$} at 60 20
\put {$q$} at 60 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 100 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 0 37 70 -39 /
\plot -70 -39 0 -39 30 -5 /
\plot 40 5 70 37 /
\barrow from 0 37 to 55 -23
\barrow from 40 5 to 55 21
\barrow from -37 37 to -25 37
\barrow from -37 -39 to -25 -39
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -32 -54 -47 /
\plot -39 -32 -39 -47 /
\plot -24 -32 -24 -47 /
\plot -9 -32 -9 -47 /
\plot -5 26 10 26 /
\plot -5 11 10 11 /
\plot -5 -4 10 -4 /
\plot -5 -19 10 -19 /
\plot -5 -34 10 -34 /
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\kkq$} at -50 20
\put {$\kkq$} at -50 -20
\put {$\kkg$} at -18 0
\put {$q$} at 70 20
\put {$q$} at 70 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 100 0
\endpicture
$$
\caption{Diagrams for production or annihilation of a pair of KK quarks. The second diagram is relevant only if the flavors are equal.}
\label{fig-diagrams-KKqKKq}
$$\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot 70 37 0 37 0 -39 70 -39 /
\setdashes
\plot -70 37 0 37 /
\plot -70 -39 0 -39 /
\setsolid
\barrow from -45 37 to -35 37
\barrow from 35 37 to 45 37
\barrow from -45 -39 to -35 -39
\barrow from 35 -39 to 45 -39
\setsolid
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\sq$} at -60 20
\put {$\sq$} at -60 -20
\put {$\go$} at -18 0
\put {$q$} at 60 20
\put {$q$} at 60 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 100 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot 70 37 40 5 /
\plot 30 -5 0 -39 0 37 70 -39 /
\setdashes
\plot -70 37 0 37 /
\plot -70 -39 0 -39 /
\setsolid
\barrow from 0 37 to 55 -23
\barrow from 40 5 to 55 21
\barrow from -37 37 to -25 37
\barrow from -37 -39 to -25 -39
\setsolid
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\sq$} at -60 20
\put {$\sq$} at -60 -20
\put {$\go$} at -18 0
\put {$q$} at 70 20
\put {$q$} at 70 -20
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 100 0
\endpicture
$$
\caption{Diagrams for production or annihilation of a pair of squarks. The second diagram is relevant only if the flavors are equal.}
\label{fig-diagrams-sqsq}
\end{figure}
\begin{table}[t]
$$\begin{array}{|c|c|c|c|}\hline
\,\mbox{process} &\,J = 0,\, J_z = 0 \,&\, J = 1,\, |J_z| = 1 \,&\, J = 1,\, J_z = 0 \,\\\hline\hline
\begin{array}{c}q q\to(\kkqL \kkqL)\\\q\mbox{ or }(\kkqR \kkqR)\end{array} &\, \,& &\, \frac{4}{729}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2\l(\frac{m_\kkq^2}{m_\kkg^2}\r)^2 \\
\begin{array}{c}q q'\to(\kkqL \kkqL')\\\q\mbox{ or }(\kkqR \kkqR')\end{array} &\, \frac{2}{729}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2\l(4 + \frac{m_\kkq^2}{m_\kkg^2}\r)^2 \,& &\, \frac{2}{729}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2\l(\frac{m_\kkq^2}{m_\kkg^2}\r)^2 \\
q q\to(\kkqL \kkqR) & &\, \frac{8}{729}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2\l(2 + \frac{m_\kkq^2}{m_\kkg^2}\r)^2 \,& \\
\begin{array}{c}q q'\to(\kkqL \kkqR')\\\q\mbox{ or }(\kkqR \kkqL')\end{array} & &\, \frac{4}{729}\l(\frac{2m_\kkq^2}{m_\kkq^2 + m_\kkg^2}\r)^2\l(2 + \frac{m_\kkq^2}{m_\kkg^2}\r)^2 \,& \\\hline
\begin{array}{c}q q'\to(\sq_L\sq'_L)\\\q\mbox{ or }(\sq_R\sq'_R)\end{array} & \ds\frac{16}{729}\l(\frac{2m_\go m_\sq}{m_\go^2 + m_\sq^2}\r)^2 & &\, \\\hline
\end{array}$$
\caption{Di-KK quark $(\kkq\kkq)$ (UED) and di-squark $(\sq\sq)$ (MSSM) production cross section prefactors $P_{ij}$ of (\ref{parton-cs}). All the bound states are in the $\vv{\bar 3}$ representation. All the numbers refer to a single flavor of that mass, and a single choice of the chiralities, and our notation assumes that $q' \neq q$. There also exist similar processes in which all the particles are replaced by their antiparticles. For simplicity, we assumed the masses of the different flavors and chiralities of the KK quarks or squarks to be the same.}
\label{tab-diKKquark}
\end{table}
\begin{figure}[t]
$$\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 0 5 0 42 37 /
\barrow from -70 0 to -25 0
\barrow from 5 0 to 28 23
\startrotation by 0.6 -0.6 about -100 35
\plot -100 35 -45 35 /
\barrow from -100 35 to -70 35
\plot -84 28 -84 42 /
\plot -69 28 -69 42 /
\plot -54 28 -54 42 /
\stoprotation
\startrotation by 0.6 0.6 about -100 -35
\plot -84 -30 -84 -45 /
\plot -69 -30 -69 -45 /
\plot -54 -30 -54 -45 /
\ellipticalarc axes ratio 2:1 -150 degrees from -92 -31 center at -99 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -77 -31 center at -84 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\stoprotation
\startrotation by 0.6 -0.6 about 37 -35
\ellipticalarc axes ratio 2:1 150 degrees from 29 -31 center at 36 -35
\ellipticalarc axes ratio 2:1 280 degrees from 14 -31 center at 21 -35
\ellipticalarc axes ratio 2:1 280 degrees from -1 -31 center at 6 -35
\ellipticalarc axes ratio 2:1 280 degrees from -16 -31 center at -9 -35
\stoprotation
\put {$\kkq$} at -70 35
\put {$\kkg$} at -65 -30
\put {$q$} at -35 15
\put {$q$} at 10 30
\put {$g$} at 5 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 70 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 70 37 /
\barrow from -70 37 to -25 37
\barrow from -70 37 to 45 37
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -5 26 10 26 /
\plot -5 11 10 11 /
\plot -5 -4 10 -4 /
\plot -5 -19 10 -19 /
\plot -5 -34 10 -34 /
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 150 degrees from 62 -31 center at 69 -35
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\kkq$} at -55 15
\put {$\kkg$} at -55 -15
\put {$\kkg$} at -18 0
\put {$q$} at 60 20
\put {$g$} at 60 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 37 2 37 2 -39 25 -10 /
\plot 45 10 70 37 /
\barrow from -70 37 to -25 37
\barrow from 2 37 to 2 -2
\barrow from 45 10 to 60 26
\plot -54 30 -54 45 /
\plot -39 30 -39 45 /
\plot -24 30 -24 45 /
\plot -9 30 -9 45 /
\plot -54 -30 -54 -45 /
\plot -39 -30 -39 -45 /
\plot -24 -30 -24 -45 /
\plot -9 -30 -9 -45 /
\plot -5 26 9 26 /
\plot -5 11 9 11 /
\plot -5 -4 9 -4 /
\plot -5 -19 9 -19 /
\plot -5 -34 9 -34 /
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\startrotation by 0.7 -0.75 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -280 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -280 degrees from 77 31 center at 84 35
\ellipticalarc axes ratio 2:1 -150 degrees from 92 31 center at 99 35
\stoprotation
\put {$\kkq$} at -55 15
\put {$\kkg$} at -55 -15
\put {$\kkq$} at -18 0
\put {$q$} at 75 20
\put {$g$} at 75 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
$$
\caption{Diagrams for production or annihilation of a KK quark-KK gluon pair.}
\label{fig-diagrams-KKqKKg}
$$\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\setdashes
\plot -100 35 -65 0 /
\setsolid
\plot -100 -35 -65 0 5 0 40 35 /
\barrow from -70 0 to -25 0
\barrow from 5 0 to 28 23
\barrow from -85 20 to -80 15
\startrotation by 0.6 0.6 about -100 -35
\ellipticalarc axes ratio 2:1 -150 degrees from -92 -31 center at -99 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -77 -31 center at -84 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\stoprotation
\startrotation by 0.6 -0.6 about 37 -35
\ellipticalarc axes ratio 2:1 150 degrees from 29 -31 center at 36 -35
\ellipticalarc axes ratio 2:1 280 degrees from 14 -31 center at 21 -35
\ellipticalarc axes ratio 2:1 280 degrees from -1 -31 center at 6 -35
\ellipticalarc axes ratio 2:1 280 degrees from -16 -31 center at -9 -35
\stoprotation
\put {$\sq$} at -75 35
\put {$\go$} at -70 -30
\put {$q$} at -35 15
\put {$q$} at 10 30
\put {$g$} at 5 -30
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -130 0 to 70 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot 70 35 0 35 0 -35 -70 -35 /
\setdashes
\plot -70 35 0 35 /
\setsolid
\barrow from -45 35 to -35 35
\barrow from 35 35 to 45 35
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\ellipticalarc axes ratio 2:1 220 degrees from 0 -35 center at 10 -35
\ellipticalarc axes ratio 2:1 280 degrees from 17 -31 center at 24 -35
\ellipticalarc axes ratio 2:1 280 degrees from 32 -31 center at 39 -35
\ellipticalarc axes ratio 2:1 280 degrees from 47 -31 center at 54 -35
\ellipticalarc axes ratio 2:1 150 degrees from 62 -31 center at 69 -35
\startrotation by 0 -1 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from -5 29 center at 5 29
\ellipticalarc axes ratio 2:1 -280 degrees from 12 25 center at 19 29
\ellipticalarc axes ratio 2:1 -280 degrees from 27 25 center at 34 29
\ellipticalarc axes ratio 2:1 -280 degrees from 42 25 center at 49 29
\ellipticalarc axes ratio 2:1 -180 degrees from 57 25 center at 64 29
\stoprotation
\put {$\sq$} at -55 15
\put {$\go$} at -55 -15
\put {$\go$} at -18 0
\put {$q$} at 60 15
\put {$g$} at 60 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
\beginpicture
\setcoordinatesystem units <0.571\tdim,0.571\tdim>
\stpltsmbl
\plot -70 -35 0 -35 25 -10 /
\plot 45 10 70 35 /
\setdashes
\plot -70 35 0 35 0 -35 /
\setsolid
\barrow from -45 35 to -35 35
\barrow from 0 7 to 0 -10
\barrow from 45 10 to 55 20
\ellipticalarc axes ratio 2:1 -150 degrees from -62 -31 center at -69 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -47 -31 center at -54 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -32 -31 center at -39 -35
\ellipticalarc axes ratio 2:1 -280 degrees from -17 -31 center at -24 -35
\ellipticalarc axes ratio 2:1 -220 degrees from 0 -35 center at -10 -35
\startrotation by 0.7 -0.75 about 0 30
\ellipticalarc axes ratio 2:1 -220 degrees from 0 35 center at 10 35
\ellipticalarc axes ratio 2:1 -280 degrees from 17 31 center at 24 35
\ellipticalarc axes ratio 2:1 -280 degrees from 32 31 center at 39 35
\ellipticalarc axes ratio 2:1 -280 degrees from 47 31 center at 54 35
\ellipticalarc axes ratio 2:1 -280 degrees from 62 31 center at 69 35
\ellipticalarc axes ratio 2:1 -280 degrees from 77 31 center at 84 35
\ellipticalarc axes ratio 2:1 -150 degrees from 92 31 center at 99 35
\stoprotation
\put {$\sq$} at -55 15
\put {$\go$} at -55 -15
\put {$\sq$} at -18 0
\put {$q$} at 75 15
\put {$g$} at 75 -15
\linethickness=0pt
\putrule from 0 -52 to 0 52
\putrule from -90 0 to 90 0
\endpicture
$$
\caption{Diagrams for production or annihilation of a squark-gluino pair. For bound states, the last diagram does not actually contribute.}
\label{fig-diagrams-sqgo}
\end{figure}
\begin{table}[t]
$$\begin{array}{|c|c|c|}\hline
\,\mbox{process} & J = \frac{3}{2},\, |J_z| = \frac{3}{2} & J = \frac{1}{2},\, |J_z| = \frac{1}{2} \\\hline\hline
\,qg\to(\kkq\kkg)_\vv{3} \,&\, \ds\frac{3\,m_\kkq\l(m_\kkg + 9m_\kkq\r)^2}{32\l(m_\kkg+m_\kkq\r)^3} \,&\, \ds\frac{9\,m_\kkq^3\l(m_\kkg + 9m_\kkq\r)^2}{64\l(m_\kkg+m_\kkq\r)^5}\, \\[12pt]
\,qg\to(\kkq\kkg)_\vv{\bar 6} \,&\, \ds\frac{m_\kkq}{16\l(m_\kkg+m_\kkq\r)} \,&\, \ds \frac{3\,m_\kkq^3}{32\l(m_\kkg+m_\kkq\r)^3} \\[15pt]\hline
\,qg\to(\sq\go)_\vv{3} \,&\, \,&\, \ds\frac{3\,m_\go m_\sq^2\l(m_\go+9m_\sq\r)^2}{32\l(m_\go+m_\sq\r)^5} \,\\[12pt]
\,qg\to(\sq\go)_\vv{\bar 6} \,&\, \,&\, \ds\frac{m_\go m_\sq^2}{16\l(m_\go+m_\sq\r)^3} \,\\\hline
\end{array}$$
\caption{KK quark-KK gluon $(\kkq\kkg)$ (UED) and squark-gluino $(\sq\go)$ (MSSM) bound states production cross section prefactors $P_{ij}$ of (\ref{parton-cs}). All the numbers refer to a single flavor and chirality. There also exist similar processes in which all the particles are replaced by their antiparticles.}
\label{tab-KKquarkKKgluon}
\end{table}
The possible diagrams for the remaining bound states are shown in figures~\ref{fig-diagrams-KKqKKq}, \ref{fig-diagrams-sqsq}, \ref{fig-diagrams-KKqKKg} and~\ref{fig-diagrams-sqgo}. The resulting cross section prefactors are given in tables~\ref{tab-diKKquark} and~\ref{tab-KKquarkKKgluon}. For all of these bound states, the annihilation will be almost entirely into dijets, without any cleaner channels (even with a small branching ratio) to consider. Furthermore, the dijet signal (see figure~\ref{fig-dijet-all}) is typically even smaller than that of the gluinonium (whose signal we will analyze in more detail in the next section), and therefore cannot be seen on top of the QCD background in most scenarios. The only exception is if the mass spectrum of the model is very degenerate to the extent that signals from many different bound states merge into one. This situation is exemplified in the right plot of figure~\ref{fig-dijet-all} which sums over the flavors and chiralities of the KK quarks or squarks, which corresponds to the overly optimistic scenario in which all the flavors and chiralities are sufficiently long-lived and close in mass within $\sim 5\%$ so that the dijet signals of all their bound states merge into a single peak in the invariant mass distribution. Similarly, we may further sum the curves corresponding to different types of bound states. Even then, the dijet signal will be extremely challenging.
\section{Detection prospects\label{sec-simulation}}
\setcounter{equation}{0}
We will now analyze to what extent the bound state signals discussed in the previous sections will be detectable at the LHC.\footnote{LHC signals of gluinonium have been also studied in~\cite{Chikovani:1996bk,Cheung:2004ad,BouhovaThacker:2004nh,BouhovaThacker:2006pj,Kats:2009bv} and squarkonium in~\cite{Drees:1993yr,Drees:1993uw,Martin:2008sv,Younkin:2009zn}.} To that end, it is informative to consider the existing experimental constraints on the masses of the various UED and MSSM particles. Direct limits from the Tevatron Run I constrain the size of the extra dimension in UED as $1/R \gtrsim 300$~GeV~\cite{Macesanu:2002db,Lin:2005ix} (the KK modes masses are $m_1 \sim 1/R$). A much stronger limit can probably be obtained from the analysis of data available today. The most stringent indirect limit, $1/R \gtrsim 600$~GeV, arises from the inclusive radiative $\bar{B}\to X_s\gamma$ decay~\cite{Haisch:2007vb}. With the assumption that the Higgs is not much heavier than $115$~GeV, the same limit is obtained from electroweak observables based on LEP data~\cite{Gogoladze:2006br}. For the MSSM, the gluino mass is constrained by collider searches to $m_\go \gtrsim 400$~GeV, and squarks are constrained to $m_\sq \gtrsim 600$~GeV, except for the sbottom (which can be lighter than $300$~GeV) and the stop (which can even be under $200$~GeV)~\cite{Feng:2009te,Khachatryan:2011tk,Collaboration:2011hh,daCosta:2011qk,Chatrchyan:2011wc,Chatrchyan:2011bz,Aad:2011yf}. It should be remembered though that MSSM and UED have multiple free parameters (the soft SUSY-breaking parameters in MSSM and the boundary terms in UED), and experimental bounds usually depend on certain arbitrary assumptions about them and thus apply in only part of the parameter space (for an example, see~\cite{Alwall:2008ve}). We therefore find it useful to consider also particles that are lighter than the bounds quoted above. Note also that many of our results do not depend on the full particle content of MSSM or UED and can be relevant to bound states in other models that are less constrained.
To simulate the bound state signals and the dominant standard model backgrounds we used \textsc{Pythia} (version 8.120)~\cite{Sjostrand:2006za,Sjostrand:2007gs} with NLO MSTW 2008 PDFs~\cite{Martin:2009iq} and \textsc{SISCone} jet algorithm (version 2.0.1)~\cite{Salam:2007xv} with cone radius $R = 1$ (except for the $t\bar t$ analysis where we used $R=0.5$), overlap parameter $f = 0.75$, no $p_T$ threshold on stable cones, and an infinite number of passes. We selected several particles defined in \textsc{Pythia} and modified their coupling constants and branching ratios such that they would behave according to the bound state effective vertices from tables~\ref{tab-KKgKKg-gogo-couplings} and~\ref{tab-KKqKKaq-sqasq-couplings}. For the purpose of simulation, we pretended all the bound states to be color singlets. We used the BSM Higgs for simulating spin-$0$ bound states, the $Z'$ for spin-$1$ bound states (and $\Upsilon$ for simulating~(\ref{bleaching})), and the KK graviton for spin-$2$ bound states. We have simulated the backgrounds without any $K$-factors since our signals do not include higher-order QCD corrections either. Such corrections to the pair production processes, the bound state wavefunctions and the annihilation processes can be large and sometimes even change the cross section by a factor of $\sim 2$. Part of these corrections have already been computed for some of the MSSM bound states~\cite{Hagiwara:2009hq,Kauth:2009ud,Martin:2009dj,Younkin:2009zn} but none for UED. Our results will need to be re-examined once these corrections are known.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.50\textwidth]{SB-dijet-M.eps}
\includegraphics[width=0.49\textwidth]{lum-dijet-M.eps}
\caption{Dijet channel: signal-to-background ratio (left) and the luminosity required for $3\sigma$ significance (right) for gluinonium (dashed lines) and KK gluonium (solid line) at the $14$ TeV LHC as a function of the resonance mass. For KK gluonium we assume $m_\kkq = m_\kkg$ while for gluinonium we present two cases.}
\label{fig-SB-dijet-M}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.49\textwidth]{SB-bbbar-M.eps}
\includegraphics[width=0.49\textwidth]{lum-bbbar-M.eps}
\caption{$b\bar b$ channel: signal-to-background ratio (left) and the luminosity required for $3\sigma$ significance (right) for gluinonium (dashed lines) and KK gluonium (solid line) at the $14$ TeV LHC as a function of the resonance mass. For KK gluonium we assume $m_\kkq = m_\kkg$ while for gluinonium we present two cases.}
\label{fig-SB-bbbar-M}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.49\textwidth]{SB-ttbar-M.eps}
\includegraphics[width=0.49\textwidth]{lum-ttbar-M.eps}
\caption{Semileptonic $t\bar t$ channel: signal-to-background ratio (left) and the luminosity required for $3\sigma$ significance (right) for gluinonium (dashed lines) and KK gluonium (solid line) at the $14$ TeV LHC as a function of the resonance mass. For KK gluonium we assume $m_\kkq = m_\kkg$ while for gluinonium we present two cases.}
\label{fig-SB-ttbar-M}
\end{center}
\end{figure}
In the analysis of the dijet channel, we require the two hardest jets within $|\eta| < 2.5$ to have $p_T > 2M/5$, the scattering angle in the partonic collision frame to satisfy $\l|\cos\theta\r| < 0.5$, and the dijet invariant mass to be within $\pm 15\%$ from the mass of the resonance. In the $b\bar b$ channel, we require two tagged jets and assume $60\%$ tagging efficiency for $b$-jets, $1\%$ mistag rate for gluon and light quark jets and $15\%$ mistag rate for $c$-jets. Based on~\cite{ATL-PHYS-PUB-2009-018,CMS-PAS-BTV-09-001}, we believe this level of $b$ tagging performance will be realistic at least for resonances with $M \lesssim 1$~TeV. This leaves the standard model $b\bar b$ production as the dominant background. Besides the tagging efficiency factor, we use the same cuts as in the dijet channel. In the $t\bar t$ channel, we consider the standard model top production processes to be the background and use the procedure described in appendix~C of~\cite{Kats:2009bv} to reconstruct the $4$-momenta of the two tops from their semileptonic decay products (which is the situation where one $W$ from $t\to Wb$ decays leptonically and the other hadronically). We then count events in which the $t\bar t$ invariant mass is within $\pm 15\%$ from the mass of the resonance, the tops have $p_T > M/3$, and $\l|\cos\theta\r| < 0.5$.
The results for the dijet, $b\bar b$ and $t\bar t$ channels, respectively, are presented figures~\ref{fig-SB-dijet-M}, \ref{fig-SB-bbbar-M} and~\ref{fig-SB-ttbar-M}. The dijet channel is unlikely to be realistic because of the very small signal-to-background ratio $S/B$ (which implies high sensitivity to systematic errors in modeling the background). In the $b\bar b$ and $t\bar t$ channels the situation regarding $S/B$ is more promising, especially for KK gluonium. As for the fluctuations in the background, the luminosity required for the statistical significance of the signal\footnote{We compute the significance as $S/\sqrt B$, and present the value of integrated luminosity for which $S/\sqrt B = 3$.} is achievable for $m_\go \lesssim 450$~GeV (if $m_\sq \gg m_\go$) or $m_\kkg \lesssim 800$~GeV in the $b\bar b$ channel, and for $m_\go \lesssim 300$~GeV (if $m_\sq \sim m_\go$) or $m_\kkg \lesssim 500$~GeV in the $t\bar t$ channel.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.49\textwidth]{SB-gammagamma-M.eps}
\includegraphics[width=0.49\textwidth]{lum-gammagamma-M.eps}
\caption{Diphoton channel: signal-to-background ratio (left) and the luminosity required for $3\sigma$ significance (right) for gluinonium (thick dashed line), KK gluonium (estimate; thick solid line), squarkonium (thin dashed line) and KK quarkonium (thin solid line) at the $14$ TeV LHC as a function of the resonance mass. For gluinonium we assumed $m_\sq = m_\go$ (and similarly for KK gluonium).}
\label{fig-SB-diphoton-M}
\end{center}
\end{figure}
In the diphoton channel we consider $q\bar q\to\gamma\gamma$ and $gg\to\gamma\gamma$ to be the dominant backgrounds.\footnote{There will also be a contribution from $\gamma$+jet and dijet events in which jets are misidentified as photons. However, for the heavy resonances considered here, this background can be made subdominant by the tight photon identification criteria with only a mild reduction of the signal (see~\cite{CMS-PAS-EXO-10-019} for a recent study by CMS).} We select events in which the two hardest photons within $|\eta| < 1.5$ have $p_T > 50$~GeV and look in the invariant mass window of $\pm 2\%$ around the mass of the resonance. The results are presented in figure~\ref{fig-SB-diphoton-M}. Since gluinonium is unlikely to be lighter than $600$~GeV (corresponding to a $300$~GeV gluino), the luminosity required to see its $\gamma\gamma$ signal is too high. On the other hand, for KK gluonium the signal is promising up to KK gluon masses of $\sim 500$~GeV. For squarkonium, the signal is viable for squark masses $\lesssim 350$~GeV. While typical squarks are expected to be at least as heavy as $\sim 600$~GeV, a stop can be much lighter. Luckily, the stop is also the squark that has the largest chance for being sufficiently long-lived so that stoponium will decay primarily by annihilation. Several models in which this happens were discussed in~\cite{Martin:2008sv}. The stoponium $\gamma\gamma$ signal can thus be within the reach of the LHC. Similarly, a KK quark with mass $\lesssim 400$~GeV can have a viable $\gamma\gamma$ signal from KK quarkonium. Here we assumed that the KK quarks or squarks that form the bound state have charge $Q = 2/3$. For KK quarks or squarks with charge $Q = -1/3$ the cross sections will be $16$ times smaller.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.49\textwidth]{SB-dilepton-M.eps}
\includegraphics[width=0.49\textwidth]{lum-dilepton-M.eps}
\caption{Dielectron channel: signal-to-background ratio (left) and the luminosity required for $3\sigma$ significance (right) for several examples of spin-$1$ KK quarkonia at the $14$ TeV LHC as a function of the resonance mass. The solid curves correspond to the left plot of figure~\ref{fig-cs-dilepton-M} and the dashed curves to the right plot.}
\label{fig-SB-dilepton-M}
\end{center}
\end{figure}
In the dilepton channel, we focus on $e^+e^-$ processes since they will have the best mass resolution. We include Drell-Yan processes as the background. We select events which include two electrons with $|\eta| < 2.5$ and $p_T > 60$~GeV and look in the invariant mass window of $\pm 2\%$ around the mass of the resonance. The resulting reach is shown in figure~\ref{fig-SB-dilepton-M} for several cases of KK quarkonium. The signal is detectable for some KK quarkonia with KK quarks corresponding to the light quark flavors if $m_\kkq \lesssim 500$~GeV. On the other hand, for a perhaps more likely case that $t^\star_R$ is light, the $e^+e^-+\mbox{jet}$ signal of $(t^\star_R\bar{t^\star_R})$ is much less promising, especially since for this case, as in the right plot of figure~\ref{fig-cs-dilepton-M}, we are already considering the more favorable situation of $m_\kkg = 2m_\kkq$ (rather than $m_\kkg \approx m_\kkq$).
\section{Conclusions\label{sec-conclusions}}
\setcounter{equation}{0}
In this paper we have addressed all the possible $S$-wave QCD bound states of pairs of particles from the UED scenario and compared them with analogous bound states of MSSM particles. KK gluonium and gluinonium were also compared with bound states of adjoint scalars, and KK quarkonium and squarkonium with bound states of color-triplet spin-$1$ particles.
We have found that if the KK gluon is sufficiently stable (which is likely to be the case if it is lighter than the KK quarks), its bound states signals at the LHC will have cross sections that are significantly larger than those of the gluinonium of the MSSM, mostly due to the larger number of spin states for KK gluonia. Besides the order-of-magnitude difference in the size of the cross sections, the angular distributions of the annihilation products can be used as a discriminator since KK gluonia will predominantly form with spin $2$ (and some spin $0$), while gluinonia will be mostly spin $0$ (and some spin $1$).
The diphoton signal of spin-$0$ KK quarkonium is twice as large as that of the squarkonium, and is potentially detectable if the constituent KK quarks are sufficiently stable and light. This difference in the cross section can help distinguish between KK quarks and squarks. Another potential discriminator is the dilepton annihilation signal of the vector KK quarkonium which might be measurable.
Despite the fact that bound state cross sections are very small, we found that in some cases the detection of the signal is realistic. For example, a $600$~GeV KK gluon can give rise to a $1.2$~TeV resonance in the $b\bar b$ channel, with $S/B \sim 3\%$ and $3\sigma$ significance reached at $\sim 30\mbox{ fb}^{-1}$. With even more luminosity, the angular distribution can hopefully be extracted and indicate the spin-$2$ nature of the bound state, confirming that the KK gluon has spin $1$. Another example is a $400$~GeV KK top whose bound state will appear as an $800$~GeV resonance in the $\gamma\gamma$ channel with $S/B \sim 10\%$ and $3\sigma$ significance at $\sim 600\mbox{ fb}^{-1}$. Furthermore, many of our results remain valid for other models that contain pair-produced particles with color representations and spins that our study has covered, but which can otherwise be slightly or completely different from the standard MSSM and UED scenarios and have weaker experimental bounds on the masses, which opens up additional possibilities.
Overall, our study has demonstrated how processes of bound state formation and annihilation, which are easy to compute for a given model and easy to reconstruct at the collider, provide additional channels for studying new particles at the LHC.
\section*{Acknowledgments}
We would like to thank Matthew Schwartz for giving us the idea for this project and for many useful discussions. We are also grateful to Matthew Strassler for several important suggestions. Our research is supported by NSF grant PHY-0804450 and DOE grant DE-FG02-96ER40959. Some of the computations in this paper were performed on the Odyssey cluster supported by the FAS Research Computing Group at Harvard University. DK is supported by the General Sir John Monash Award.
|
\section{Introduction}
For the notations, readers are refereed to \cite{Roga1}. So-called quantum operations are completely positive (CP) and trace-preserving (TP) linear super-operator $\Phi$, it is also called quantum channel, stochastic CP super-operator, by which the decoherence induced in an $d$-level quantum system $\cH_d$ are described by the map entropy $\S^{\mathrm{map}}(\Phi)$. When stochastic CP super-operator is unit-preserving, it is called bi-stochastic. In particular, depolarizing channels own an important role, for instance, Roga and his colleagues in \cite{Roga1} obtained a result which states that \emph{among all quantum channels with a given minimal Renyi output entropy of order two the depolarizing channel has the smallest map entropy Renyi entropy}. The present notes aims to give another proof, based on a theorem in \cite{Pedersen}. For convenience, it is listed as follows:
\begin{lem}(\cite{Pedersen})\label{average}
For any linear operator $M$ on an $n$-dimensional complex Hilbert space, the uniform average of $|\innerm{\psi}{M}{\psi}|^2$
over state vectors $\ket{\psi}$ on the unit sphere $S^{2n-1}$ in $\Complex^n$ is given by\\ \indent
$\int_{S^{2n-1}}|\innerm{\psi}{M}{\psi}|^2 d\mu(\psi)=\frac{1}{n(n+1)}[\tr{MM^\dagger}+|\tr{M}|^2]$,\\
where $d\mu$ is the normalized measure on the sphere.
\end{lem}
\section{Main Results}
\begin{prop}\label{prop:1}
If $\S_2^{\mathrm{min}}(\Lambda_n)=\S_2^{\mathrm{min}}(\Phi)$, then $\S_2^{\mathrm{map}}(\Lambda_n)\leqslant\S_2^{\mathrm{map}}(\Phi)$.
\end{prop}
\begin{proof}
Now let\\ \indent
$\S_2^{\mathrm{min}}(\Phi)=\S_2^{\mathrm{min}}(\Lambda_n)=-\log(1-\epsilon)\qquad\forall\epsilon\in(0,1)$.\\
Then\\ \indent
$\S_2^{\mathrm{map}}(\Lambda_n)=-\log[1-(1+\frac1n)\epsilon]$.\\
It suffices to prove $\S_2^{\mathrm{map}}(\Phi)\geqslant-\log[1-(1+\frac1n)\epsilon]$ whenever $\S_2^{\mathrm{min}}(\Phi)=-\log(1-\epsilon)$. This is equivalent to the following statement:\\ \indent
$\tr{[\Phi(\out{\varphi}{\varphi})]^2}\leqslant1-\epsilon \Longrightarrow \tr{[\rho(\Phi)]^2}\leqslant1-(1+\frac1n)\epsilon$,\\
where $\rho(\Phi)=\frac1n \jam{\Phi}$ is the Jamio{\l}kowski state and $\jam{\Phi}=\Phi\ot\identity(\out{\identity}{\identity})$ is the dynamical matrix of quantum channel $\Phi$.
Now by the Kraus decomposition of $\Phi$:\\ \indent
$\Phi(\sigma)=\sum_{i}K_i \sigma K_i^\dagger,\quad\sum_i K_i^\dagger K_i=\identity$.\\
Thus\\ \indent
$\tr{[\Phi(\out{\varphi}{\varphi})]^2}= \sum_{i,j}|\innerm{\varphi}{K_i^\dagger K_j}{\varphi}|^2\leqslant1-\epsilon,\quad \forall \ket{\varphi}$\\
and\\ \indent
$\tr{[\rho_{\Phi}]^2}=\frac{1}{n^2}\sum_{i,j}|\tr{K_i^\dagger K_j}|^2$.\\
It follows from the Lemma \ref{average} that
\begin{eqnarray}
\nonumber&&\int_{S^{2n-1}}\sum_{i,j}|\innerm{\varphi}{K_i^\dagger K_j}{\varphi}|^2 d\mu(\varphi)\leqslant(1-\epsilon)\int_{S^{2n-1}}d\mu(\varphi)=1-\epsilon\\
\nonumber&\Longleftrightarrow&
\sum_{i,j}\int_{S^{2n-1}}|\innerm{\varphi}{K_i^\dagger K_j}{\varphi}|^2 d\mu(\varphi)
\\
\nonumber&=&\frac{1}{n(n+1)}\sum_{i,j}\left[\tr{[K_i^\dagger K_j][K_i^\dagger K_j]^\dagger}+|\tr{K_i^\dagger K_j}|^2\right]
\\&=&\frac{1}{n(n+1)}\sum_{i,j}\left[\tr{K_i K_i^\dagger K_j K_j^\dagger}+|\tr{K_i^\dagger K_j}|^2\right]\leqslant1-\epsilon.\label{Eqn1}
\end{eqnarray}
Therefore it follows from Schwarcz's inequality that
\begin{eqnarray}
\nonumber n^2&=&[\tr{\identity}]^2=[\Tr{\sum_{i} K_i^\dagger K_i}]^2=[\Tr{\sum_{i}K_i K_i^\dagger}]^2\\
\nonumber&\leqslant&\tr{\identity}\Tr{\left[\sum_{i}K_i K_i^\dagger\right]^\dagger\left[\sum_{j}K_j K_j^\dagger\right]}\\
\nonumber&=&n\sum_{i,j}\tr{K_i K_i^\dagger K_j K_j^\dagger}\\
&\Longleftrightarrow& \sum_{i,j}\tr{K_i K_i^\dagger K_j K_j^\dagger}\geqslant n.\label{Eqn2}
\end{eqnarray}
Combining Eq.~(\ref{Eqn1}) with Eq.~(\ref{Eqn2}) gives that\\ \indent
$\frac{1}{n^2}\sum_{i,j}|\tr{K_i^\dagger K_j}|^2\leqslant1-(1+\frac1n)\epsilon$.
\end{proof}
\begin{remark}
In \cite{Roga1}, Roga \emph{et al.} used the following fact that let $\mu$ be the Haar measure on the unitary group $\mathbf{U}(n)$ and let $A$ be a $n^2\times n^2$ matrix, then
$$\int_{\mathbf{U}(n)}U^{\ot 2}A(U^\dagger)^{\ot 2}d\mu(U)=\left(\frac{\tr{A}}{n^2-1}-
\frac{\tr{AS}}{n(n^2-1)}\right)^2\identity-\left(\frac{\tr{A}}{n(n^2-1)}-\frac{\tr{AS}}{n^2-1}\right)^2 S,$$
where $S$ stands for the swap operator: $S\ket{\mu\nu}=\ket{\nu\mu}$.
\end{remark}
\begin{cor}
If $\S_2^{\mathrm{map}}(\Lambda_n)=\S_2^{\mathrm{map}}(\Phi)$, then $\S_2^{\mathrm{min}}(\Lambda_n)\geqslant\S_2^{\mathrm{min}}(\Phi)$.
\end{cor}
\begin{proof}
Since the function $f(t)=-\log\left(\frac{1+n e^{-t}}{n+1}\right)(n\geqslant2)$ is a contiguous, increasing function, it follows that\\ \indent
$\S_2^{\mathrm{min}}(\Lambda_n)=-\log\left(\frac{1+n e^{-\S_2^{\mathrm{map}}(\Lambda_n)}}{n+1}\right)$;\\
i.e., $\S_2^{\mathrm{min}}(\Lambda_n)$ is a monotonic increasing function of $\S_2^{\mathrm{map}}(\Lambda_n)$. This fact together the above proposition \ref{prop:1} implies that the conclusion.
\end{proof}
It will be useful throughout the present paper to make use of a simple correspondence between the spaces $\lin{\cH,\cK}$ and $\cK\ot\cH$, for given Hilbert spaces $\cH$ and $\cK$. The mapping \\ \indent $\mbox{vec}:\lin{\cH,\cK}\longrightarrow\cK\ot\cH$\\
can be defined to be the linear mapping: \\ \indent$\col{A}=\sum_{m\mu}A_{m\mu}\ket{m}\ot\ket{\mu}=\sum_{m\mu}A_{m\mu}\ket{m\mu}$\\
for every operator $A\in\lin{\cH,\cK}$, where $\set{\ket{m}}$ and $\set{\ket{\mu}}$ are the standard basis for $\cH$ and $\cK$, respectively. When the vec mapping is generalized to multipartite spaces, caution should be given to the bipartite case (multipartite situation similarly). Specifically, for given complex Euclidean spaces $\cH_{A/B}$ and $\cK_{A/B}$,\\ \indent
$\mbox{vec}:\lin{\cH_{A}\ot\cH_{B},\cK_{A}\ot\cK_{B}}\longrightarrow
\cK_{A}\ot\cH_{A}\ot\cK_{B}\ot\cH_{B}$\\
is defined to be the linear mapping that represents a change of bases from the standard basis of $\lin{\cH_{A}\ot\cH_{B},\cK_{A}\ot\cK_{B}}$ to the standard basis of $\cK_{A}\ot\cH_{A}\ot\cK_{B}\ot\cH_{B}$. Concretely, \\ \indent
$\col{\out{m}{n}\ot\out{\mu}{\nu}}:=\ket{mn}\ot\ket{\mu\nu}$, \\ where $\set{\ket{n}}$ is the standard basis for $\cH_{A}$ and $\set{\ket{\nu}}$ is the standard basis for $\cH_{B}$, while $\set{\ket{m}}$ is the standard basis for $\cK_{A}$ and $\set{\ket{\mu}}$ is the standard basis for $\cK_{B}$. The mapping is determined for every operator $X\in\lin{\cH_{A}\ot\cH_{B},\cK_{A}\ot\cK_{B}}$ by linearity. Note that if $X=A\ot B$, where $A\in\lin{\cH_{A},\cK_{A}}$ and $B\in\lin{\cH_{B},\cK_{B}}$, then $\col{A\ot B}=\col{A}\ot\col{B}$.
\begin{prop}\label{product channel}
For any CP super-operators $\Phi$ and $\Psi$, with corresponding their Kraus representations: $\Phi=\sum_{i}Ad_{M_{i}}$ and $\Psi=\sum_{j}Ad_{N_{j}}$, respectively. Then:
\begin{enumerate}
\item $\Phi\ot\Psi=\sum_{i,j}Ad_{M_{i}\ot N_{j}}$, and $\jam{\Phi\ot\Psi}=\jam{\Phi}\ot \jam{\Psi}$;
\item $\jam{\Phi\circ\Psi}=\Phi\ot\identity(\jam{\Psi})=\identity\ot\Psi^\trans(\jam{\Phi})
=\Phi\ot\Psi^\trans(\col{\identity}\row{\identity})$.
\end{enumerate}
\end{prop}
\begin{prop}
Let $\Phi$ and $\Psi$ are CP stochastic super-operators. For any $p\geqslant0$, the Renyi map entropy satisfies the following addition identity:\\ \indent
$\S^{\mathrm{map}}_p(\Phi\ot\Psi)=\S^{\mathrm{map}}_p(\Phi)+\S^{\mathrm{map}}_p(\Psi)$.
\end{prop}
\begin{proof}
By the Lemma \ref{product channel}, $\rho(\Phi\ot\Psi)=\rho(\Phi)\ot\rho(\Psi)$, which implies that $\|\rho(\Phi)\ot\rho(\Psi)\|_p=\|\rho(\Phi)\|_p\|\rho(\Psi)\|_p$ since $\|\cdot\|_p$ is multiplicative.
Now
\begin{eqnarray*}
\S^{\mathrm{map}}_p(\Phi\ot\Psi)&=&\frac{p}{1-p}\log\|\rho(\Phi)\ot\rho(\Psi)\|_p\\
&=&\frac{p}{1-p}\log\|\rho(\Phi)\|_p+\frac{p}{1-p}\log\|\rho(\Psi)\|_p\\
&=&\S^{\mathrm{map}}_p(\Phi)+\S^{\mathrm{map}}_p(\Psi).
\end{eqnarray*}
\end{proof}
\begin{lem}\label{dualchannel}(\cite{Lin})
Let $\Phi$ be CP super-operator with their Kraus operators $\set{K_i}$. The corresponding transpose of $\Phi$ is $\Phi^{\trans}$ for which the Kraus operators $\sec{K_i^{\trans}}$; the dual of $\Phi$ is $\Phi^\dagger$ for which the Kraus operators $\set{K_i^\dagger}$. Then\\ \indent
$\jam{\Phi^\trans}=S\jam{\Phi}S$ and $\jam{\Phi^{\dagger}}=S\jam{\Phi}^{\trans}S$,\\
where $S$ is the swap operator.
\end{lem}
\begin{lem}(\cite{Roga2})
Let $\Phi,\Psi$ be CP bi-stochastic super-operators. Then the dynamical subadditivity inequality is valid:\\ \indent
$\max\set{\S^{\mathrm{map}}(\Phi),\S^{\mathrm{map}}(\Psi)}\leqslant
\S^{\mathrm{map}}(\Phi\circ\Psi)\leqslant\S^{\mathrm{map}}(\Phi)+\S^{\mathrm{map}}(\Psi)$.
\end{lem}
\begin{proof} In the present proof, a simple approach is given, which is different from the one in \cite{Roga2}. Specifically, since $\Phi$ is bi-stochastic, it is easily seen that \\ \indent
$\S(\Phi\circ\Psi)=\S(\Phi\ot\identity(\rho_{\Psi}))\geqslant\S(\Psi),
\S(\Phi\circ\Psi)=\S(\identity\ot\Psi^\trans(\rho_{\Phi}))\geqslant\S(\Phi)$.\\
Now from the Lindblad's entropic inequality; i.e.,\\ \indent
$\abs{\S(\hat{\sigma}(\rho,\Lambda))-\S(\rho)}\leqslant
\S(\Lambda(\rho))\leqslant\S(\hat{\sigma}(\rho,\Lambda))+\S(\rho)$,\\
it follows that\\ \indent
$\S(\Phi\ot\identity(\rho_{\Psi}))\leqslant\S(\rho_{\Psi})+
\S(\hat{\sigma}(\rho_{\Psi},\Phi\ot\identity))\Longleftrightarrow
\S(\Phi\circ\Psi)\leqslant\S(\Phi)+\S(\Psi)$, \\
where $\S(\hat{\sigma}(\rho_{\Psi},\Phi\ot\identity))=\S(\Phi)$ can be easily checked.
\end{proof}
\begin{prop}
For any bi-stochastic CP super-operators $\Phi,\Psi$ and any maximally entangled state $\frac{1}{\sqrt{n}}\col{U}$, where $U$ is a $n\times n$ complex unitary matrix, the following inequality for von Neumann entropy holds:\\ \indent
$\max\set{\S^{\mathrm{map}}(\Phi),\S^{\mathrm{map}}(\Psi)}\leqslant
\S((\Phi\ot\Psi)(\col{U}\row{U}))\leqslant\S^{\mathrm{map}}(\Phi)+\S^{\mathrm{map}}(\Psi)$
\end{prop}
\begin{proof}
It suffices to show that when $U=\identity$. Thus let $U=\identity$. Clearly, it follows from the Lemma \ref{dualchannel} that $\S^{\mathrm{map}}(\Psi)=\S^{\mathrm{map}}(\Psi^\trans)$ when $\Psi$ are bi-stochastic (Therefore so does $\Psi^\trans$). Since\\ \indent
$(\Phi\ot\Psi)(\col{\identity}\row{\identity})=
(\Phi\circ\Psi^{\trans}\ot\identity)(\col{\identity}\row{\identity})=\jam{\Phi\circ\Psi^\trans}$,\\
which implies that \\ \indent
$\S((\Phi\ot\Psi)(\col{\identity}\row{\identity}))=\S^{\mathrm{map}}(\Phi\circ\Psi^\trans)$.\\
Now under the condition that $\Phi,\Psi$ are bi-stochastic, it is obtained from \cite{Roga2} that\\ \indent
$|\S^{\mathrm{map}}(\Phi)-\S^{\mathrm{map}}(\Psi)|\leqslant\max\set{\S^{\mathrm{map}}(\Phi),\S^{\mathrm{map}}(\Psi)}
\leqslant\S^{\mathrm{map}}(\Phi\circ\Psi^\trans)
\leqslant\S^{\mathrm{map}}(\Phi)+\S^{\mathrm{map}}(\Psi)$.
\end{proof}
\begin{remark}
If the Lindblad's inequality employed, then
\begin{eqnarray*}
|\S^{\mathrm{map}}(\Phi)-\S^{\mathrm{map}}(\Psi)|&=&|\S(\rho(\Phi))-\S(\widehat{\sigma}[\identity\ot\Psi,\rho(\Phi)])|
\leqslant\S((\identity\ot\Psi)(\rho(\Phi)))\\
&\leqslant&\S(\rho(\Phi))+\S(\widehat{\sigma}[\identity\ot\Psi,\rho(\Phi)])
=\S^{\mathrm{map}}(\Phi)+\S^{\mathrm{map}}(\Psi).
\end{eqnarray*}
Thus the result can be generalized to the CP stochastic maps \cite{Roga1}.
\end{remark}
\begin{remark}
There is a \emph{conjecture} which can be stated as follows: if $\Phi,\Psi\in\T(\cH)$ are bi-stochastic CP super-operators, then \\ \indent
$\S(\rho)+\S(\Phi\circ\Psi(\rho))\leqslant\S(\Phi(\rho))+\S(\Psi(\rho))$ \\
for any $\rho\in\D(\cH)$, where $\D(\cH)$ stands for all density matrix acting on a $N$-dimensional Hilbert space $\cH$ and $\T(\cH)$ all linear super-operators from $\lin{\cH}$ to $\lin{\cH}$.
And what is a characterization of the saturation for the above inequality. If this conjecture is correct, then it can be employed to give a simple proof to the strong dynamical subadditivity for bi-stochastic CP super-operators.
\end{remark}
|
\section{Introduction}
Several astrophysical observations of distant type Ia supernovae have revealed that the content of the universe is made of about $70\%$ of dark energy, $25\%$ of dark matter and $5\%$ of baryonic (visible) matter (Riess et al. 1998, Perlmutter et al. 1999, de Bernardis et al. 2000, Hanany et al. 2000). Thus, the overwhelming preponderance of matter and energy in the universe is believed to be dark i.e. unobservable by telescopes. The dark energy is responsible for the accelerated expansion of the universe. Its origin is mysterious and presumably related to the cosmological constant. Dark energy is usually interpreted as a vacuum energy and it behaves like a fluid with negative pressure. Dark matter also is mysterious. The suggestion that dark matter may constitute a large part of the universe was raised by Zwicky in 1937. He realized that some mass was ``missing'' in order to account for observations. This missing mass problem was confirmed later by more accurate measurements (Borriello \& Salucci 2001). The rotation curves of neutral hydrogen clouds in spiral galaxies measured from the Doppler effect are found to be roughly flat with a typical rotational velocity $v_\infty\sim 200\, {\rm km/s}$ up to the maximum observed radius of about $50$ kpc. This mass profile is much more extended than the distribution of starlight which typically converges within $\sim 10$ kpc. This implies that galaxies are surrounded by an extended halo of dark matter whose mass $M(r)\sim r v_\infty^2/G$ increases linearly with radius.
Although some authors like Milgrom (1983) propose a modification of Newton's law (MOND theory) to explain the rotation curves of spiral galaxies without invoking dark matter, the dark matter hypothesis is favored by most astrophysicists.
The nature of dark matter (DM) is one of the most important puzzles in modern physics and cosmology. A wide ``zoology'' of exotic particles that could form dark matter has been proposed. In particular, many grand unified theories in particle physics predict the existence of various exotic bosons (e.g. axions, scalar neutrinos, neutralinos) that should be present in considerable abundance in the universe and comprise (part of) the cosmological missing mass (Primack et al. 1998, Overduin \& Wesson 2004). Even if the bosonic particles have never been detected in accelerator experiments, they are considered as leading candidates of dark matter and might play a significant role in the evolution and in the structure of the universe.
If dark matter is made of bosons, they should have formed compact gravitating Bose-Einstein condensates (BEC) such as boson stars. Boson stars were introduced by Kaup (1968) and Ruffini \& Bonazzola (1969) in the sixties. Early works on boson stars (Thirring 1983, Breit et al. 1984, Takasugi \& Yoshimura 1984, van der Bij \& Gleiser 1987) were motivated by the axion field that was proposed as a possible solution to the strong CP problem in QCD. For particles with mass $m\sim 1\, {\rm GeV/c^2}$, the maximum mass of a boson star, called the Kaup mass, is much smaller than the solar mass ($M_{Kaup}\sim 10^{-19}M_\odot$!) so that these {\it mini boson stars}, like axion black holes, are not very astrophysically relevant. Some authors (Baldeschi et al. 1983, Sin 1994, Hu et al. 2000) have proposed that dark matter halos could be giant systems of ``Bose liquid'' but in that case the mass of the bosons must be extremely small ($m\sim 10^{-24}\, {\rm eV}$) to yield masses consistent with the mass of galactic halos. Such an ultralight scalar field was called ``fuzzy cold dark matter'' (FCDM) by Hu et al. (2000) who discussed its overall cosmological behavior. On the other hand, Colpi et al. (1986) have shown that if the bosons have a self-interaction, then the mass of the boson stars can considerably increase, even for a small self-interaction. For $m\sim 1\, {\rm GeV/c^2}$ it becomes of the order of the solar mass so that dark matter could be made of numerous boson stars. On the other hand, for $m\sim 1\, {\rm eV/c^2}$, the mass of the boson stars becomes of the order of the mass of the galactic halos. Therefore, some authors (Lee \& Koh 1996, Peebles 2000, Goodman 2000, Arbey et al. 2003, B\"ohmer \& Harko 2007) proposed that dark matter halos themselves could be in the form of gigantic self-gravitating Bose-Einstein condensates with short-range interactions described by a single wave function $\psi({\bf r},t)$. In the Newtonian limit, which is relevant at the galactic scale, the evolution of this wave function is governed by the Gross-Pitaevskii-Poisson (GPP) system (B\"ohmer \& Harko 2007). Using the Madelung (1927) transformation, the GP equation turns out to be equivalent to hydrodynamic (Euler) equations involving an isotropic classical pressure due to short-range interactions (scattering) and an anisotropic quantum pressure (or a quantum potential) arising from the Heisenberg uncertainty principle. For a standard BEC with quartic self-interaction, the equation of state is that of a polytrope with index $n=1$. At large scales, scattering and quantum effects are negligible and one recovers the classical hydrodynamic equations of cold dark matter (CDM) models which are remarkably successful in explaining the large-scale structure of the universe. At small scales, gravitational collapse is prevented by the (repulsive) scattering or by the uncertainty principle. This may be a way to solve the problems of the CDM model such as the cusp problem and the missing satellite problem (Hu et al. 2000).
In our previous papers (Chavanis 2011, Chavanis \& Delfini 2011), we performed an exhaustive study of the equilibrium configurations of a Newtonian self-gravitating BEC with short-range interactions. For a given value of the scattering length, we obtained the mass-radius relation $M(R)$ connecting the non-interacting case (corresponding to small masses) studied by Ruffini \& Bonazzola (1969) to the Thomas-Fermi (TF) limit (corresponding to large masses) investigated by B\"ohmer \& Harko (2007). We also considered the case of attractive self-interaction. This corresponds to a negative scattering length ($a_s<0$) yielding a negative pressure. In that case, we found the existence of a maximum mass $M_{max}=1.012\hbar/\sqrt{|a_s|Gm}$ above which the system becomes unstable. It turns out that this mass is ridiculously small (being possibly as small as the Planck mass $M_P=2.18\, 10^{-8} {\rm kg}!$) meaning that a self-gravitating BEC with attractive short-range interactions is extremely unstable (except if the mass of the bosons is extraordinarily small like in Hu et al. 2000). We proposed that the scattering length could be negative in the early universe and that it could help for the formation of structures\footnote{This is a potentially interesting idea because, in an expanding universe, the condensation process is often regarded to be too slow to account for the formation of structures (Bonnor 1957, Harrison 1967). An attractive self-interaction could substantially enhance the instability.}. Then, the scattering length could become positive and help to stabilize the structures against gravitational collapse. Of course, the mechanism by which the scattering length changes sign remains to be established so that this idea is highly speculative. We may note that some atoms in terrestrial BEC experiments are reported to have negative scattering lengths (Dalfovo et al. 1999). Therefore, the possibility of negative pressure for a BEC can be contemplated. Furthermore, it has been experimentally demonstrated that, under certain conditions, it is possible to manipulate the sign and value of the scattering length (Fedichev et al. 1996). The extension of these ideas to cosmology remains an (important) problem that we shall not discuss further here. {\it We shall assume that the dark matter in the universe is a BEC with repulsive ($a_s>0$) or attractive ($a_s<0$) self-interaction and theoretically explore the consequences of this hypothesis.}
If dark matter halos are BECs, they have probably formed by Jeans instability. The gravitational instability of a scalar field equivalent to a BEC was considered by Khlopov et al. (1985), Bianchi {et al} (1990), Hu et al. (2000) and Sikivie \& Yang (2009). However, these authors started their analysis from relativistic field equations and did not take into account the self-interaction of the particles. In our previous paper (Chavanis 2011), we studied the Jeans instability of a self-gravitating BEC with short-range interactions described by the Gross-Pitaevskii-Poisson system in a static universe. We considered both attractive and repulsive self-interaction and found that when the scattering length is negative the growth rate of the instability increases with respect to the case where the scattering length is positive or zero. The next step is to study the Jeans instability of a BEC in an expanding universe. This is the topic of the present paper.
In the first part of the paper (Secs. \ref{sec_gpp}-\ref{sec_eds}), we neglect relativistic effects and use the equations of Newtonian cosmology introduced by Milne (1934) and McCrea \& Milne (1934). We study the linear development of perturbations of a self-gravitating BEC with short-range interactions in an expanding Einstein-de Sitter (EdS) universe. In the TF approximation, the equation of state of a BEC with quartic self-interaction is that of a polytrope of index $\gamma=2$ ($n=1$). In the non-interacting case, we find that the BEC behaves similarly to a polytrope of index $\gamma=5/3$ ($n=3/2$) but with a quantum mechanically modified Jeans length. In the two cases, the equation for the density contrast can be solved analytically in terms of Bessel functions.
In the second part of the paper (Sec. \ref{sec_pressure}), we take
special relativistic effects into account and use the equations of
Newtonian cosmology with pressure introduced by McCrea (1951). In that
case, the evolution of the cosmic fluid depends on the equation of
state. We consider a BEC dark matter with equation of state $p=k\rho
c^2$ and add the contribution of radiation, baryons and dark energy
(cosmological constant). For a BEC dark matter with repulsive
self-interaction $k>0$ (positive pressure) the scale factor increases
more rapidly than in the standard $\Lambda$CDM model where dark matter
is pressureless ($k=0$) while for a BEC dark matter with attractive
self-interaction $k<0$ (negative pressure) it increases less rapidly.
We study the linear development of the perturbations in these two
cases and show that the perturbations grow faster in a BEC dark matter
than in a pressureless dark matter.
In the third part of the paper (Sec. \ref{sec_dark}), we consider a
``dark fluid'' with a generalized equation of state $p=(\alpha
\rho+k\rho^2)c^2$ having a component $p=k\rho^2 c^2$ similar to a BEC
dark matter and a component $p=\alpha\rho c^2$ mimicking the effect of
the cosmological constant (dark energy). We find optimal parameters
that give a good agreement with the standard $\Lambda$CDM model.
While our series of papers on this subject was in course of redaction,
Harko (2011) published a paper where he also considered the formation
of structures in an expanding universe made of BEC dark matter. Our
contribution is complementary to Harko's work and confirms his main
results. In addition, we address the following issues: (i) we study
the effect of the quantum pressure in the first part of the paper;
(ii) we use different relativistic hydrodynamic equations to model the
cosmic fluid in the second part of the paper; (iii) we study a
generalized equation of state in the third part of the paper; (iv) we
consider both positive and negative scattering lengths throughout the
paper. The papers of B\"ohmer \& Harko (2007) and Harko (2011) show
that a cosmic BEC can be an interesting model for the dark matter of
the universe. Our series of papers, which develop this idea, go in the
same direction.
Finally, in a different perspective, Widrow \& Kaiser (1993) have proposed to describe a classical collisionless self-gravitating system by the Schr\"odinger-Poisson system. In this approach, the constant $\hbar$ is not the Planck constant, but rather an adjustable parameter that controls the spatial resolution $\lambda_{deB}$ through a de Broglie relation $\lambda_{deB}=\hbar/mv$. It is argued that when $\hbar\rightarrow 0$, the Vlasov-Poisson system is recovered and that a finite value of $\hbar$ provides a small-scale regularization of the dynamics. In that case, the Schr\"odinger-Poisson system has nothing to do with quantum mechanics since it aims at describing the evolution of classical collisionless matter under the influence of gravity (in static or expanding universes). Still, on a mathematical point of view, these equations are equivalent to those describing self-gravitating BECs without self-interaction. Therefore, the results of the first part of our paper can have application in that context, independently of quantum mechanics.
\section{The Gross-Pitaevskii-Poisson system}
\label{sec_gpp}
Following B\"ohmer \& Harko (2007), we assume that dark matter is a Bose-Einstein condensate. A self-gravitating BEC with short-range interactions is described by the Gross-Pitaevskii-Poisson system
\begin{equation}
\label{gpp1}
i\hbar \frac{\partial\psi}{\partial t}=-\frac{\hbar^2}{2m}\Delta\psi+m(\Phi+h(\rho))\psi,
\end{equation}
\begin{equation}
\label{gpp2}
\Delta\Phi=4\pi G Nm |\psi|^2-\Lambda,
\end{equation}
where $\rho=Nm|\psi|^2$ is the density, $\Phi$ the gravitational
potential and $h(\rho)=\int u_{SR}({\bf r}-{\bf r}')\rho({\bf r}',t)\,
d{\bf r}'$ an effective potential taking into account small-scale
interactions. For the sake of generality, we have included the
cosmological constant $\Lambda$ in the Poisson equation\footnote{We
may note that the cosmological constant $\Lambda$ plays the same role
as a global rotation $\Omega$ of the universe if $\Phi$ is interpreted
as an effective gravitational potential
$\Phi_{eff}=\Phi-\frac{1}{2}({\bf \Omega}\times {\bf r})^2$ in the
rotating frame. Indeed, the Poisson equation becomes
$\Delta\Phi_{eff}=4\pi G \rho-2\Omega^2$ which allows the
identification $\Lambda=2\Omega^2$.}. We write the wave function in
the form $\psi({\bf r},t)=A({\bf r},t)e^{iS({\bf r},t)/\hbar}$ where
$A$ and $S$ are real, and make the Madelung (1927) transformation
\begin{equation}
\label{gpp3}
\rho=Nm|\psi|^2=NmA^2, \qquad {\bf u}=\frac{1}{m}\nabla S,
\end{equation}
where $\rho({\bf r},t)$ is the density field and ${\bf u}({\bf r},t)$ the velocity field. We note that the flow is irrotational since $\nabla\times {\bf u}={\bf 0}$. With this transformation, it can be shown that the Gross-Pitaevskii equation (\ref{gpp1}) is equivalent to the barotropic Euler equations with an additional term $Q$ called the quantum potential (or quantum pressure). Indeed, one obtains the set of equations
\begin{equation}
\label{gpp4}
\frac{\partial\rho}{\partial t}+\nabla\cdot (\rho {\bf u})=0,
\end{equation}
\begin{equation}
\label{gpp5}
\frac{\partial {\bf u}}{\partial t}+({\bf u}\cdot \nabla){\bf u}=-\frac{1}{\rho}\nabla p-\nabla\Phi-\frac{1}{m}\nabla Q,
\end{equation}
\begin{equation}
\label{gpp6}
\Delta\Phi=4\pi G \rho-\Lambda,
\end{equation}
with
\begin{equation}
\label{gpp7}
Q=-\frac{\hbar^2}{2m}\frac{\Delta \sqrt{\rho}}{\sqrt{\rho}}.
\end{equation}
The pressure $p(\rho)$ is determined by the effective potential $h(\rho)$, playing the role of an enthalpy (Chavanis 2011), through the relation $p'(\rho)=\rho h'(\rho)$. For a contact pair interaction $u_{SR}({\bf r}-{\bf r}')=g\delta({\bf r}-{\bf r}')$, the effective potential $h(\rho)=g\rho$ where $g={4\pi a_s\hbar^2}/{m^3}$ is the pseudo-potential and $a_s$ is the s-scattering length (Dalfovo et al. 1999). For the sake of generality, we allow $a_s$ to be positive or negative. The corresponding equation of state is
\begin{equation}
\label{gpp8}
p=\frac{2\pi a_s\hbar^2}{m^3}\rho^{2}.
\end{equation}
It corresponds to a polytropic equation of state of the form
\begin{equation}
\label{gpp9}
p=K\rho^{\gamma},\qquad \gamma=1+\frac{1}{n},
\end{equation}
with polytropic index $n=1$ (i.e. $\gamma=2$) and polytropic constant $K={2\pi a_s\hbar^2}/{m^3}$.
The effective potential corresponding to the polytropic equation of state (\ref{gpp9}) is
\begin{equation}
\label{gpp10}
h(\rho)=\frac{K\gamma}{\gamma-1}\rho^{\gamma-1}.
\end{equation}
This leads to a GP equation of the form
\begin{equation}
\label{gpp11}
i\hbar \frac{\partial\psi}{\partial t}=-\frac{\hbar^2}{2m}\Delta\psi+m(\Phi+\kappa |\psi|^{2/n})\psi,
\end{equation}
where $\kappa=K(n+1)(Nm)^{1/n}$. This is the usual form of the GP
equation considered in the literature (Sulem \& Sulem 1999). The
standard BEC, with a quartic interaction $p\propto \rho^2\propto
|\psi|^4$, corresponds to $n=1$. We may also recall that classical
and ultra-relativistic fermion stars are equivalent to polytropes with
index $n=3/2$ and $n=3$ (Chandrasekhar 1939). The GPP system
(\ref{gpp11})-(\ref{gpp2}) with $n=3/2$ has been studied by Bilic {et
al.} (2001) in relation to the formation of white dwarf stars by
gravitational collapse. They showed that the quantum pressure can
regularize the dynamics at small-scales.
For an isothermal equation of state
\begin{equation}
\label{gpp12}
p=\rho \frac{k_B T_{eff}}{m},
\end{equation}
the effective potential is
\begin{equation}
\label{gpp13}
h(\rho)=\frac{k_B T_{eff}}{m}\ln\rho,
\end{equation}
and the GP equation reads
\begin{equation}
\label{gpp14}
i\hbar \frac{\partial\psi}{\partial t}=-\frac{\hbar^2}{2m}\Delta\psi+(m\Phi+{2k_B T_{eff}}\ln|\psi|)\psi.
\end{equation}
Interestingly, we note that a nonlinear Schr\"odinger equation with a logarithmic potential similar to Eq. (\ref{gpp14}) has been introduced long ago by Bialynicki \& Mycielski (1976) as a possible generalization of the Schr\"odinger equation in quantum mechanics.
{\it Remark:} for a BEC at $T=0$, the pressure arising in the Euler equation (\ref{gpp5}) has a meaning different from the kinetic pressure of a normal fluid at finite temperature. It is due to the self-interaction of the particles encapsulated in the effective potential $h(\rho)$ and not to thermal motion (since $T=0$). As a result, the pressure can be {\it negative} (!) contrary to a kinetic pressure. This is the case, in particular, for a BEC described by the equation of state (\ref{gpp8}) when the scattering length $a_s$ is negative. In terrestrial BEC experiments, some atoms like $^7{\rm Li}$ have a negative scattering length (Fedichev et al. 1996). On the other hand, the constant $T_{eff}$ appearing in Eq. (\ref{gpp12}) is just an ``effective'' temperature since it arises from a particular form of self-interaction and has nothing to do with the kinetic temperature (which here is $T=0$). In particular, this effective temperature can be negative. In that respect, we note that linear equations of state $p=\alpha\rho c^2$ with $\alpha<0$ have been introduced heuristically to account for the accelerated expansion of the universe (see the review of Peebles \& Ratra 2003 and Sec. \ref{sec_dark}).
\section{Newtonian cosmology}
\label{sec_nc}
In this first part of the paper, we use the Newtonian cosmology introduced by Milne and McCrea \& Milne in 1934. The great advantage of the Newtonian treatment is its simplicity\footnote{It is surprising to realize that Newtonian cosmology was developed after, and was influenced by, the cosmological models based on Einstein's theory of general relativity. It could have been developed much earlier. As Milne writes: ``It seems to have escaped previous notice that whereas the theory of the expanding universe is generally held to be one of the fruits of the theory of relativity, actually all the phenomena observable at present could have been predicted by the founders of mathematical hydrodynamics in the eighteenth century, or even by Newton himself''.}. Furthermore, the equations are identical with those derived using the theory of general relativity, provided the pressure is negligible in comparison with the energy density $\rho c^2$ where $c$ is the speed of light (we shall go beyond this limitation in the second part of the paper). This makes this approach attractive. Furthermore, its simplicity allows us to introduce novel ingredients such as the quantum pressure which is derived from the classical Gross-Pitaevskii equation. The validity and limitation of Newtonian cosmology have been discussed by various authors such as Layzer (1954), McCrea (1955), Callan et al. (1965) and Harrison (1965).
We recall the basics of Newtonian cosmology. We consider a spatially homogeneous solution of Eqs. (\ref{gpp4})-(\ref{gpp7}) of the form
\begin{equation}
\label{nc1}
\rho({\bf r},t)=\rho_b(t),\qquad {\bf u}({\bf r},t)=\frac{\dot a}{a}{\bf r},
\end{equation}
where $a(t)$ is the scale factor and $H=\dot a/a$ is the Hubble ``constant'' (in fact a function of time).
The Euler equations reduce to
\begin{equation}
\label{nc2}
\frac{d\rho_b}{dt}+3\rho_b\frac{\dot a}{a}=0,
\end{equation}
\begin{equation}
\label{nc3}
\frac{\ddot a}{a}{\bf r}=-\nabla\Phi_b.
\end{equation}
The pressure $p$ and the quantum potential $Q$ do not appear in the theory of the homogeneous model since they enter Eq. (\ref{gpp5}) only through their gradients. Equation (\ref{nc2}) leads to the relation
\begin{equation}
\label{nc4}
\rho_b a^3\sim 1,
\end{equation}
which corresponds to the conservation of mass. Taking the divergence of Eq. (\ref{nc3}) and using the Poisson equation (\ref{gpp6}), we obtain the cosmological equation
\begin{equation}
\label{nc5}
\frac{d^2 a}{dt^2}=-\frac{4}{3}\pi G \rho_b a+\frac{\Lambda}{3}a.
\end{equation}
Using Eq. (\ref{nc4}), its first integral is
\begin{equation}
\label{nc6}
\left (\frac{da}{dt}\right )^2=\frac{1}{3}(8\pi G\rho_b+\Lambda) a^2-\kappa,
\end{equation}
where $\kappa$ is a constant of integration. Equations (\ref{nc4})-(\ref{nc6}) are the Newtonian equations of an isotropic and homogeneous universe. They coincide with the equations derived by Friedmann (1922,1924) for $\Lambda=0$ and $\kappa=\pm 1$ and by Einstein \& de Sitter (1932) for $\Lambda=0$ and $\kappa=0$ from the theory of general relativity when the pressure is small compared to the energy density $\rho c^2$. In that case, $\kappa$ is the curvature constant and space is flat ($\kappa=0$), elliptical ($\kappa=1$) or hyperbolic ($\kappa=-1$)\footnote{Note that the constant $\kappa$ in Eq. (\ref{nc6}) can be set to unity by a suitable normalization of the parameters.}.
The Einstein (1917) static universe corresponds to
\begin{equation}
\label{nc7}
a=1, \qquad \rho_b=\frac{\Lambda}{4\pi G},\qquad \kappa=\Lambda.
\end{equation}
However, this universe is unstable against perturbations in $a$ (Eddington 1930, Harrison 1967). The Einstein-de Sitter (EdS) universe corresponds to $\Lambda=0$ and $\kappa=0$. In that case, Eqs. (\ref{nc5}) and (\ref{nc6}) reduce to
\begin{equation}
\label{nc8}
\ddot a=-\frac{4}{3}\pi G \rho_b a,\qquad {\dot a}^2=\frac{8}{3}\pi G\rho_b a^2.
\end{equation}
This yields
\begin{equation}
\label{nc9}
a\propto t^{2/3},\qquad H=\frac{\dot a}{a}=\frac{2}{3t}, \qquad \rho_b=\frac{1}{6\pi Gt^2}.
\end{equation}
Note that both inflationary theory (Guth 1981) and observations favor a flat universe ($\kappa=0$).
In that case, Eq. (\ref{nc6}) can be written $H^2=\frac{8}{3}\pi G(\rho_b+\rho_{\Lambda})$, where $\rho_{\Lambda}=\Lambda/8\pi G$ is the dark energy density, and it gives a relationship between Hubble's constant and the total density of the universe. With the present-day value of the Hubble
constant $H_0=2.273\, 10^{-18}\, {\rm s}^{-1}$, the Einstein-de Sitter
model ($\kappa=\Lambda=0$, $p\ll \rho c^2$) leads to an age of the universe $t_0=2/3H_0\sim
9.3$ billion years while the $\Lambda$CDM model, which is the standard
model of cosmology, predicts $13.75$ billion years. The effect of dark energy will be
considered in Secs. \ref{sec_contribution} and \ref{sec_dark}.
{\it Remark:} We can obtain equations (\ref{nc4})-(\ref{nc6}) directly if we use the ``naive" picture that the universe is a uniform sphere of radius $a(t)$, density $\rho_b(t)$ and mass $M$. The conservation of mass $M=\frac{4}{3}\pi \rho_b a^3$ leads to Eq. (\ref{nc4}). On the other hand, applying the Gauss theorem at the border of the sphere, Newton's equation can be written ${d^2 a}/{dt^2}=-{GM}/{a^2}+{\Lambda}a/{3}$ leading to Eq. (\ref{nc5}). This is the equation of motion of a fictive particle of position $a$ in a potential $V(a)=-GM/a-\Lambda a^2/6$. Its first integral is given by Eq. (\ref{nc6}) where $E=-\kappa/2$ can be regarded as the energy of the fictive particle. The case $E=0$ (EdS universe) corresponds to the situation where the velocity of the fictive particle is equal to the escape velocity. In this naive picture, $a$ may be interpreted as the ``radius'' of the universe. In fact, a better derivation (which does not assume a finite sphere) is to apply Newton's equation to an arbitrary fluid particle located in ${\bf r}$ and write ${d^2 r}/{dt^2}=-{G (4\pi\rho_b r^3/3)}/{r^2}+{\Lambda}r/{3}$. Setting ${\bf r}=a(t){\bf x}$ and dividing by $x$ yields Eq. (\ref{nc5}).
\section{Quantum Euler-Poisson system in an expanding universe}
\label{sec_qep}
We shall now study the instability of the homogeneous background and the growth of perturbations that ultimately lead to the large-scale structures of the universe. We shall work in the comoving frame (Peebles 1980). To that purpose, we set
\begin{equation}
\label{qep1}
{\bf r}=a(t){\bf x}, \qquad {\bf u}=\frac{\dot a}{a}{\bf r}+{\bf v},
\end{equation}
where ${\bf v}$ is the peculiar velocity. For the moment, we allow arbitrary deviations from the background flow. Let us first write the Poisson equation (\ref{gpp6}) in the comoving frame. Integrating Eq. (\ref{nc3}) and using Eq. (\ref{nc5}), we find that the background gravitational potential is
\begin{equation}
\label{qep2}
\Phi_b=-\frac{1}{2}\frac{\ddot a}{a} r^2=-\frac{1}{2}{\ddot a} a x^2=\frac{2}{3}\pi G\rho_b(t)r^2-\frac{\Lambda}{6}r^2.
\end{equation}
This result can also be obtained by integrating the Poisson equation for a homogeneous system or by using the Gauss theorem. If we introduce the new potential $\phi=\Phi-\Phi_b$, i.e.
\begin{equation}
\label{qep3}
\phi=\Phi+\frac{1}{2}{\ddot a} a x^2,
\end{equation}
the Poisson equation (\ref{gpp6}) becomes
\begin{equation}
\label{qep4}
\Delta\phi=4\pi G a^2 (\rho-\rho_b),
\end{equation}
where the Laplacian is taken with respect to ${\bf x}$. Now, following Peebles (1980) and taking the quantum pressure into account, we find that the hydrodynamic equations (\ref{gpp4}) and (\ref{gpp5}) can be written in the comoving frame as
\begin{equation}
\label{qep5}
\frac{\partial\rho}{\partial t}+\frac{3\dot a}{a}\rho+\frac{1}{a}\nabla\cdot (\rho {\bf v})=0,
\end{equation}
\begin{eqnarray}
\label{qep6}
\frac{\partial {\bf v}}{\partial t}+\frac{1}{a}({\bf v}\cdot \nabla){\bf v}+\frac{\dot a}{a}{\bf v}=-\frac{1}{\rho a}\nabla p-\frac{1}{a}\nabla\phi\nonumber\\
+\frac{\hbar^2}{2m^2a^3}\nabla \left (\frac{\Delta \sqrt{\rho}}{\sqrt{\rho}}\right ).
\end{eqnarray}
It is convenient to write the density in the form
\begin{eqnarray}
\label{qep7}
\rho=\rho_b(t)\left\lbrack 1+\delta({\bf x},t)\right \rbrack,
\end{eqnarray}
where $\rho_b\propto 1/a^3$ and $\delta({\bf x},t)$ is the density contrast (Peebles 1980). Substituting Eq. (\ref{qep7}) into Eqs. (\ref{qep4})-(\ref{qep6}), we obtain the quantum barotropic Euler-Poisson system in an expanding universe
\begin{equation}
\label{qep8}
\frac{\partial\delta}{\partial t}+\frac{1}{a}\nabla\cdot ((1+\delta) {\bf v})=0,
\end{equation}
\begin{eqnarray}
\label{qep9}
\frac{\partial {\bf v}}{\partial t}+\frac{1}{a}({\bf v}\cdot \nabla){\bf v}+\frac{\dot a}{a}{\bf v}=-\frac{1}{\rho a}\nabla p-\frac{1}{a}\nabla\phi\nonumber\\
+\frac{\hbar^2}{2m^2a^3}\nabla \left (\frac{\Delta \sqrt{1+\delta}}{\sqrt{1+\delta}}\right ),
\end{eqnarray}
\begin{equation}
\label{qep10}
\Delta\phi=4\pi G \rho_b a^2 \delta.
\end{equation}
For $\hbar=p=0$, we recover the usual Euler equations of a cold gas ($T=0$) in an expanding universe (Peebles 1980)\footnote{If the universe is a classical collisionless fluid, the evolution of this fluid is fundamentally described by the Vlasov equation (Gilbert 1966). In that case, the hydrodynamic equations without pressure ($\hbar=p=0$) are based on some approximations (Peebles 1980) and they are not valid for all times. Alternatively, if dark matter is a BEC, the hydrodynamic equations (\ref{qep8})-(\ref{qep10}) are rigorously equivalent to the GPP system for all times.}. For a barotropic equation of state $p=p(\rho)$, the Euler equation (\ref{qep9}) can be rewritten
\begin{eqnarray}
\label{qep11}
\frac{\partial {\bf v}}{\partial t}+\frac{1}{a}({\bf v}\cdot \nabla){\bf v}+\frac{\dot a}{a}{\bf v}=-\frac{c_s^2}{(1+\delta)a}\nabla \delta-\frac{1}{a}\nabla\phi\nonumber\\
+\frac{\hbar^2}{2m^2a^3}\nabla \left (\frac{\Delta \sqrt{1+\delta}}{\sqrt{1+\delta}}\right ),
\end{eqnarray}
where $c_s^2=p'(\rho)=p'(\rho_b(1+\delta))$ is the square of the velocity of sound in the evolving system. It generically depends on position and time. For an isothermal equation of state $c_s^2=k_BT/m$ is a constant and for a polytropic equation of state $c_s^2=K\gamma\rho^{\gamma-1}$. For a standard BEC with a quartic self-interaction, described by the equation of state (\ref{gpp8}), the square of the velocity of sound is
\begin{eqnarray}
\label{qep11b}
c_s^2=\frac{4\pi a_s\hbar^2\rho}{m^3},
\end{eqnarray}
and we obtain the system of equations
\begin{equation}
\label{qep12}
\frac{\partial\delta}{\partial t}+\frac{1}{a}\nabla\cdot ((1+\delta) {\bf v})=0,
\end{equation}
\begin{eqnarray}
\label{qep13}
\frac{\partial {\bf v}}{\partial t}+\frac{1}{a}({\bf v}\cdot \nabla){\bf v}+\frac{\dot a}{a}{\bf v}=-\frac{4\pi a_s\hbar^2\rho_b}{m^3 a}\nabla \delta\nonumber\\
-\frac{1}{a}\nabla\phi+\frac{\hbar^2}{2m^2a^3}\nabla \left (\frac{\Delta \sqrt{1+\delta}}{\sqrt{1+\delta}}\right ),
\end{eqnarray}
\begin{equation}
\label{qep14}
\Delta\phi=4\pi G \rho_b a^2\delta.
\end{equation}
\section{Linearized equations}
\label{sec_lin}
Of course, $\delta=\phi=0$ and ${\bf v}={\bf 0}$ is a solution of the quantum Euler-Poisson system (\ref{qep8})-(\ref{qep10}) in the comoving frame, corresponding to the pure background flow (\ref{nc1}). However, this solution may be unstable to some perturbations. If we consider small perturbations $\delta\ll 1$, $\phi\ll 1$, $|{\bf v}|\ll 1$ and linearize the foregoing equations we obtain
\begin{equation}
\label{lin1}
\frac{\partial\delta}{\partial t}+\frac{1}{a}\nabla\cdot {\bf v}=0,
\end{equation}
\begin{eqnarray}
\label{lin2}
\frac{\partial {\bf v}}{\partial t}+\frac{\dot a}{a}{\bf v}=-\frac{1}{a}c_s^2\nabla \delta-\frac{1}{a}\nabla\phi
+\frac{\hbar^2}{4m^2a^3}\nabla (\Delta\delta),
\end{eqnarray}
\begin{equation}
\label{lin3}
\Delta\phi=4\pi G \rho_b \delta a^2,
\end{equation}
where $c_s^2=p'(\rho_b(t))$ now denotes the square of the velocity of sound in the homogeneous background flow. It is just a function of time. Taking the time derivative of Eq. (\ref{lin1}) multiplied by $a$, the divergence of Eq. (\ref{lin2}), and using Eq. (\ref{lin3}), these equations can be combined into a single equation governing the evolution of the density contrast
\begin{eqnarray}
\label{lin4}
\frac{\partial^2\delta}{\partial t^2}+2\frac{\dot a}{a}\frac{\partial\delta}{\partial t}=\frac{c_s^2}{a^2}\Delta\delta+4\pi G\rho_b \delta-\frac{\hbar^2}{4m^2a^4}\Delta^2\delta.
\end{eqnarray}
For $\hbar=0$, we recover the equation first derived by Bonnor (1957). In his case, the pressure $p$ is a kinetic pressure. Expanding the solution in Fourier modes of the form $\delta({\bf x},t)=\delta_{\bf k}(t)e^{i{\bf k}\cdot {\bf x}}$, we obtain
\begin{eqnarray}
\label{lin5}
\ddot\delta+2\frac{\dot a}{a}\dot\delta+\left (\frac{\hbar^2k^4}{4m^2a^4}+\frac{c_s^2k^2}{a^2}-4\pi G\rho_b\right )\delta=0,
\end{eqnarray}
where, for brevity, we have noted $\delta(t)$ for $\delta_{\bf k}(t)$. For a standard BEC, using $c_s^2={4\pi a_s \hbar^2\rho_b}/{m^3}$, the foregoing equation can be rewritten
\begin{eqnarray}
\label{lin6}
\ddot\delta+2\frac{\dot a}{a}\dot\delta+\left (\frac{\hbar^2k^4}{4m^2a^4}+\frac{4\pi a_s \hbar^2 \rho_b k^2}{m^3a^2}-4\pi G\rho_b\right )\delta=0.\nonumber\\
\end{eqnarray}
\section{Quantum Jeans length}
\label{sec_qjl}
\subsection{Expanding universe}
\label{sec_e}
In the non-interacting case $a_s=c_s=0$, the equation for the density contrast reduces to
\begin{eqnarray}
\label{e1}
\ddot\delta+2\frac{\dot a}{a}\dot\delta+\left (\frac{\hbar^2k^4}{4m^2a^4}-4\pi G\rho_b\right )\delta=0.
\end{eqnarray}
From this relation, we can define a time-dependent quantum Jeans wavenumber
\begin{eqnarray}
\label{e2}
k_Q=\left (\frac{16\pi G\rho_b m^2a^4}{\hbar^2}\right )^{1/4}.
\end{eqnarray}
Note that the proper quantum Jeans wavenumber obtained by writing $\delta\propto e^{i{\bf k}_*\cdot {\bf r}}$ is $k_Q^*=({16\pi G\rho_b m^2}/{\hbar^2})^{1/4}$.
Recalling Eq. (\ref{nc4}), we can write $k_Q=\kappa_Q a^{1/4}$ where $\kappa_Q=({16\pi G\rho_b a^3 m^2}/{\hbar^2})^{1/4}$ is a constant. The quantum Jeans length $\lambda_Q=2\pi /k_Q$ decreases with time like $a^{-1/4}$ so that, in the comoving frame, the system becomes unstable at smaller and smaller scales as the universe expands (note that, on the contrary, the proper quantum Jeans scale increases with time like $a^{3/4}$). In the cold, non-quantum universe, there is no Jeans length: all the scales are unstable. Therefore, quantum effects can stabilize the system at small scales and avoid the density cusps (Hu et al. 2000).
In the TF limit where the quantum potential can be neglected, the equation for the density contrast reduces to
\begin{eqnarray}
\label{e3}
\ddot\delta+2\frac{\dot a}{a}\dot\delta+\left (\frac{c_s^2k^2}{a^2}-4\pi G\rho_b\right )\delta=0.
\end{eqnarray}
From this relation, we can define a time-dependent classical Jeans wavenumber
\begin{eqnarray}
\label{e4}
k_J=\left (\frac{4\pi G\rho_b a^2}{c_s^2}\right )^{1/2}.
\end{eqnarray}
The proper Jeans wavenumber is $k_J^*=({4\pi G\rho_b}/{c_s^2})^{1/2}$. The evolution of $k_J(t)$ depends on the equation of state. For an isothermal equation of state for which $c_s^2=k_BT/m$, recalling Eq. (\ref{nc4}), we can write $k_J=\kappa_J a^{-1/2}$ where $\kappa_J=(4\pi G\rho_b a^3 m/k_BT)^{1/2}$ is a constant. The classical Jeans length $\lambda_J=2\pi/k_J$ increases with time like $a^{1/2}$ so that, in the comoving frame, the system becomes unstable at larger and larger scales as the universe expands. For a polytropic equation of state for which $c_s^2=K\gamma \rho_b^{\gamma-1}$, we can write $k_J=\kappa_J a^{(3\gamma-4)/2}$ where $\kappa_J=\lbrack 4\pi G(\rho_b a^3)^{2-\gamma}/K\gamma\rbrack^{1/2}$ is a constant. The classical Jeans wavelength behaves like $\lambda_J\propto a^{(4-3\gamma)/2}$. For $\gamma<4/3$, the Jeans wavelength grows with time and for $\gamma>4/3$, it decreases with time. For $\gamma=4/3$, the Jeans wavenumber is constant in time, i.e. $k_J=\kappa_J$. Finally, for a standard BEC, $c_s^2={4\pi a_s \hbar^2\rho_b}/{m^3}$, and the Jeans wavenumber can be written
\begin{eqnarray}
\label{e5}
k_J=\left (\frac{Gm^3a^2}{a_s\hbar^2}\right )^{1/2}.
\end{eqnarray}
It is independent on the density and can be written $k_J=\kappa_J a$ with $\kappa_J=(Gm^3/a_s\hbar^2)^{1/2}$. The Jeans length decreases like $\lambda_J\propto 1/a$.
{\it Remark:} the study of the Jeans instability in an expanding universe exhibits a critical value of the polytropic index $\gamma_{crit}=4/3$ (i.e. $n_{crit}=3$). It is interesting to note that the same critical index arises when one studies the dynamical stability of spatially inhomogeneous polytropic spheres with respect to the barotropic Euler-Poisson system. It has been established that a polytropic star is stable for $\gamma\ge 4/3$ and unstable otherwise (Binney \& Tremaine, 1987). Surprisingly, the same index appears in the stability analysis of a spatially homogeneous polytropic gas in an expanding universe. Note that the static study of Jeans (see the following section) does not directly exhibit a critical value of the polytropic index\footnote{The critical value $\gamma_{crit}=4/3$ only appears when one introduces the Jeans mass $M_J\sim \rho \lambda_J^3\propto \rho^{(3\gamma-4)/2}$, but it does not play any particular role in the stability analysis of the homogeneous fluid.}.
\subsection{Static universe}
\label{sec_s}
If we assume that Eq. (\ref{lin5}) remains valid in a static universe, and take $a=1$, we get
\begin{eqnarray}
\label{s1}
\ddot\delta+\left (\frac{\hbar^2k^4}{4m^2}+{c_s^2k^2}-4\pi G\rho_b\right )\delta=0.
\end{eqnarray}
Writing the perturbation in the form $\delta(t)\propto e^{-i\omega t}$, we obtain the dispersion relation
\begin{eqnarray}
\label{s2}
\omega^2=\frac{\hbar^2k^4}{4m^2}+{c_s^2k^2}-4\pi G\rho_b.
\end{eqnarray}
This dispersion relation has been studied in our previous paper (Chavanis 2011). The generalized Jeans wavenumber is given by
\begin{equation}
\label{s3}
k_c^2=\frac{2m^2}{\hbar^2}\left\lbrack \sqrt{c_s^4+\frac{4\pi G\hbar^2\rho_b}{m^2}}-c_s^2\right\rbrack.
\end{equation}
In the non-interacting case ($a_s=c_s=0)$, we recover the quantum Jeans wavenumber (\ref{e2}) with $a=1$ and in the TF limit, we recover the classical Jeans wavenumber (\ref{e4}) with $a=1$. The dispersion relation can be written $\omega^2/4\pi G\rho_b={k^4}/{k_Q^4}+{k^2}/{k_J^2}-1$ and the generalized Jeans wavenumber $k_c^2=({k_Q^4}/{2k_J^2})\lbrack \pm (1+{4k_J^4}/{k_Q^4})^{1/2}-1\rbrack$ with $+$ when $c_s^2\ge 0$ and $-$ when $c_s^2<0$. For $\lambda<\lambda_c$, $\omega$ is real and the perturbation oscillates with a pulsation $\omega$; for $\lambda>\lambda_c$, $\omega$ is purely imaginary and the perturbation grows exponentially rapidly with a growth rate $\gamma=\sqrt{-\omega^2}$. Jeans' stability criterion is therefore $\lambda<\lambda_c$. For $c_s^2>0$, the maximum growth rate corresponds to $k=0$ (infinite wavelengths) and is given by $\gamma=\sqrt{4\pi G\rho_b}$. On the other hand, considering a BEC with negative scattering length, it is found (Chavanis 2011) that the maximum growth rate corresponds to $k_*=(8\pi |a_s|\rho_b/m)^{1/2}$ and is given by
\begin{eqnarray}
\label{s4}
\gamma_*=\sqrt{\frac{16\pi^2a_s^2\hbar^2\rho_b^2}{m^4}+4\pi G\rho_b}.
\end{eqnarray}
Note that $k_*=({k_Q^4}/{2|k_J^2|})^{1/2}$ and $\gamma_*=\sqrt{4\pi G\rho} (1+{k_Q^4}/{4k_J^4})^{1/2}$. For $|a_s|\gg (Gm^4/4\pi\hbar^2\rho_b)^{1/2}$, we find that $\gamma_*\sim 4\pi|a_s|\hbar\rho_b/m^2\gg \sqrt{4\pi G\rho_b}$. Therefore, an attractive short-range interaction ($a_s<0$) increases the growth rate of the Jeans instability.
Jeans' classical analysis (Jeans 1902, 1929) suffers from the defect that in general there is no initial stationary state that is in a uniform non-rotating fluid. Therefore, using Eq. (\ref{s1}) in a static universe is referred to as the ``Jeans swindle''\footnote{Kiessling (2003) provides a vindication of the ``Jeans swindle''. He argues that, when considering an infinite and homogeneous distribution of matter, the Poisson equation must be modified so as to correctly define the gravitational force. He proposes to use a regularization of the form $\Delta\Phi-k_0^2\Phi=4\pi G\rho$ where $k_0$ is an inverse screening length that ultimately tends to zero ($k_0\rightarrow 0$), or a regularization of the form $\Delta\Phi=4\pi G(\rho-\overline{\rho})$ where $\overline{\rho}$ is the mean density. In his point of view, this is not a swindle but just the right way to make the problem mathematically rigorous and have a well-defined gravitational force.}. As noted by Bonnor (1957), the Jeans procedure is only valid in the static Einstein universe\footnote{As is well-known, Einstein (1917) introduced a cosmological constant in the equations of general relativity in order to recover a homogeneous static universe. He also considered, as a preamble of his paper, a modification of the classical Poisson equation in the form $\Delta\Phi-k_0^2\Phi=4\pi G\rho$ because he (incorrectly) believed that, in the Newtonian world model, the cosmological constant is equivalent to a screening length (see Spiegel 1998, Kiessling 2003 and Chavanis \& Delfini 2010 for historical details).}. However, in this last case, the background density of the universe is $\rho_b=\Lambda/4\pi G$. The justification does not apply to a uniform mass of gas of different density. Furthermore, the Einstein universe is strongly unstable. Therefore, this justification is not valid and it is necessary to develop the Jeans instability analysis in an expanding universe.
{\it Remark:} Jeans' analysis could be valid if the growth rate of the instability were much larger than the rate of the expansion of the universe. According to Eq. (\ref{nc8}), the rate of expansion of the universe is $H=\dot a/a=(8\pi G\rho/3)^{1/2}$. On the other hand, when $c_s^2>0$, the maximum growth rate of the Jeans instability is $\gamma=\sqrt{4\pi G\rho}$. They are exactly of the same order which leads to the usual conclusion that Jeans' instability analysis is not valid (see, e.g., Weinberg 1972). However, when $c_s^2<0$, the maximum growth rate, given by Eq. (\ref{s4}) can be much larger. In that case, the expansion of the universe could be ignored and the static Jeans instability analysis could be valid.
\section{Solution in the Einstein-de Sitter universe}
\label{sec_eds}
The evolution of the perturbations is more complicated to analyze in an expanding universe than in a static universe (and turns out to be very different). For simplicity, we consider the case of an Einstein-de Sitter universe. In that case, it is possible to obtain analytical solutions of the linearized equation (\ref{lin5}) for the density contrast in some particular cases. Measuring the evolution in terms of $a$ instead of $t$ and using Eq. (\ref{nc8}), Eq. (\ref{lin5}) is transformed into
\begin{eqnarray}
\label{eds1}
\frac{d^2\delta}{da^2}+\frac{3}{2a}\frac{d\delta}{da}+\frac{3}{2a^2}\left (\frac{\hbar^2k^4}{16\pi G\rho_b m^2a^4}+\frac{c_s^2k^2}{4\pi G\rho_b a^2}-1\right )\delta=0.\nonumber\\
\end{eqnarray}
For $\hbar=c_s=0$, we recover the case of a {\it cold classical universe} in which the pressure is zero and quantum effects are neglected. Equation (\ref{eds1}) reduces to
\begin{eqnarray}
\label{eds2}
\frac{d^2\delta}{da^2}+\frac{3}{2a}\frac{d\delta}{da}-\frac{3}{2a^2}\delta=0,
\end{eqnarray}
and its solutions are
\begin{eqnarray}
\label{eds3}
\delta_+\propto a, \qquad \delta_-\propto a^{-3/2}.
\end{eqnarray}
We note that the growth of the perturbations is algebraic in an expanding universe while the ``naive'' Jeans instability analysis in a static universe predicts an exponential growth. Therefore, the results of the two analysis are very different. As mentioned long ago by Bonnor (1957) and others, this (slow) algebraic growth may be a problem to form large-scale structures sufficiently rapidly.
Assuming a polytropic equation of state $p=K\rho^{\gamma}$ for which $c_s^2=K\gamma\rho_b^{\gamma-1}$ and introducing the variables of Sec. \ref{sec_e}, we can rewrite Eq. (\ref{eds1}) in the form
\begin{eqnarray}
\label{eds4}
\frac{d^2\delta}{da^2}+\frac{3}{2a}\frac{d\delta}{da}+\frac{3}{2a^2}\left (\frac{k^4}{\kappa_Q^4 a}+\frac{k^2}{\kappa_J^2a^{3\gamma-4}}-1\right )\delta=0.\quad
\end{eqnarray}
Making the change of variables $\delta(a)=f(a)/a$, we find that the differential equation for $f$ is
\begin{eqnarray}
\label{eds5}
\frac{d^2 f}{da^2}-\frac{1}{2a}\frac{df}{da}+\frac{1}{a^2}\left (\frac{3k^4}{2\kappa_Q^4 a}+\frac{3k^2}{2\kappa_J^2 a^{3\gamma-4}}-1\right )f=0.\quad
\end{eqnarray}
We have not been able to find the general solution of this equation in
terms of simple functions so that we shall consider particular
cases. Some of the solutions presented below have already been
discussed in the literature (see, e.g., Harrison 1967) but we shall
give more detail and put emphasis on the
solutions describing self-gravitating BECs that were not discussed
previously.
\subsection{The TF approximation}
\label{sec_tf}
In the TF approximation where the quantum potential can be neglected, Eq. (\ref{eds5}) reduces to
\begin{eqnarray}
\label{tf1}
\frac{d^2 f}{da^2}-\frac{1}{2a}\frac{df}{da}+\frac{1}{a^2}\left (\frac{3k^2}{2\kappa_J^2a^{3\gamma-4}}-1\right )f=0.
\end{eqnarray}
This is a particular case of the equation studied by Savedoff \& Vila (1962) in relation to the work of Bonnor (1957)\footnote{In this section, we allow $p$ to have the interpretation of an ordinary pressure, i.e. we consider not only a gas of BECs but also a more conventional gas with a kinetic pressure.}. It can be solved in terms of Bessel functions (see Appendix \ref{sec_sol}). For $\gamma\neq 4/3$, $f(a)$ is given by Eq. (\ref{sol7}) and the evolution of the density contrast is
\begin{eqnarray}
\label{tf2}
\delta(a)\propto \frac{1}{a^{1/4}}J_{\pm \frac{5}{2(4-3\gamma)}}\left (\frac{\sqrt{6}}{4-3\gamma} \frac{k}{\kappa_J} a^{\frac{4-3\gamma}{2}}\right ).
\end{eqnarray}
Using the asymptotic expansions of the Bessel functions, we note that
the density contrast is oscillating for $k\gg k_J(a)$ and growing for
$k\ll k_J(a)$ (on the timescale over which these inequalities are fulfilled).
Therefore, wavelengths which are long with respect to
$\lambda_J(a)$ behave like in a cold universe whereas for the short
wavelengths the perturbations have an oscillatory behavior. In this
sense, the results are similar to those obtained with the classical
Jeans analysis. They are in fact different because (i) the Jeans
length varies with time, (ii) the oscillating solutions can
be growing or decaying and (iii) the growth is algebraic instead
of exponential. In fact, the results more crucially depend on the
value of the polytropic index $\gamma$ than on the Jeans length
$\lambda_J$. We must distinguish four cases. (i) $\gamma<4/3$: for
$a\rightarrow 0$ the perturbations grow like in a cold gas
($\delta_+\sim a$) and, for $a\rightarrow +\infty$, they undergo damped
oscillations (they decay like $1/a^{(5-3\gamma)/4}$); (ii)
$4/3<\gamma<5/3$: for $a\rightarrow 0$ the perturbations diverge like
$1/a^{(5-3\gamma)/4}$ while oscillating implying that we must start
the stability analysis at $a_i>0$. For $a\rightarrow a_i$, the
perturbations undergo damped oscillations (they decay like
$1/a^{(5-3\gamma)/4}$) and, for $a\rightarrow +\infty$, they grow like
in a cold gas ($\delta_+\sim a$); (iii) $\gamma=5/3$: for
$a\rightarrow 0$ the perturbations oscillate and, for $a\rightarrow
+\infty$, they grow like in a cold gas ($\delta_+\sim a$); (iv)
$\gamma>5/3$: for $a\rightarrow 0$ the perturbations undergo growing
oscillations (they grow like $a^{(3\gamma-5)/4}$) and for
$a\rightarrow +\infty$ they grow like in a cold gas ($\delta_+\sim
a$). Considering the regime $a\ll 1$, the perturbations grow
for $\gamma<4/3$, decay for $4/3<\gamma<5/3$ and grow for
$\gamma>5/3$. Considering the regime $a\gg 1$, the perturbations
decay for $\gamma<4/3$ and grow for $\gamma>4/3$. For
$\gamma<4/3$, the perturbations start to grow but finally decay. In a
sense, the system is asymptotically stable. However, the growth of the
perturbations in the initial stage can trigger nonlinear effects that
may induce instabilities.
For $\gamma=4/3$ ($n=3$), the density contrast behaves like
\begin{eqnarray}
\label{tf3}
\delta(a)\propto a^{-\frac{1}{4}\pm \sqrt{\frac{25}{16}-\frac{3k^2}{2\kappa_J^2}}}.
\end{eqnarray}
This polytropic index corresponds to a gas of photons at temperature $T$ or to a gas of relativistic fermions at zero temperature. In that case, the Jeans length is independent on time: $k_J=\kappa_J$. For $k>(25/24)^{1/2}k_J$ the density contrast behaves like
\begin{eqnarray}
\label{tf4}
\delta(a)\propto \frac{1}{a^{1/4}}\cos\left (\sqrt{\frac{3k^2}{2\kappa_J^2}-\frac{25}{16}}\ln a+\phi\right ).
\end{eqnarray}
It diverges for $a\rightarrow 0$ implying that we must start the stability analysis at $a_i>0$. For $a\ge a_i$, the perturbations decay like ${a^{-1/4}}$ while making oscillations. For $k_J<k<(25/24)^{1/2}k_J$, the perturbations decay algebraically without oscillating. For $k<k_J$, the perturbations grow algebraically (more precisely $\delta_+$ grows and $\delta_-$ decays). For $k\rightarrow 0$, the perturbations behave like in a cold gas ($\delta_+$ grows like $a$ and $\delta_-$ decays like $a^{-2/3}$). This situation is relatively close to the classical Jeans stability analysis but (i) the growth of the perturbation is algebraic instead of exponential; (ii) the oscillatory solutions decay with time; (iii) there is a small interval $k_J<k<(25/24)^{1/2}k_J$ that has no counterpart in the static Jeans analysis.
Let us consider particular equations of state with a physical meaning. The index $\gamma=1$ corresponds to an {\it isothermal universe}. The density contrast is given by
\begin{eqnarray}
\label{tf5}
\delta(a)\propto \frac{1}{a^{1/4}}J_{\pm \frac{5}{2}}\left ({\sqrt{6}} \frac{k}{\kappa_J} a^{{1}/{2}}\right ).
\end{eqnarray}
For the index $\gamma=5/3$ ($n=3/2$), we obtain
\begin{eqnarray}
\label{tf6}
\delta(a)\propto \frac{1}{a^{1/4}}J_{\pm \frac{5}{2}}\left (\sqrt{6} \frac{k}{\kappa_J} \frac{1}{a^{{1}/{2}}}\right ).
\end{eqnarray}
This polytropic index corresponds to a gas of non-relativistic fermions at zero temperature. For the polytropic index $\gamma=2$ ($n=1$), we have
\begin{eqnarray}
\label{tf7}
\delta(a)\propto \frac{1}{a^{1/4}}J_{\pm \frac{5}{4}}\left (\sqrt{\frac{3}{2}} \frac{k}{\kappa_J} \frac{1}{a}\right ).
\end{eqnarray}
This polytropic index corresponds to a BEC with quartic self-interaction in the TF approximation (see Sec. \ref{sec_gpp}). In that case, $\kappa_J=(Gm^3/a_s\hbar^2)^{1/2}$. The previous expression assumes that $a_s>0$. In the case where $a_s<0$, we obtain
\begin{eqnarray}
\label{tf8}
\delta(a)\propto \frac{1}{a^{1/4}}I_{\pm \frac{5}{4}}\left (\sqrt{\frac{3}{2}}\frac{k}{\kappa_J} \frac{1}{a}\right ).
\end{eqnarray}
with $\kappa_J=(Gm^3/|a_s|\hbar^2)^{1/2}$. For $a\rightarrow +\infty$, the perturbations behave like in a cold gas. For $a\rightarrow 0$, we find that
\begin{eqnarray}
\label{tf9}
\delta(a)\propto a^{1/4}e^{\sqrt{\frac{3}{2}}\frac{k}{k_J}\frac{1}{a}}\rightarrow +\infty.
\end{eqnarray}
In a sense, this exponential divergence demonstrates that an attractive self-interaction accelerates the growth of the perturbations. In that case, we can neglect the expansion of the universe and we are led back to the static study of Sec. \ref{sec_s}.
Some curves representing the evolution of the density contrast $\delta(a)$ in the different cases listed above are represented in Figs. \ref{gamma1}-\ref{gamma2} for illustration.
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{gamma1.eps}
\caption{Evolution of the perturbation $\delta(a)$ in an isothermal universe ($\gamma=1$). This corresponds to the case $\gamma<4/3$. We have taken $k=\kappa_J$.}
\label{gamma1}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{gamma4sur3.eps}
\caption{Evolution of the perturbation $\delta(a)$ in a photonic or in a relativistic fermionic universe ($\gamma=4/3$). We have represented the solution $\delta_+$.}
\label{gamma4sur3}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{gamma1.55.eps}
\caption{Evolution of the perturbation $\delta(a)$ in a polytropic universe with $\gamma=1.55$. This corresponds to the case $4/3<\gamma<5/3$. We have taken $k=\kappa_J$.}
\label{gamma1.55}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{gamma5sur3.eps}
\caption{Evolution of the perturbation $\delta(a)$ in a non-relativistic fermionic universe or in a BEC universe without self-interaction ($\gamma=5/3$). We have taken $k=\kappa_J$.}
\label{gamma5sur3}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{gamma2.eps}
\caption{Evolution of the perturbation $\delta(a)$ in a BEC universe with quartic self-interaction in the TF limit ($\gamma=2$). This corresponds to the case $\gamma>5/3$. We have taken $k=\kappa_J$. }
\label{gamma2}
\end{center}
\end{figure}
\subsection{The non-interacting BEC case}
\label{sec_ni}
In the non-interacting case $a_s=c_s=0$, which corresponds to a {\it BEC universe} without self-interaction, Eq. (\ref{eds5}) reduces to
\begin{eqnarray}
\label{ni1}
\frac{d^2 f}{da^2}-\frac{1}{2a}\frac{df}{da}+\frac{1}{a^2}\left (\frac{3k^4}{2\kappa_Q^4 a}-1\right )f=0.
\end{eqnarray}
Comparing this equation with Eq. (\ref{tf1}), we see that the dependance in $a$ is the same as for a polytrope $\gamma=5/3$. Therefore, $f(a)$ is given by Eq. (\ref{sol7}) with $\lambda={3k^4}/{2\kappa_Q^4}$ and $\alpha=-1/2$ yielding
\begin{eqnarray}
\label{ni2}
\delta(a)\propto \frac{1}{a^{1/4}}J_{\pm \frac{5}{2}}\left (\sqrt{6} \frac{k^2}{\kappa_Q^2} \frac{1}{a^{{1}/{2}}}\right ).
\end{eqnarray}
For $a\rightarrow 0$ the perturbations oscillate and for $a\rightarrow +\infty$ they behave like in a cold gas ($\delta_+$ grows like $a$). In a sense, a cold bosonic universe behaves similarly to a cold non-relativistic fermionic universe (compare Eqs. (\ref{ni2}) and (\ref{tf6})). However, the dependence in $k$ is different ($k^2$ instead of $k$) and the expression of the Jeans length also is different (compare Eqs. (\ref{e2}) and (\ref{e4})).
\subsection{The non-relativistic fermionic case $\gamma=5/3$}
\label{sec_fermi}
A non-relativistic gas of fermions at $T=0$ is usually described by a polytropic equation of state $p=K\rho^{\gamma}$ with index $\gamma=5/3$ ($n=3/2$) and polytropic constant $K=(1/5)(3/8\pi)^{2/3}h^2/m^{8/3}$ (Chandrasekhar 1939). This equation of state arises from the Pauli exclusion principle. Such a semi-classical description is valid in the TF approximation which is exact when $N\rightarrow +\infty$. In some cases, it can be relevant to go beyond the TF approximation and take into account the quantum pressure arising from the Heisenberg principle. This can regularize the dynamics at small scales as shown by Bilic et al. (2001) in their study of the formation of fermion stars by gravitational collapse. In our problem, this general situation is described by Eq. (\ref{eds4}) with $\gamma=5/3$. This yields
\begin{eqnarray}
\label{fermi1}
\frac{d^2\delta}{da^2}+\frac{3}{2a}\frac{d\delta}{da}+\frac{3}{2a^2}\left (\frac{k^4}{\kappa_Q^4 a}+\frac{k^2}{\kappa_J^2a}-1\right )\delta=0.
\end{eqnarray}
The equation for $f(a)$ is
\begin{eqnarray}
\label{fermi2}
\frac{d^2 f}{da^2}-\frac{1}{2a}\frac{df}{da}+\frac{1}{a^2}\left (\frac{3k^4}{2\kappa_Q^4a}+\frac{3k^2}{2\kappa_J^2 a}-1\right )f=0.
\end{eqnarray}
As noted in Sec. \ref{sec_ni}, the two terms have the same dependence on $a$. Therefore, $f(a)$ is given by Eq. (\ref{sol7}) with
\begin{eqnarray}
\label{fermi3}
\lambda=\frac{3k^4}{2\kappa_Q^4}+\frac{3k^2}{2\kappa_J^2},\qquad \alpha=-\frac{1}{2}.
\end{eqnarray}
We obtain
\begin{eqnarray}
\label{fermi4}
\delta(a)\propto \frac{1}{a^{1/4}}J_{\pm \frac{5}{2}}\left (\sqrt{\frac{6k^4}{\kappa_Q^4}+\frac{6k^2}{\kappa_J^2}} \frac{1}{a^{{1}/{2}}}\right ).
\end{eqnarray}
For $a\rightarrow 0$ the perturbations oscillate and, for $a\rightarrow +\infty$, they behave like in a cold gas.
\subsection{The relativistic fermionic case $\gamma=4/3$}
\label{sec_fermirel}
A relativistic gas of fermions at $T=0$ is usually described by a polytropic equation of state $p=K\rho^{\gamma}$ with index $\gamma=4/3$ ($n=3$) and polytropic constant $K=(1/4)(3/8\pi)^{1/3}hc/m^{4/3}$ (Chandrasekhar 1939). If we go beyond the TF approximation and take the quantum pressure into account, the evolution of the density contrast is described by Eq. (\ref{eds4}) with $\gamma=4/3$. This yields
\begin{eqnarray}
\label{frel1}
\frac{d^2\delta}{da^2}+\frac{3}{2a}\frac{d\delta}{da}+\frac{3}{2a^2}\left (\frac{k^4}{\kappa_Q^4 a}+\frac{k^2}{\kappa_J^2}-1\right )\delta=0.\quad
\end{eqnarray}
The equation for $f(a)$ is
\begin{eqnarray}
\label{frel2}
\frac{d^2 f}{da^2}-\frac{1}{2a}\frac{df}{da}+\frac{1}{a^2}\left (\frac{3k^4}{2\kappa_Q^4 a}+\frac{3k^2}{2\kappa_J^2}-1\right )f=0.\quad
\end{eqnarray}
Its solution is given by Eq. (\ref{sol13}) and the evolution of the density contrast is
\begin{eqnarray}
\label{frel3}
\delta(a)\propto \frac{1}{a^{1/4}}J_{\pm \frac{1}{2}\sqrt{25-24\frac{k^2}{\kappa_J^2}}}\left (\sqrt{6} \frac{k^2}{\kappa_Q^2} \frac{1}{a^{1/2}}\right ).
\end{eqnarray}
We note that the classical Jeans scale appears in the index of the Bessel function while the quantum Jeans scale appears in the argument of the Bessel function.
\subsection{The general case (asymptotics)}
\label{sec_genasy}
We now treat the general case, taking into account quantum pressure
and polytropic pressure (we assume $a_s>0$). Since we cannot solve
Eq. (\ref{eds4}) analytically, we just give asymptotic results.
Let us first consider the limit $a\rightarrow 0$. (i) If $\gamma<5/3$, the polytropic pressure is negligible in front of the quantum pressure and the perturbations oscillate (see Sec. \ref{sec_ni}). (ii) If $\gamma=5/3$, the polytropic pressure and the quantum pressure are of the same order, and the perturbations oscillate (see Sec. \ref{sec_fermi}). (iii) If $\gamma>5/3$, the quantum pressure is negligible in front of the polytropic pressure and the perturbations undergo growing oscillations $\delta\propto a^{(3\gamma-5)/4}$ (see Sec. \ref{sec_tf}). Therefore, the perturbations oscillate for $\gamma\le 5/3$ and grow for $\gamma>5/3$. The quantum pressure has a stabilizing role for $\gamma\le 4/3$ (see Sec. \ref{sec_tf}).
Let us now consider the limit $a\rightarrow +\infty$. (i) If $\gamma<4/3$, the quantum pressure is negligible in front of the polytropic pressure and the perturbations undergo damped oscillations and decay like $1/a^{(5-3\gamma)/4}$ (see Sec. \ref{sec_tf}). (ii) If $\gamma=4/3$, the quantum pressure is negligible in front of the polytropic pressure and we are led back to the critical case of Sec. \ref{sec_tf}: for $k>(25/24)^{1/2}k_J$ the perturbations decay like ${a^{-1/4}}$ while making oscillations, for $k_J<k<(25/24)^{1/2}k_J$ the perturbations decay algebraically without oscillating and for $k<k_J$ the perturbations grow algebraically. (iii) For $\gamma>4/3$, the perturbations behave like in a cold gas and grows like $\delta_+\sim a\rightarrow +\infty$. Therefore, the perturbations decay for $\gamma<4/3$ and grow for $\gamma>4/3$.
\section{Bose-Einstein condensate universe}
\label{sec_pressure}
\subsection{Newtonian cosmology with pressure}
\label{sec_basic}
We now take into account pressure effects (coming from special relativity) in the evolution of the cosmic fluid. We base our study on the set of hydrodynamic equations
\begin{equation}
\label{b1}
\frac{\partial\rho}{\partial t}+\nabla\cdot (\rho {\bf u})+\frac{p}{c^2}\nabla\cdot {\bf u}=0,
\end{equation}
\begin{equation}
\label{b2}
\frac{\partial {\bf u}}{\partial t}+({\bf u}\cdot \nabla){\bf u}=-\frac{\nabla p}{\rho+\frac{p}{c^2}}-\nabla\Phi,
\end{equation}
\begin{equation}
\label{b3}
\Delta\Phi=4\pi G \left (\rho+\frac{3p}{c^2}\right )-\Lambda,
\end{equation}
where $c$ is the velocity of light. For $c\rightarrow +\infty$, we recover the classical Euler-Poisson system (Binney \& Tremaine 1987). These equations were introduced by McCrea (1951). His initial continuity equation, which contained an incorrect pressure gradient term, was corrected by Lima {et al.} (1997) and we took this correction into account in writing Eq. (\ref{b1}). These equations lead to the correct relativistic equations for the cosmic evolution. Indeed, assuming a homogeneous and isotropic solution of the form $\rho({\bf r},t)=\rho_b(t)$, $p({\bf r},t)=p_b(t)$ and ${\bf u}({\bf r},t)=(\dot a/a){\bf r}$, Eqs. (\ref{b1})-(\ref{b3}) reduce to
\begin{equation}
\label{b5}
\frac{d\rho_b}{dt}+3\frac{\dot a}{a}\left (\rho_b+\frac{p_b}{c^2}\right )=0,
\end{equation}
\begin{equation}
\label{b6}
\frac{\ddot a}{a}=-\frac{4\pi G}{3} \left (\rho_b+\frac{3p_b}{c^2}\right )+\frac{\Lambda}{3},
\end{equation}
\begin{equation}
\label{b7}
\left (\frac{\dot a}{a}\right )^2=\frac{8\pi G}{3}\rho_b-\frac{\kappa}{a^2}+\frac{\Lambda}{3}.
\end{equation}
These are precisely the Friedmann equations that can be derived from the theory of general relativity (Weinberg 1972). The perturbation theory based on Eqs. (\ref{b1})-(\ref{b3}) has been discussed by Reis (2003) who showed that, depending on the equation of state, the results may agree or disagree with the general relativistic approach. We shall leave this problem open since it is not our purpose here to investigate the validity of these equations in detail. Furthermore, our main results will concern the evolution of the cosmic fluid governed by the Friedmann equations. Since the Friedmann equations can be derived from the theory of general relativity, the validity of these equations is not questioned. Note that equations different from Eqs. (\ref{b1})-(\ref{b3}) have been considered by Pace {et al.} (2010) and Harko (2011). Their equations involve an additional term $\dot p {\bf u}/c^2$ in the Euler equation (\ref{b2}). However, this term apparently leads to equations for the cosmic evolution that are different from the Friedmann equations, so that this term will not be considered here.
\subsection{The cosmic evolution of a BEC universe}
\label{sec_cosmic}
We shall assume that the dark matter is a BEC with quartic self-interaction\footnote{In this part of the paper, we neglect the quantum pressure.} described by the barotropic equation of state (\ref{gpp8}). For convenience, we write this equation in the form
\begin{equation}
\label{b4}
p=kc^2\rho^2,
\end{equation}
where $k=2\pi a_s\hbar^2/m^3c^2$ is a constant that can be positive ($a_s>0$) or negative ($a_s<0$). In a first step, we concentrate on the dark matter component and neglect radiation, baryonic matter and dark energy (in particular we take $\Lambda=0$). These additional terms will be considered later. For the equation of state (\ref{b4}), the Friedmann equation (\ref{b5}) becomes
\begin{equation}
\label{cos1}
\frac{d\rho_b}{dt}+3\frac{\dot a}{a}\rho_b (1+k\rho_b)=0.
\end{equation}
This equation can be integrated into
\begin{equation}
\label{cos2}
\rho_b=\frac{A}{a^3-kA},
\end{equation}
where $A$ is a constant. If we make the physical requirement that equation (\ref{cos2}) has solutions for large $a$, then we must impose $A>0$. In that case, $\rho_b\sim A/a^3$ for $a\rightarrow +\infty$ which returns Eq. (\ref{nc4}). When $k>0$, which is the case previously considered by Harko (2011), the density exists only for $a>a_*=(kA)^{1/3}$. For $a\rightarrow a_*$, $\rho_b\rightarrow +\infty$. When $k<0$, the density is defined for all $a$ and has a finite value $\rho_b=1/|k|$ when $a\rightarrow 0$.
Combining Eqs. (\ref{b6}) and (\ref{b4}), we obtain
\begin{equation}
\label{cos3}
\frac{\ddot a}{a}=-\frac{4\pi G\rho_b}{3}(1+3k\rho_b).
\end{equation}
When $k>0$, the universe is always decelerating ($\ddot a<0$). When $k<0$, the universe is accelerating ($\ddot a>0$) for $\rho_b>\rho_c\equiv 1/3|k|$ and decelerating ($\ddot a<0$) for $\rho_b<\rho_c$. Using Eq. (\ref{cos2}), this can be re-expressed in terms of the radius\footnote{We shall sometimes call the scale factor $a$ the ``radius'', having in mind the naive picture of a universe behaving like a homogeneous sphere of radius $a$. This is of course an abuse of language in the context of the theory of general relativity.}: the universe is accelerating for $a<a_c\equiv (2A|k|)^{1/3}$ and decelerating for $a>a_c$.
In order to determine the temporal evolution of $a(t)$, we shall assume $\kappa=0$ (flat space) like in the Einstein-de Sitter universe. Combining Eqs. (\ref{b7}) and (\ref{cos2}), we get
\begin{equation}
\label{cos4}
\dot a=\left (\frac{8\pi GA}{3}\right )^{1/2}\frac{a}{\sqrt{a^3-kA}}.
\end{equation}
We note that for $a\rightarrow +\infty$, this equation reduces to Eq. (\ref{nc8}) with Eq. (\ref{nc4}) and we recover the Einstein-de Sitter solution (\ref{nc9}). It is convenient to define $a_*=(|k|A)^{1/3}$ and introduce $R=a/a_*$. In that case, the density is given by
\begin{equation}
\label{cos7b}
\rho_b=\frac{1}{|k|}\frac{1}{R^3\mp 1},
\end{equation}
and Eq. (\ref{cos4}) can be rewritten
\begin{equation}
\label{cos5}
\dot R=\frac{KR}{\sqrt{R^3\mp 1}},
\end{equation}
where the upper sign $-$ corresponds to $k>0$ and the lower sign $+$ corresponds to $k<0$. We have defined the constant
\begin{equation}
\label{cos6}
K=\left (\frac{8\pi G}{3|k|}\right )^{1/2}=\left (\frac{4Gm^3c^2}{3|a_s|\hbar^2}\right )^{1/2}.
\end{equation}
Note that it depends on the mass $m$ of the bosons and on their scattering length $a_s$. Introducing the dimensionless parameter $\lambda=a_s/\lambda_c=a_smc/\hbar$ (Chavanis 2011) measuring the strength of the short-range interactions ($\lambda_c$ is the Compton wavelength of the bosons), we can rewrite Eq. (\ref{cos6}) in the form
\begin{equation}
\label{cos7}
K=\frac{2}{\sqrt{3}}\left (\frac{m}{M_P}\right )^2\frac{1}{\sqrt{|\lambda|}}t_P^{-1},
\end{equation}
where $M_P=(\hbar c/G)^{1/2}$ is the Planck mass, $l_P=(\hbar G/c^3)^{1/2}$ is the Planck length and $t_P=l_P/c$ is the Planck time.
For $k>0$, which is the case previously considered by Harko (2011), the solution of Eq. (\ref{cos5}) is
\begin{equation}
\label{cos8}
\sqrt{R^3-1}-\arctan \sqrt{R^3-1}=\frac{3}{2}Kt,
\end{equation}
where the constant of integration has been set equal to zero. In this model, the universe starts at a finite time $t=0$ (see Sec. \ref{sec_tl} for a revision of this statement) with a finite radius $R(0)=1$ and an infinite density $\rho_b(0)=\infty$. For $t\rightarrow 0$, $R\simeq 1+(3/4)^{1/3}(Kt)^{2/3}$. The universe is expanding, always decelerating, and asymptotically approaches the Einstein-de Sitter universe $R\sim (3Kt/2)^{2/3}$ (see Fig. \ref{becscalingPLUS}).
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{becscalingPLUS.eps}
\caption{Evolution of the scale factor in a BEC universe with $k>0$. The dashed line corresponds to the Einstein-de Sitter universe ($k=0$) that is reached asymptotically.}
\label{becscalingPLUS}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{becscalingMOINS.eps}
\caption{Evolution of the scale factor in a BEC universe with $k<0$. The dashed line corresponds to the Einstein-de Sitter universe ($k=0$). We note that a BEC universe with $k<0$ converges more rapidly towards the EdS solution than a BEC universe with $k>0$. The bullet (locating the inflexion point) corresponds to the time $t_c$ at which the universe starts to decelerate.}
\label{becscalingMOINS}
\end{center}
\end{figure}
For $k<0$, the solution of Eq. (\ref{cos5}) is
\begin{equation}
\label{cos9}
\sqrt{R^3+1}-\ln \left (\frac{1+\sqrt{R^3+1}}{R^{3/2}}\right )=\frac{3}{2}Kt,
\end{equation}
where the constant of integration has been set equal to zero. In this model, the universe starts from $t\rightarrow -\infty$ with a vanishing radius $R(-\infty)=0$ and a finite density $\rho_b(-\infty)=1/|k|$. For $R<R_c\equiv 2^{1/3}\simeq 1.26$, the universe is accelerating and for $R>R_c\equiv 2^{1/3}$, it is decelerating. For $R\rightarrow +\infty$, we recover the classical Einstein-de Sitter universe $R\sim (3Kt/2)^{2/3}$ (see Fig. \ref{becscalingMOINS}). The change of regime between the phase of acceleration and the phase of deceleration manifests itself by an inflexion point ($\ddot R=0$). The time at which the universe starts decelerating is
\begin{equation}
\label{cos10}
\frac{3}{2}Kt_c= \sqrt{3}-\ln\left (\frac{1+\sqrt{3}}{\sqrt{2}}\right )\simeq 1.074.
\end{equation}
At $t=0$, $R(0)\simeq 0.7601$ and $\rho_b(0)=0.6948/|k|$. For $t\rightarrow -\infty$,
$R\rightarrow 0$ and the asymptotic expansion of
Eq. (\ref{cos9}) yields
\begin{equation}
\label{cos11}
R(t)\propto e^{Kt}.
\end{equation}
For $t\ll t_c$, the universe expands exponentially rapidly, decelerates after $t_c$ and coincides with the Einstein-de Sitter universe for $t\gg t_c$. In the accelerating phase, the rate of the expansion is given by Eq. (\ref{cos6}).
{\it Important remark:} in our mathematical description of the evolution of the cosmic fluid, we have extrapolated the solutions (\ref{cos8}) and (\ref{cos9}) far away in the past. In the model with $k>0$,
the universe emerges at a primordial time $t=0$ from a big-bang singularity where the density is infinite. In the model with $k<0$, the universe has always existed (up to $t\rightarrow -\infty$) and there is no big bang singularity at $t=0$. This is due to the negative pressure allowing for the presence of an inflexion point that changes the concavity of the curve $a(t)$. Of course, this extrapolation to $t\rightarrow -\infty$ is not physically justified since we have ignored important effects like the radiation which dominates in the early universe. We shall see in Sec. \ref{sec_contribution} that the contribution of the radiation strongly alters the results at early times in the case $k<0$. Therefore, our description of the solution (\ref{cos9}) must be considered only on a formal mathematical basis. Still it shows that one can construct cosmological models (solution to the Einstein equations) without primordial singularity.
\subsection{Other representations}
\label{sec_rep}
The previous representation shows that there exists a {\it single} universal curve $a(t)$ in each case $a_s>0$, $a_s=0$ (EdS) and $a_s<0$, provided that the units of time and length are appropriately chosen. However, these units depend on $a_s$. Alternatively, it may be useful to choose units of time and length that are independent on $a_s$ and plot the curve $a(t)$ for different values of $a_s$. This is the representation that has been chosen by Harko (2011) for the case $k>0$ and that we shall discuss and generalize in this section.
In Eq. (\ref{cos2}), the constant $A$ can be determined by the present-day density $\rho_b=\rho_0$ and the corresponding scale factor $a=a_0$. Writing $\rho_0=A/(a_0^3-kA)$, we obtain
\begin{eqnarray}
\label{rep1}
A=\frac{\rho_0 a_0^3}{1+k\rho_0}.
\end{eqnarray}
We note that $\rho_0<1/|k|$ when $k<0$ as we have already indicated. The relation (\ref{cos2}) between the density and the scale factor can be rewritten
\begin{eqnarray}
\label{rep2}
\rho_b=\frac{q}{k}\frac{1}{(\frac{a}{a_0})^3-q},
\end{eqnarray}
where we have introduced the dimensionless parameter
\begin{eqnarray}
\label{rep3}
q=\frac{k\rho_0}{1+k\rho_0}.
\end{eqnarray}
We note that $q$ is positive for $k>0$ and negative for $k<0$. Furthermore, $q=0$ for $k=0$, $q\rightarrow 1$ for $k\rightarrow +\infty$ and $q\rightarrow -\infty$ for $k\rightarrow -1/\rho_0$.
Using these results, the differential equation (\ref{cos4}) giving the evolution of the scale factor with the time can be rewritten
\begin{eqnarray}
\label{rep4}
\dot a=\left (\frac{8\pi G q}{3k}\right )^{1/2}\frac{a}{\sqrt{(\frac{a}{a_0})^3-q}}.
\end{eqnarray}
From Eq. (\ref{b7}), we can define the present-day critical density by $(\rho_c)_0=3H_0^2/8\pi G$ where $H_0$ is the present-day Hubble constant. If $\kappa=0$, which seems to be the case for our universe, $(\rho_c)_0$ represents the total density including baryonic matter, radiation, dark matter and dark energy. It is customary to introduce the present-day density parameter $\Omega_0=\rho_0/(\rho_c)_0$. Then, we can write
\begin{eqnarray}
\label{rep5}
\left (\frac{8\pi G q}{3k}\right )^{1/2}=\frac{H_0\sqrt{\Omega_0}}{\sqrt{1+k\rho_0}}.
\end{eqnarray}
Since the denominator depends on $k$, it must be expressed in terms of $q$ (which is our control parameter). Using Eq. (\ref{rep3}), we obtain
\begin{eqnarray}
\label{rep6}
k\rho_0=\frac{q}{1-q}.
\end{eqnarray}
so that Eq. (\ref{rep4}) can be rewritten
\begin{eqnarray}
\label{rep7}
\dot a=H_0\sqrt{\Omega_0}(1-q)^{1/2}\frac{a}{\sqrt{(\frac{a}{a_0})^3-q}}.
\end{eqnarray}
The density is
\begin{eqnarray}
\label{rep2b}
\rho_b=\rho_0\frac{1-q}{(\frac{a}{a_0})^3-q},
\end{eqnarray}
We can make the connection with the notations of Sec. \ref{sec_cosmic} by setting
\begin{eqnarray}
\label{rep8}
a_*=a_0 |q|^{1/3},\qquad K=H_0\sqrt{\Omega_0}\left (\frac{1-q}{|q|}\right )^{1/2}.
\end{eqnarray}
In that case, Eq. (\ref{rep7}) takes the form of Eq. (\ref{cos5}) and its solutions are given by Eqs. (\ref{cos8}) and (\ref{cos9}). Returning to the notations of this section, we find for $0\le q<1$ that
\begin{eqnarray}
\label{rep9}
\frac{1}{\sqrt{1-q}}\biggl\lbrace \sqrt{(\frac{a}{a_0})^3-q}-\sqrt{q}\arctan \sqrt{\frac{(\frac{a}{a_0})^3-q}{q}}\biggr\rbrace\nonumber\\
=\frac{3}{2}\sqrt{\Omega_0}H_0t,
\end{eqnarray}
which was previously obtained by Harko (2011). On the other hand, for $q\le 0$, we obtain
\begin{eqnarray}
\label{rep10}
\frac{1}{\sqrt{1-q}}\biggl\lbrace \sqrt{(\frac{a}{a_0})^3-q}-\sqrt{|q|}\ln \left (\frac{\sqrt{|q|}+ \sqrt{(\frac{a}{a_0})^3-q}}{(\frac{a}{a_0})^{3/2}}\right )\biggr\rbrace\nonumber\\
=\frac{3}{2}\sqrt{\Omega_0}H_0 t.\qquad
\end{eqnarray}
Finally, for $q=0$, we recover the Einstein-de Sitter model
\begin{eqnarray}
\label{rep11}
\frac{a}{a_0}=\left (\frac{3}{2}\sqrt{\Omega_0}\right )^{2/3}(H_0t)^{2/3}.
\end{eqnarray}
The curves giving the scale factor $a(t)$ as a function of time for different values of $q$ are plotted in Figs. \ref{becharkoPLUS} and \ref{becharkoMOINS}. The interest of this representation is that it shows how the BEC models (\ref{rep9}) and (\ref{rep10}) approach the Einstein-de Sitter model (\ref{rep11}) as $q\rightarrow 0$. On the other hand, for larger values of $q$ (positive or negative), there can be substantial differences between a BEC universe and the classical pressureless Einstein-de Sitter universe.
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{becharkoPLUS.eps}
\caption{Scale factor $a(t)$ as a function of time $t$ for a BEC universe with $k\ge 0$. The units of space and time are normalized by present-day quantities that do not depend on $k$. The different curves correspond to $q=0$ (EdS, dashed), $q=0.01$, $q=0.1$ and $q=0.5$. The straight line corresponds to $a/a_0=1$. Its intersection with the curve $a(t)$ defines the age of the universe.}
\label{becharkoPLUS}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{becharkoMOINS.eps}
\caption{Scale factor $a(t)$ as a function of time $t$ for a BEC universe with $k\le 0$. The units of space and time are normalized by present-day quantities that do not depend on $k$. The different curves correspond to $q=0$ (EdS, dashed), $q=-0.01$, $q=-0.1$ and $q=-0.5$.}
\label{becharkoMOINS}
\end{center}
\end{figure}
To see that, let us consider the asymptotic behaviors of Eqs. (\ref{rep9}) and (\ref{rep10}). For $t=0$, we find that
\begin{eqnarray}
\label{rep16}
\frac{a}{a_0}=R_{\pm}|q|^{1/3},
\end{eqnarray}
with $R_+=1$ for $k>0$ and $R_{-}\simeq 0.7601$ for $k<0$. For $t\rightarrow +\infty$, we get
\begin{eqnarray}
\label{rep12}
\frac{a}{a_0}\sim \left (\frac{3}{2}\sqrt{\Omega_0}\right )^{2/3}(1-q)^{1/3}(H_0t)^{2/3}.
\end{eqnarray}
For $0<q<1$, the scale factor is asymptotically smaller than in an EdS universe (corresponding to $q=0$). Since a BEC universe with $k>0$ initially starts with a radius $a(0)>0$, this implies that the curves must cross each other at some point (such a crossing is shown in Fig. \ref{becharkoPLUS} for $q=0.5$). Of course, the crossing point occurs at larger and larger times as $q\rightarrow 0$. For $q<0$, the scale factor is always larger than in an EdS universe.
For $q<0$, the inflexion point indicating when the universe starts decelerating is located at
\begin{eqnarray}
\label{rep17}
\frac{a_c}{a_0}=2^{1/3}|q|^{1/3},
\end{eqnarray}
\begin{eqnarray}
\label{rep17b}
\frac{3}{2}\sqrt{\Omega_0}H_0t_c=\left\lbrack \sqrt{3}-\ln\left (\frac{1+\sqrt{3}}{\sqrt{2}}\right )\right\rbrack \left (\frac{|q|}{1-q}\right )^{1/2}.\nonumber\\
\end{eqnarray}
For $q\rightarrow -\infty$, $a_c\rightarrow +\infty$ and $\frac{3}{2}\sqrt{\Omega_0}H_0t_c\rightarrow 1.074$. On the other hand for $t\rightarrow -\infty$, the scale factor behaves like
\begin{eqnarray}
\label{rep18}
\frac{a}{a_0}\propto e^{\sqrt{\frac{1+|q|}{|q|}\Omega_0}H_0t}.
\end{eqnarray}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{age.eps}
\caption{Age of a BEC universe as a function of $q\ge 0$. The Einstein-de Sitter model corresponds to $q=0$. }
\label{age}
\end{center}
\end{figure}
Let us naively assume that all the content of the universe is in the
form of BEC dark matter and take $\Omega_0=1$. The age of a BEC
universe corresponds to the time $t_0$ at which $a/a_0=1$. For $0\le q<1$,
\begin{eqnarray}
\label{rep13}
H_0 t_0=\frac{2}{3}\left\lbrace 1-\sqrt{\frac{q}{1-q}}\arctan \sqrt{\frac{1-q}{q}}\right\rbrace,
\end{eqnarray}
and for $q\le 0$,
\begin{eqnarray}
\label{rep13b}
H_0 t_0=\frac{2}{3}\biggl\lbrace 1-\sqrt{\frac{-q}{1-q}}\ln \left ({\sqrt{-q}+ \sqrt{1-q}}\right )\biggr\rbrace.
\end{eqnarray}
The age of the EdS universe ($q=0$) is
\begin{eqnarray}
\label{rep15}
H_0 t_0=\frac{2}{3}.
\end{eqnarray}
As can be seen in Fig. \ref{age}, the age of a BEC universe with $q\neq 0$ is smaller than the age of the EdS universe ($q=0$).
\subsection{Contribution of radiation, baryons and dark energy}
\label{sec_contribution}
Several observational results indicate that the universe is flat ($\kappa=0$) and that its present-day expansion is accelerating ($\ddot a >0$). Furthermore, the estimated age of the universe is
$t_0\sim 13.75\, {\rm Gyrs}$ which is about $H_0^{-1}$. If we assume that the universe is made only of a pressureless
fluid and if we take $\Lambda=0$, we are led to the EdS
universe. However, this model leads to a decelerating expansion and
predicts a too small age of the universe $t_0=2/3H_0\sim 9.3\, {\rm Gyrs}$. If we assume that dark
matter is a BEC, we again find that the present-day universe is
decelerating. Furthermore, its predicted age is even smaller than in
the EdS universe (see Sec. \ref{sec_rep}). Therefore, we must invoke some form of dark energy to understand
the present-day acceleration of the universe and its age. In the
standard $\Lambda$CDM model, dark energy is due to the cosmological
constant. On the other hand, to be complete, we must include the contribution of dark matter, baryonic matter and radiation. In that case, Eq. (\ref{rep7}) is replaced by
\begin{eqnarray}
\label{rep19}
\frac{\dot a}{a}=H_0\sqrt{\frac{\Omega_{B,0}}{(a/a_0)^3}+\frac{\Omega_{rad,0}}{(a/a_0)^4}
+\frac{\Omega_{DM,0}(1-q)}{(a/a_0)^3-q}+\Omega_{\Lambda}},\nonumber\\
\end{eqnarray}
where $\Omega_{B,0}$, $\Omega_{rad,0}$, $\Omega_{DM,0}$ and $\Omega_{\Lambda}$ are the present-day values of the density parameters of the baryonic matter, radiation, dark matter and dark energy. We have assumed that dark matter is in the form of BEC. The case of a pressureless dark matter is recovered for $q=0$. Following Harko (2011), we have adopted the numerical values $\Omega_{B,0}=0.0456$, $\Omega_{rad,0}=8.24\, 10^{-5}$, $\Omega_{DM,0}=0.228$ and $\Omega_{\Lambda}=0.726$ (Hinshaw et al. 2009).
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{darkPOS.eps}
\caption{Scale factor $a(t)$ as a function of time $t$ for a universe filled with baryonic matter, radiation, BEC dark matter with $k\ge 0$, and dark energy. The different curves correspond to $q=0$ (pressureless dark matter, dashed), $q=0.01$, $q=0.1$, $q=0.303471$ and $q=0.5$. }
\label{darkPOS}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{darkNEG.eps}
\caption{Scale factor $a(t)$ as a function of time $t$ for a universe filled with baryonic matter, radiation, BEC dark matter with $k\le 0$, and dark energy. The different curves correspond to $q=0$ (pressureless dark matter, dashed), $q=-0.01$, $q=-0.1$ and $q=-0.5$. }
\label{darkNEG}
\end{center}
\end{figure}
The evolution of the scale factor $a(t)$ is represented in Figs. \ref{darkPOS} and \ref{darkNEG} for $k>0$ and $k<0$ respectively. In both cases, the universe is initially decelerating but finally enters in a phase of acceleration due to the effect of dark energy (cosmological constant). In the case $k>0$, previously treated by Harko (2011), the scale factor increases more rapidly than in the standard model ($k=0$). Furthermore, the density of dark matter becomes infinite at a finite radius so that radiation does not dominate in the early universe. Accordingly, the universe starts from $a(0)>0$ at $t=0$. In the case $k<0$, the scale factor increases less rapidly than in the standard model ($k=0$). Furthermore, the density of dark matter does not diverges when $a\rightarrow 0$ so that radiation dominates in the early universe. Accordingly, the universe starts from $a(0)=0$ with an infinite (radiation) density.
\subsection{Theory of linear perturbations}
\label{sec_tl}
The theory of linear perturbations based on the Newtonian equations with pressure (\ref{b1})-(\ref{b3}) has been performed by Reis (2003) in the general case, so that we shall directly apply his results to the present situation. Note that Harko (2011) considers other equations and obtains different results. Nevertheless, our main conclusions will be the same. For an arbitrary equation of state, the evolution of the density contrast is given by (Reis, 2003):
\begin{eqnarray}
\label{tl1}
\ddot\delta-H\left\lbrack 3 (2w-c_{eff}^2-c_s^2)-2\right\rbrack\dot\delta\nonumber\\
+3H^2\biggl\lbrack \frac{1}{H}\dot{c_{eff}^2}
+(3c_s^2-6w-1)c_{eff}^2\nonumber\\
+\frac{3w^2}{2}-4w-\frac{1}{2}
+3c_s^2\biggr\rbrack \delta-\frac{c_{eff}^2c^2}{a^2}\Delta\delta=0,
\end{eqnarray}
where $H=\dot a/a$ is the Hubble constant, $w=p_b/\rho_b c^2$, $c_{eff}^2=(\delta p/\delta \rho)/c^2$ and $c_s^2=p_b'(\rho_b)/c^2$. These last quantities represent the velocity of sound normalized by the velocity of light. For a barotropic fluid, $c_{eff}^2=c_s^2=p'(\rho_b)/c^2$. On the other hand, from the Friedmann equations (\ref{b5})-(\ref{b7}), it is easy to establish the relation
\begin{eqnarray}
\label{tl2}
\frac{\dot w}{1+w}=-3H(c_s^2-w).
\end{eqnarray}
We shall now focus on the equation of state (\ref{b4}). For this equation of state, we obtain $w=k\rho_b$ and $c_{eff}^2=c_s^2=2w$. In that case, Eq. (\ref{tl1}) reduces to
\begin{equation}
\label{tl3}
\ddot\delta+2H (3w+1)\dot\delta-3H^2 \left (\frac{9w^2}{2}+6w+\frac{1}{2}\right ) \delta
-\frac{2wc^2}{a^2}\Delta\delta=0,
\end{equation}
where we have used $\dot w/H=-3w(1+w)$ according to Eq. (\ref{tl2}). Measuring the evolution in terms of $a$ rather than in terms of $t$, and using the Friedmann equations (\ref{b6}) and (\ref{b7}), we can rewrite Eq. (\ref{tl3}) in the form
\begin{equation}
\label{tl4}
a^2\frac{d^2 \delta}{da^2}+\frac{3}{2}(1+3w) a \frac{d\delta}{da}-\frac{3}{2} (9w^2+12w+1)\delta -\frac{2wc^2}{H^2a^2}\Delta\delta=0.
\end{equation}
Finally, according to Eq. (\ref{cos7b}), we have
\begin{eqnarray}
\label{tl5}
w=\frac{\pm 1}{R^3\mp 1},
\end{eqnarray}
where the upper sign corresponds to $k>0$ and the lower sign corresponds to $k<0$. If, for simplicity, we ignore the Laplacian term in Eq. (\ref{tl4}), we obtain
\begin{eqnarray}
\label{tl6}
R^2\frac{d^2 \delta}{dR^2}+\frac{3}{2}(1+3w) R \frac{d\delta}{dR}-\frac{3}{2} (9w^2+12w+1)\delta=0.\nonumber\\
\end{eqnarray}
For $R\rightarrow +\infty$, $w\rightarrow 0$ and Eq. (\ref{tl6}) reduces to
\begin{eqnarray}
\label{tl7}
R^2\frac{d^2 \delta}{dR^2}+\frac{3}{2} R \frac{d\delta}{dR}-\frac{3}{2}\delta=0.
\end{eqnarray}
This returns the usual equation (\ref{eds2}) of a pressureless fluid in the Einstein-de Sitter universe. Its solutions are given by Eq. (\ref{eds3}).
Let us now consider the case of small $R$. For $k>0$, we have seen that the universe exists only for $R\ge 1$. Now, for $R\rightarrow 1$, $w\rightarrow +\infty$ so that the differential equation (\ref{tl6}) is ill-defined. This means that we must start the perturbation analysis at $t=t_i>0$. Harko (2011) remarks that the requirement $c_{eff}=c_s<1$ (meaning that the velocity of sound must be smaller than the velocity of light) implies $R>3^{1/3}$ and $w<1/2$. Therefore, we shall assume that the universe starts at $t_*$ such that $R(t_*)=3^{1/3}$. According to Eq. (\ref{cos8}), we have
\begin{equation}
\label{cos10edf}
\frac{3}{2}Kt_*= \sqrt{2}-\arctan\sqrt{2}\simeq 0.4589.
\end{equation}
For $t\rightarrow t_*$, Eq. (\ref{tl6}) can be approximated by
\begin{eqnarray}
\label{tl6bis}
R^2\frac{d^2 \delta}{dR^2}+\frac{15}{4} R \frac{d\delta}{dR}-\frac{111}{8}\delta=0,
\end{eqnarray}
and its solutions are
\begin{eqnarray}
\label{tl11bis}
\delta \propto R^{\frac{-11+\sqrt{1009}}{8}},\qquad \delta \propto R^{\frac{-11-\sqrt{1009}}{8}}.
\end{eqnarray}
The growth of the perturbations is faster than in a cold EdS universe ($a^{2.6}$ instead of $a$). On the other hand, for $k<0$, the universe starts at $R=0$ for $t\rightarrow -\infty$ implying $w=-1$.
In that case, Eq. (\ref{tl6}) can be approximated by
\begin{eqnarray}
\label{tl10}
R^2\frac{d^2 \delta}{dR^2}-3 R \frac{d\delta}{dR}+3\delta=0,
\end{eqnarray}
and its solutions are
\begin{eqnarray}
\label{tl11}
\delta \propto R^{3},\qquad \delta \propto R.
\end{eqnarray}
Contrary to a classical cold universe, the two solutions of the differential equation (\ref{tl10}) are growing. Furthermore, the growth of the perturbations is faster than in an EdS universe ($a^3$ instead of $a$). Combining Eq. (\ref{tl11}) with Eq. (\ref{cos11}), we get
\begin{eqnarray}
\label{tl12}
\delta(t) \propto e^{3Kt},\qquad \delta(t) \propto e^{Kt},
\end{eqnarray}
for $t\rightarrow -\infty$.
In conclusion, these simple estimates indicate that the growth of the
perturbations is faster in a BEC universe than in a pressureless
universe. This conclusion was previously reached by Harko (2011) for
$k>0$ based on different equations. Our approach confirms this
conclusion and extends it to $k<0$ (It may be recalled that the
Laplacian term in Eq. (\ref{tl3}) has been neglected in our simple
analysis. Now, we have seen in Sec. \ref{sec_s} that it is precisely
this term that leads to an increase of the maximum growth rate in the
(static) Jeans problem when $k<0$. Therefore, the inclusion of this
term in our analysis should enhance the growth of perturbations in the
case of attractive self-interaction).
\section{Dark fluid with a generalized equation of state}
\label{sec_dark}
\subsection{Linear equation of state}
\label{sec_leos}
One possibility to understand the present-day acceleration of the
expansion of the universe is to invoke a form of dark energy arising
from a non-zero value of the cosmological constant $\Lambda$ (see
Sec. \ref{sec_contribution}). However, the physical meaning of
$\Lambda$ is not clearly understood. A major problem is that most
quantum field theories predict a huge cosmological constant from the
energy of the quantum vacuum, more than $100$ order of magnitude too
large. Therefore, other approaches have been developed to understand
the phase of acceleration without invoking the cosmological
constant. One possibility is to consider a ``dark fluid'' with a
negative pressure. In this context, many workers (see the review of
Peebles \& Ratra 2003) have considered a linear equation of state
\begin{equation}
\label{dark0}
p=\alpha \rho c^2,
\end{equation}
with $-1\le \alpha\le 1$. For this equation of state, the Friedmann equations (\ref{b5}), (\ref{b6}) and (\ref{b7}) with $\kappa=\Lambda=0$ reduce to
\begin{equation}
\label{dark1}
\frac{d\rho_b}{dt}+3\frac{\dot a}{a}(1+\alpha)\rho_b=0,
\end{equation}
\begin{equation}
\label{dark2}
\frac{\ddot a}{a}=-\frac{4\pi G}{3} (1+3\alpha)\rho_{b},\qquad \left (\frac{\dot a}{a}\right )^2=\frac{8\pi G}{3}\rho_b.
\end{equation}
Equation (\ref{dark1}) leads to the relation $\rho_{b}a^{3(1+\alpha)}\sim 1$. We note that the effect of a cosmological constant $\Lambda$ is equivalent to a fluid with an equation of state $p=-\rho c^2$ ($\alpha=-1$) since, in that case, $\rho_b$ is constant and can be written $\rho_b=\rho_\Lambda\equiv \Lambda/8\pi G$. This equation of state leads to an exponential growth of the scale factor
\begin{equation}
\label{dark3bb}
a=a_0 e^{\sqrt{\Lambda/3}t}.
\end{equation}
On the other hand, for $-1<\alpha\le 1$, Eqs. (\ref{dark1}) and (\ref{dark2}) generate a model of the form
\begin{equation}
\label{dark3}
a\propto t^{2/\lbrack 3(1+\alpha)\rbrack},\qquad H=\frac{\dot a}{a}=\frac{2}{3(1+\alpha)t},
\end{equation}
\begin{equation}
\label{dark4}
\rho_b=\frac{1}{6\pi G(1+\alpha)^2 t^2}.
\end{equation}
According to Eq. (\ref{dark2}-a), this universe is accelerating for $-1<
\alpha<\alpha_{c}\equiv -1/3$ and decelerating for
$\alpha>\alpha_{c}$. For $\alpha=\alpha_c$, the scale factor increases
linearly with time ($a\propto t$). The EdS universe corresponds to
$\alpha=0$. For the general model (\ref{dark3})-(\ref{dark4}), the age
of the universe is $t_0=2/\lbrack 3(1+\alpha)H_{0}\rbrack$ where
$H_0=2.273\, 10^{-18}\, {\rm s}^{-1}$ is the present-day value of the
Hubble constant. We note that for the critical value $\alpha_c=-1/3$,
the age of the universe is $t_0=H_0^{-1}=13.95\, {\rm Gyrs}$ which is
close to the value $13.75\, {\rm Gyrs}$ predicted by the
$\Lambda$CDM model. However, the deceleration parameter $Q=-{\ddot
a}a/{\dot a}^2=(1+3\alpha)/2$ vanishes for $\alpha=-1/3$ while its
present-day value is $\sim -0.5$. We also note that the linear equation
of state (\ref{dark0}) does not allow for a transition between a phase
of deceleration and a phase of acceleration, while such a transition
is likely in our universe. It may be therefore interesting to
generalize this model.
\subsection{Generalized equation of state}
\label{sec_geos}
We consider a generalized equation of state of the form
\begin{equation}
\label{dark5}
p=(\alpha \rho+k\rho^2) c^2,
\end{equation}
with $-1\le \alpha\le 1$ and $k$ positive or negative (the case $\alpha=-1$ is specifically treated in Appendix \ref{sec_eosg}). In this approach, a single ``dark fluid'' combines the properties of a BEC dark matter described by the equation of state (\ref{b4}) and of a dark energy described by the equation of state (\ref{dark0}). We note that the equation of state of the BEC dark matter ($\propto \rho^2$) dominates in the early universe where the density is high while the dark energy ($\propto \rho$) dominates in the present-day universe where the density is low. The equation of state of the BEC also dominates at the scale of dark matter halos. This makes the study of this equation of state interesting. Another nice feature of this equation of state is that it admits fully analytical solutions\footnote{Eq. (\ref{dark5}) can be written as $p_b=w(t)\rho_b c^2$ where $w(t)$ is a function of time. Many authors have considered an equation of state of that form with some prescribed function $w(t)$. In our model, $w(t)$ is not an {\it ad hoc} function but it is explicitly given by $w(t)=\alpha+k\rho_b(t)$.}.
For the equation of state (\ref{dark5}), the Friedmann equation (\ref{b5}) becomes
\begin{equation}
\label{dark6}
\frac{d\rho_b}{dt}+3\frac{\dot a}{a}\rho_b (1+\alpha+k\rho_b)=0.
\end{equation}
This equation can be integrated into
\begin{equation}
\label{dark7}
\rho_b=\frac{A(1+\alpha)}{a^{3(1+\alpha)}-kA},
\end{equation}
where $A>0$ is a constant. For $a\rightarrow +\infty$, $\rho_b\sim
A(1+\alpha)/a^{3(1+\alpha)}$. When $k>0$, the density
exists only for $a>a_*=(kA)^{1/\lbrack 3(1+\alpha)\rbrack}$. For $a\rightarrow a_*$,
$\rho_b\rightarrow +\infty$. When $k<0$, the density is defined for
all $a$ and has a finite value $\rho_b=(1+\alpha)/|k|$ when $a\rightarrow 0$.
Combining Eqs. (\ref{b6}) and (\ref{dark5}), and taking $\Lambda=0$, we obtain
\begin{equation}
\label{dark8}
\frac{\ddot a}{a}=-\frac{4\pi G\rho_b}{3}(1+3\alpha+3k\rho_b).
\end{equation}
Let us define a critical density and a critical scale factor
\begin{equation}
\label{dark9}
\rho_c= -\frac{1+3\alpha}{3k},\qquad a_c=\left (\frac{-2kA}{1+3\alpha}\right )^{\frac{1}{3(1+\alpha)}},
\end{equation}
corresponding to a possible inflexion point ($\ddot a=0$) in the curve
$a(t)$. When $k>0$ and $\alpha\ge -1/3$, the universe is always
decelerating ($\ddot a<0$). When $k>0$ and $\alpha<-1/3$, the universe
is decelerating for $\rho_b>\rho_c$ (i.e. $a<a_c$) and accelerating
for $\rho_b<\rho_c$ (i.e. $a>a_c$). When $k<0$ and $\alpha\le -1/3$,
the universe is always accelerating ($\ddot a>0$). When $k<0$ and
$\alpha>-1/3$, the universe is accelerating for $\rho_b>\rho_c$ (i.e.
$a<a_c$) and decelerating for $\rho_b<\rho_c$ (i.e. $a>a_c$).
In order to determine the temporal evolution of $a(t)$, we shall assume that the universe is flat ($\kappa=0$). Combining Eqs. (\ref{b7}) and (\ref{dark7}), we get
\begin{equation}
\label{dark10}
\dot a=\left \lbrack\frac{8\pi GA}{3}(1+\alpha)\right \rbrack^{1/2}\frac{a}{\sqrt{a^{3(1+\alpha)}-kA}}.
\end{equation}
For $a\rightarrow +\infty$, we recover the solution (\ref{dark3})-(\ref{dark4}). It is convenient to define $a_*=(|k|A)^{1/\lbrack 3(1+\alpha)\rbrack}$ and introduce $R=a/a_*$. In that case, the density is given by
\begin{equation}
\label{dark14}
\rho_b=\frac{1+\alpha}{|k|}\frac{1}{R^{3(1+\alpha)}\mp 1},
\end{equation}
and Eq. (\ref{dark10}) can be rewritten
\begin{equation}
\label{dark11}
\dot R=\frac{KR}{\sqrt{R^{3(1+\alpha)}\mp 1}},
\end{equation}
where the upper sign $-$ corresponds to $k>0$ and the lower sign $+$ corresponds to $k<0$. On the other hand, we have defined the constant
\begin{equation}
\label{dark12}
K=\left \lbrack \frac{8\pi G}{3|k|}(1+\alpha)\right \rbrack^{1/2}.
\end{equation}
For $k>0$, the solution of Eq. (\ref{dark11}) is
\begin{equation}
\label{dark16}
\sqrt{R^{3(1+\alpha)}-1}-\arctan \sqrt{R^{3(1+\alpha)}-1}=\frac{3(1+\alpha)}{2}Kt,
\end{equation}
where the constant of integration has been set equal to zero. In this model, the universe starts at a finite time $t=0$ (see Sec. \ref{sec_gtlp} for a revision of this statement) with a finite radius $R(0)=1$ and an infinite density $\rho_b(0)=\infty$. For $t\rightarrow 0$, $R\simeq 1+\lbrack 3/(4(1+\alpha))\rbrack^{1/3}(Kt)^{2/3}$. For $t\rightarrow +\infty$, it asymptotically approaches the solution (\ref{dark3})-(\ref{dark4}) i.e. $R\sim (3(1+\alpha)Kt/2)^{2/\lbrack 3(1+\alpha)\rbrack}$. When $\alpha\ge -1/3$, the universe is always decelerating ($\ddot R<0$). When $\alpha<-1/3$, the universe is decelerating for $\rho_b>\rho_c$ (i.e. $R<R_c^{+}\equiv ({-2}/(1+3\alpha))^{{1}/\lbrack {3(1+\alpha)}\rbrack}$) and accelerating for $\rho_b<\rho_c$ (i.e. $R>R_c^+$). The time at which the universe starts accelerating is
\begin{equation}
\label{dark17}
\frac{3(1+\alpha)}{2}Kt_c=\sqrt{\frac{-3(1+\alpha)}{1+3\alpha}}-\arctan \sqrt{\frac{-3(1+\alpha)}{1+3\alpha}}.
\end{equation}
Some possible evolutions of $R(t)$ corresponding to $\alpha<-1/3$, $\alpha=1/3$ and $\alpha>1/3$ are represented in Fig. \ref{generalisationPOS}.
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{generalisationPOS.eps}
\caption{Evolution of the scale factor in the case $k>0$ for different values of $\alpha$ (specifically $\alpha=-2/3$, $\alpha=-1/3$ and $\alpha=1/2$).}
\label{generalisationPOS}
\end{center}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{generalisationNEG.eps}
\caption{Evolution of the scale factor in the case $k<0$ for different values of $\alpha$ (specifically $\alpha=-2/3$, $\alpha=-1/3$ and $\alpha=1/2$).}
\label{generalisationNEG}
\end{center}
\end{figure}
For $k<0$, the solution of Eq. (\ref{dark11}) is
\begin{equation}
\label{dark18}
\sqrt{R^{3(1+\alpha)}+1}-\ln \left (\frac{1+\sqrt{R^{3(1+\alpha)}+1}}{R^{3(1+\alpha)/2}}\right )=\frac{3(1+\alpha)}{2}Kt,
\end{equation}
where the constant of integration has been set equal to zero. In this model, the universe starts from $t\rightarrow -\infty$ with a vanishing radius $R(-\infty)=0$ and a finite density $\rho_b(-\infty)=(1+\alpha)/|k|$. For $t\rightarrow +\infty$, it asymptotically approaches the solution (\ref{dark3})-(\ref{dark4}) i.e. $R\sim (3(1+\alpha)Kt/2)^{2/\lbrack 3(1+\alpha)\rbrack}$. When $\alpha\le -1/3$, the universe is always accelerating ($\ddot a>0$). When $\alpha>-1/3$, the universe is accelerating for $\rho_b>\rho_c$ (i.e. $R<R_c^{-}\equiv ({2}/(1+3\alpha))^{{1}/\lbrack {3(1+\alpha)}\rbrack}$) and decelerating for $\rho_b<\rho_c$ (i.e. $a>a_c$). The time at which the universe starts decelerating is
\begin{equation}
\label{dark19}
\frac{3(1+\alpha)}{2}Kt_c=\sqrt{\frac{3(1+\alpha)}{1+3\alpha}}-\ln \left (\frac{\sqrt{1+3\alpha}+\sqrt{3(1+\alpha)}}{\sqrt{2}}\right ).
\end{equation}
For $t\rightarrow -\infty$,
$R\rightarrow 0$ and the asymptotic expansion of
Eq. (\ref{dark18}) yields
\begin{equation}
\label{dark20}
R(t)\propto e^{Kt}.
\end{equation}
Some possible evolutions of $R(t)$ corresponding to $\alpha<-1/3$, $\alpha=1/3$ and $\alpha>1/3$ are represented in Fig. \ref{generalisationNEG}.
\subsection{Optimal parameters}
\label{sec_op}
We can easily extend the analysis of Sec. \ref{sec_rep} to the equation of state (\ref{dark5}). The density can be written
\begin{eqnarray}
\label{dark21}
\rho_b=\rho_0\frac{1-q}{(\frac{a}{a_0})^{3(1+\alpha)}-q},
\end{eqnarray}
where
\begin{eqnarray}
\label{dark22}
q=\frac{k\rho_0}{1+\alpha+k\rho_0}, \qquad {\rm implying} \qquad k\rho_0=\frac{q(1+\alpha)}{1-q}.
\end{eqnarray}
We note that $q$ is positive for $k>0$ and negative for $k<0$. Furthermore, $q=0$ for $k=0$, $q\rightarrow 1$ for $k\rightarrow +\infty$ and $q\rightarrow -\infty$ for $k\rightarrow -(1+\alpha)/\rho_0$. Using these results, the differential equation (\ref{b7}) giving the evolution of the scale factor with the time can be rewritten
\begin{eqnarray}
\label{dark23}
\dot a=H_0\sqrt{\Omega_0}(1-q)^{1/2}\frac{a}{\sqrt{(\frac{a}{a_0})^{3(1+\alpha)}-q}}.
\end{eqnarray}
We can make the connection with the previous notations by setting
\begin{eqnarray}
\label{dark24}
a_*=a_0 |q|^{1/\lbrack 3(1+\alpha)\rbrack},\qquad K=H_0\sqrt{\Omega_0}\left (\frac{1-q}{|q|}\right )^{1/2}.\qquad
\end{eqnarray}
In that case, Eq. (\ref{dark23}) takes the form of Eq. (\ref{dark11}) and its solutions are given by Eqs. (\ref{dark16}) and (\ref{dark18}). Returning to the notations of this section, we find that the evolution of $a(t)$ is given by Eqs. (\ref{rep9})-(\ref{rep11}) where all the $3$ are replaced by $3(1+\alpha)$. The other equations can be easily generalized.
Let us assume that all the content of the universe is in the dark fluid with the generalized equation of state (\ref{dark5}) so that $\Omega_0=1$. Let us call $t_0$ the time at which $a/a_0=1$. For $0<q<1$,
\begin{eqnarray}
\label{dark25}
H_0 t_0=\frac{2}{3(1+\alpha)}\left\lbrace 1-\sqrt{\frac{q}{1-q}}\arctan \sqrt{\frac{1-q}{q}}\right\rbrace,
\end{eqnarray}
for $q<0$,
\begin{eqnarray}
\label{dark26}
H_0 t_0=\frac{2}{3(1+\alpha)}\left\lbrace 1-\sqrt{\frac{-q}{1-q}}
\ln(\sqrt{-q}+\sqrt{1-q})\right\rbrace,\nonumber\\
\end{eqnarray}
and for $q=0$,
\begin{eqnarray}
\label{dark27}
H_0 t_0=\frac{2}{3(1+\alpha)}.
\end{eqnarray}
These expressions basically give the age of the universe $t_0$ as a function of $q$ and $\alpha$.
The deceleration parameter is defined by
\begin{eqnarray}
\label{dark28}
Q=-\frac{{\ddot a} a}{{\dot a}^2}=\frac{1+3w(t)}{2},
\end{eqnarray}
where $w(t)\equiv p_b/\rho_b c^2$. For the equation of state (\ref{dark5}), $w(t)=\alpha+k\rho_b$. Using Eq. (\ref{dark22}), we find that the present-day value of the deceleration parameter can be expressed in our model as
\begin{eqnarray}
\label{dark29}
Q_0=\frac{3\alpha+2q+1}{2(1-q)}.
\end{eqnarray}
The observed value of $Q_0$ is close to $-0.5$. If we take $Q_0=-1/2$, Eq. (\ref{dark29}) yields $\alpha=-(q+2)/3$. On the other hand, in the standard model, the age of the universe is about $13.75\, {\rm Gyrs}$. If we take $t_0=H_0^{-1}$ (which is close to this value) and use the previous relation between $\alpha$ and $q$, Eqs. (\ref{dark25}) and (\ref{dark26}) determine the optimal value of $q$, then $\alpha$. For $0<q<1$, the optimal value of $q$ is given by
\begin{eqnarray}
\label{dark30}
\frac{1+q}{2}-\sqrt{\frac{q}{1-q}}\arctan \sqrt{\frac{1-q}{q}}=0.
\end{eqnarray}
We find $q=0.303471$ and $\alpha=-0.76782367$. For $q<0$, the optimal value of $q$ is given by
\begin{eqnarray}
\label{dark31}
\frac{1+q}{2}-\sqrt{\frac{-q}{1-q}}\ln(\sqrt{-q}+\sqrt{1-q}) =0.
\end{eqnarray}
We find $q=-0.383589$ and $\alpha=-0.53880367$. The evolution of the scale factor $a(t)$ in a universe filled with a dark fluid with equation of state (\ref{dark5}) is represented in Fig. \ref{modeleoptimal} for the optimal values of $(q,\alpha)$ obtained previously. It gives a relatively good agreement with the standard $\Lambda$CDM model based on a non-vanishing value of the cosmological constant.
\begin{figure}[!h]
\begin{center}
\includegraphics[clip,scale=0.3]{modeleoptimalNEG.eps}
\caption{The full line represents the scale factor $a(t)$ as a function of time $t$ for a universe filled with a dark fluid with equation of state (\ref{dark5}) with optimal parameters ($q=0.303471$, $\alpha=-0.76782367$) and ($q=-0.383589$, $\alpha=-0.53880367$). The long-dashed line corresponds to the standard model (see Figs. \ref{darkPOS} and \ref{darkNEG}). By construction, these three curves give the same values of the age of the universe $t_0\sim H_0^{-1}$ and of the present-day deceleration parameter $Q_0\sim -0.5$. The model with $k>0$ is decelerating then accelerating. Using Eqs. (\ref{dark17}) and (\ref{dark24}), we find that the acceleration starts at $H_0t_c\sim 0.2$ which is earlier than the value $H_0t_c\sim 0.5$ predicted by the standard model. The model with $k<0$ is always accelerating. We have also indicated the model corresponding to a linear equation of state ($q=0$) with $\alpha=-1/3$ yielding $a/a_0=H_0t$. It gives the correct age of the universe but has a vanishing deceleration parameter. Finally, we have indicated the (pressureless) EdS model corresponding to $\alpha=q=0$ yielding $a/a_0=(3H_0t/2)^{2/3}$. It does not give the correct age of the universe and is decelerating instead of accelerating. The other models with $\alpha=0$ and $q\neq 0$ do not do better. This shows that we must combine the two equations of state $\alpha\rho$ and $k\rho^2$ to obtain a good agreement with the standard model, i.e. we need an equation of state with two parameters. }
\label{modeleoptimal}
\end{center}
\end{figure}
\subsection{Theory of linear perturbations}
\label{sec_gtlp}
In this section, we extend to the case of the generalized equation of
state (\ref{dark5}) the theory of linear perturbations developed in
Sec. \ref{sec_tl}. For this equation of state, $w=\alpha+k\rho_b$ and
$c_{eff}^2=c_s^2=2w-\alpha$. According to
Eq. (\ref{dark14}), we have
\begin{eqnarray}
\label{dark32}
w=\alpha\pm\frac{1+\alpha}{R^{3(1+\alpha)}\mp 1},
\end{eqnarray}
where the upper sign corresponds to $k>0$ and the lower sign corresponds to $k<0$. The generalization of Eq. (\ref{tl6}) is
\begin{eqnarray}
\label{dark33}
R^2\frac{d^2 \delta}{dR^2}+\frac{3}{2}(1+3w-4\alpha) R \frac{d\delta}{dR}\nonumber\\
-\frac{3}{2} (9w^2+12w+1-8\alpha-6\alpha^2)\delta=0.
\end{eqnarray}
For $R\rightarrow +\infty$, $w\rightarrow \alpha$ and Eq. (\ref{dark33}) reduces to
\begin{eqnarray}
\label{dark34}
R^2\frac{d^2 \delta}{dR^2}+\frac{3}{2}(1-\alpha) R \frac{d\delta}{dR}-\frac{3}{2}(\alpha+1)(3\alpha+1)\delta=0.\nonumber\\
\end{eqnarray}
This is the equation for the density contrast corresponding to the linear equation of state (\ref{dark0}). Its solutions are
\begin{eqnarray}
\label{dark35}
\delta \propto R^{1+3\alpha},\qquad \delta \propto R^{-\frac{3}{2}(1+\alpha)}.
\end{eqnarray}
Let us now consider the case of small $R$. For $k>0$, the requirement $c_{eff}=c_s<1$ implies $R>\lbrack(3+\alpha)/(1-\alpha)\rbrack^{1/\lbrack 3(1+\alpha)\rbrack}$ and $w<(\alpha+1)/2$. Therefore, we shall assume that the universe starts at $t_*$ such that $R(t_*)=\lbrack(3+\alpha)/(1-\alpha)\rbrack^{1/\lbrack 3(1+\alpha)\rbrack}$. According to Eq. (\ref{dark16}), we have
\begin{equation}
\label{dark36}
\frac{3(1+\alpha)}{2}Kt_*=\sqrt{\frac{2(1+\alpha)}{1-\alpha}}-\arctan \sqrt{\frac{2(1+\alpha)}{1-\alpha}}.
\end{equation}
For $t\rightarrow t_*$, Eq. (\ref{dark33}) can be approximated by
\begin{eqnarray}
\label{dark37}
R^2\frac{d^2 \delta}{dR^2}+\frac{15}{4}(1-\alpha) R \frac{d\delta}{dR}+\frac{3}{8}(15\alpha^2-10\alpha-37)\delta=0,\nonumber\\
\end{eqnarray}
and its solutions are
\begin{eqnarray}
\label{dark38}
\delta_{\pm} \propto R^{\frac{-11+15\alpha\pm\sqrt{-135\alpha^2-90\alpha+1009}}{8}}.
\end{eqnarray}
For $k<0$, the universe starts at $R=0$ for $t\rightarrow -\infty$ implying $w=-1$.
In that case, Eq. (\ref{dark33}) can be approximated by
\begin{eqnarray}
\label{dark39}
R^2\frac{d^2 \delta}{dR^2}-3(1+2\alpha) R \frac{d\delta}{dR}+3(1+\alpha)(1+3\alpha)\delta=0,\quad
\end{eqnarray}
and its solutions are
\begin{eqnarray}
\label{dark40}
\delta \propto R^{3(1+\alpha)},\qquad \delta \propto R^{1+3\alpha}.
\end{eqnarray}
\section{Conclusion}
\label{sec_conclusion}
Following the proposal of B\"ohmer \& Harko (2007) and others (see a short historic in Chavanis 2011), we have assumed that the dark matter in the universe is a self-gravitating BEC with short-range interactions, and we have theoretically explored the consequences of this hypothesis. For the sake of generality, we have considered the case of positive and negative scattering lengths.
At the level of dark matter halos, a positive scattering length, equivalent to a repulsive self-interaction generating a positive pressure, is able to stabilize the halos with respect to gravitational collapse. This leads to dark matter halos without density cusps, equivalent to polytropes of index $n=1$ (more generally, the barotropic equation of state is fixed by the form of the self-interaction so that other configurations are possible). Alternatively, if the scattering length is negative, equivalent to an attractive self-interaction generating a negative pressure, the dark matter is very unstable and collapses above a very small critical mass $M_{max}=1.012\hbar/\sqrt{|a_s|Gm}$ (Chavanis 2011). When these ideas are applied to an infinite homogeneous cosmic fluid (Jeans problem), it is found that a negative scattering length can increase the maximum growth rate of the instability and accelerate the formation of structures. The virtues of these results could be combined by assuming that the scattering length changes sign in the course of the evolution. It could be initially negative to help with the formation of structures and become positive (due to a change of density, magnetic fields, radiation,...) to prevent complete gravitational collapse. The mechanism of this change of sign is, however, unknown so that this idea remains highly speculative. However, some terrestrial experiments have demonstrated that certain atoms can have negative scattering lengths, that their scattering length can depend on the magnetic field, and that it is possible in principle to manipulate the value and the sign of $a_s$ (Fedichev et al. 1996). Therefore, a BEC is a serious candidate for which the pressure can be positive and/or negative (see Appendix \ref{sec_det} for further remarks about the values of the BEC parameters).
At the cosmological level, we have constructed models of universe composed of BEC dark matter with attractive or repulsive self-interaction. We have first studied the academic situation where the universe is made only of BEC dark matter. A BEC universe with positive scattering length, having a positive pressure, is not qualitatively very different from a classical Einstein-de Sitter universe. It also emerges at a primordial time $t=0$ from a big-bang singularity where the density is infinite, and undergoes a decelerating expansion asymptotically equivalent to the EdS universe. A difference, however, is that the initial scale factor $a(0)$ is finite. On the other hand, a BEC universe with negative scattering length, having a negative pressure, markedly differs from previous models. It starts from $t\rightarrow -\infty$ with a vanishing radius and a finite density, has an initial accelerating expansion then decelerates and asymptotically behaves like the EdS universe. This model universe exists for any time in the past and there is no big-bang singularity. When we add the effect of radiation, baryonic matter and dark energy (via the cosmological constant), the picture is different. In that case, a BEC universe with attractive or repulsive self-interaction starts from a singularity at $t=0$ where the density is infinite. It first experiences a phase of decelerating expansion followed by a phase of accelerating expansion. For $k\rightarrow 0$ we recover the standard $\Lambda$CDM model but for $k\neq 0$, the evolution of the scale factor in a BEC universe can be substantially different. The model with $k>0$ expands more rapidly than the standard model ($k=0$). The initial scale factor is finite ($a(0)>0$) and the radiation never dominates. The model with $k<0$ expands less rapidly than the standard model. The initial scale factor vanishes ($a(0)=0$) and the radiation dominates leading to a decelerating expansion. In both models, the dark energy dominates at large times leading to an accelerating expansion. Finally, we have considered a ``dark fluid'' with generalized equation of state $p=(\alpha\rho+k\rho^2)c^2$ having a component $p=k\rho^2 c^2$ similar to a BEC dark matter and a component $p=\alpha\rho c^2$ mimicking the effect of the cosmological constant (dark energy). We have found optimal parameters $(\alpha,k)$ that give a good agreement with the standard model. We have studied the growth of perturbations in these different models and confirmed the previous observation of Harko (2011) that the density contrast increases more rapidly in a BEC universe than in the standard model.
In conclusion, the idea that dark matter could be a BEC is fascinating and probably deserves further research.
|
\section{Introduction.}
\label{intro}
The physics of the heavy mesons and baryons with open and hidden heavy
quarks is very reach and hot topic. Understanding the heavy-meson
physics is important for evaluation of the components of the $CKM$-matrix,
verification of the Standard Model and probing the physics beyond
it, as well as production of different exotic meson states. Currently
the experiments with $B$- and $D$-mesons are intensively studied
by Belle~\cite{Belle},
BaBar~\cite{BABAR}
and CDF collaborations, where unprecedented integrated luminocities
were achieved, as well as neutrino-production of open and hidden charm
in neutrino-hadron processes studied by K2K~\cite{K2K},
MiniBoone~\cite{MiniBooNE},
NuTeV~\cite{NuTeV}
and Minerva~\cite{Minerva} collaborations.
Theoretically, in pre-QCD era some success was achieved by the quantum-mechanical
models which use effective potentials to describe heavy hadrons and
their excitations (see e.g.~\cite{Eichten:1979ms} and references
therein). However, such description inevitably introduces undefined
phenomenological constants. The relation of these constants to QCD
parameters is quite obscure: due to interaction with gluons and virtual
light quark pairs all the constants contain nonperturbative dynamics.
The numerical values of these constants are determined from fits to
experimental data, which limits the predictive power of such models.
An advanced version of the potential model is NRQCD \cite{Bodwin:1994jh},
however in this model light quarks and their interactions with heavy
quarks via gluons is done in a phenomenological way. For this reason
it is limited to description of systems with two heavy quarks. Alternatively,
the heavy mesons are described in the Heavy Quark Effective Theory
(HQET) proposed in~\cite{Isgur:1989}, which treats
the heavy mesons using the pQCD methods but does not take into account
nonperturbative effects.
We propose to study the heavy quark physics in
the framework of the instanton vacuum model. This model was developed
in~\cite{Diakonov} and provided
a consistent description of the light mesons physics~\cite{Musakhanov}.
One of the most prominent advances of the instanton vacuum model
is the correct description of the spontaneous breaking of the chiral
symmetry ($S\chi$SB), which is responsible for properties of most
hadrons and nuclei ~\cite{Leutwyler:2001hn}. The $S\chi$SB is due
to specific properties of QCD vacuum, which is known to be one of
the most complicated objects due to perturbative as well as non-perturbative
fluctuations and is a very important object of investigations by methods
of Nonperturbative Quantum Chromo Dynamics (NQCD). In the instanton
picture $S\chi$SB is due to the delocalization of single-instanton
quark zero modes in the instanton medium. One of the advantages of
the instanton vacuum is that it is characterized by only two parameters:
the average instanton size $\rho\sim0.3\,{\rm fm}$ and the average
inter-instanton distance $R\sim1\,{\rm fm}$. These essential numbers
were suggested in~\cite{Shuryak:1981ff} and were derived from $\Lambda_{\overline{{\rm MS}}}$
in ~\cite{Diakonov}. These values were recently confirmed
by lattice measurements \cite{lattice}.
In case of the heavy quarks, the instanton vacuum description was
discussed in~\cite{Diakonov:1989un,Chernyshev:1995gj}. For the heavy quarks
even the charmed quark mass $m_{c}\sim1.5$~GeV is larger than the typical parameters
of the instanton media--the inverse instanton size $\rho^{-1}\approx600$~MeV
and the interinstanton distance $R^{-1}\approx200$~MeV and thus the quark mass determines the
dynamics of the heavy quarks.
\section{Light quark determinant with the quark sources term.}
\label{Light quark}
Instanton vacuum field is assumed as a superposition of $N_{+}$
instantons and $N_{-}$ antiinstantons: \begin{eqnarray}
A_{\mu}(x)=\sum_{I}^{N_{+}}A_{\mu}^{I}(\xi_{I},x)+\sum_{A}^{N_{-}}A_{\mu}^{A}(\xi_{A},x).\label{A}\end{eqnarray}
Here $\xi=(\rho,z,U)$ are (anti)instanton collective coordinates--
size, position and color orientation (see reviews ~\cite{Diakonov,Schafer:1996wv}.
The main parameters of the model are the average inter-instanton distance
$R$ and the average instanton size $\rho$. The estimates of these
quantities are \begin{eqnarray}
& & \rho\simeq0.33\, fm,\, R\simeq1\,{\rm fm},\mbox{(phenomenological)}~~\mbox{\cite{Diakonov,Schafer:1996wv}},\nonumber \\
& & \rho\simeq0.35\, fm,\, R\simeq0.95\,{\rm fm},\mbox{(variational)}~~\mbox{\cite{Diakonov}},\nonumber \\
& & \rho\simeq0.36\, fm,\, R\simeq0.89\,{\rm fm},~\mbox{(lattice)}~\mbox{\cite{lattice}}\label{classicalParameters}\end{eqnarray}
and have $\sim10-15\%$ uncertainty.
Our main approximation is the interpolation formula for the light quark propagator in a single instanton field:
\begin{eqnarray}\label{Si}
&&S_{i}=S_{0}+S_{0}\hat{p}\frac{|\Phi_{0i}><\Phi_{0i}|}{c_{i}}\hat{p}S_{0},\\ \nonumber
&&S_{0}=\frac{1}{\hat{p}+im},\,\,\,
c_{i}=im<\Phi_{0i}|\hat{p}S_{0}|\Phi_{0i}>\,.
\end{eqnarray}
The advantage of this interpolation is shown by the projection of
$S_{i}$ to the zero-modes: \begin{eqnarray}
S_{i}|\Phi_{0i}>=\frac{1}{im}|\Phi_{0i}>,\,\,\,<\Phi_{0i}|S_{i}=<\Phi_{0i}|\frac{1}{im}\end{eqnarray}
as it must be, while the similar projection of $S_{i}$ given by
~\cite{Diakonov} has a wrong component, negligible only in
the $m\rightarrow0$ limit.
Summation of the re-scattering series leads to the light quark propagator in the instanton vacuum:
\begin{equation}
S =S_{0} -S_{0}\sum_{i,j}\hat p |\Phi_{0i}>
<\Phi_{0i}|\frac{1}{B}|\Phi_{0j}>
<\Phi_{0j}|\hat p S_{0},
\label{propagator1}
\end{equation}
where $B=\hat pS_0\hat p.$
Here $\tilde {\rm Tr}$ means the trace on the flavor and only on
zero-mode ($|\Phi_{0j}> $) space.
The explicit form of the matrix $ B(m)$ on the flavor and only on
zero-modes ($|\Phi_{0j}> $) space is:
\begin{equation}
B^{fg}_{ij}=\delta_{fg}
<\Phi_{0i}|\hat p \, S_{0,f}
\hat p |\Phi_{0j}> .
\end{equation}
Then the low-frequency part of the light quark determinant~\cite{Musakhanov} is
\begin{equation}
{{\rm Det}}_{\rm low} [m] = {\rm det} B(m).
\label{detB}
\end{equation}
Making few further steps~\cite{Musakhanov} we get
the fermionized representation of low-frequencies
light quark determinant in the presence of the quark sources, which is relevant for our problems, in the
form:
\begin{eqnarray}
&&{\rm Det}_{\rm low}\exp(-\xi^{+}S\xi)=\int\prod_{f}D\psi_{f}D\psi_{f}^{\dagger} \prod_{\pm,f}^{N_{\pm}}V_{\pm,f}[\psi^{\dagger},\psi]
\nonumber\\\label{part-func}
&&\times\exp\int\sum_{f}\left(\psi_{f}^{\dagger}(\hat{p}\,+\, im_{f})\psi_{f}+\psi_{f}^{\dagger}\xi_{f}+\xi_{f}^{+}\psi_{f}\right),
\end{eqnarray}
where
\begin{eqnarray}\label{V}
V_{\pm,f}[\psi^{\dagger},\psi]=&&i\int d^{4}x\left(\psi_{f}^{\dagger}(x)\,\hat{p}\Phi_{\pm,0}(x;\zeta_{\pm})\right)\\\nonumber
&&\times\int d^{4}y\left(\Phi_{\pm,0}^{\dagger}(y;\zeta_{\pm})(\hat{p}\,\psi_{f}(y)\right).
\end{eqnarray}
The light quark partition function $Z[\xi_f,\xi_f^+]$ is given by the averaging of ${\rm Det}_{\rm low}\exp(-\xi^{+}S\xi)$ over the collective coordinates of the instantons $\zeta_\pm$ as:
\begin{eqnarray}\nonumber
Z[\xi_f,\xi_f^+]=\int D\zeta &{\rm Det}_{\rm low}\exp(-\xi^{+}S\xi),\,\,\, D\zeta=\prod_\pm d\zeta_\pm.
\end{eqnarray}
The averaging over collective coordinates $\zeta_{\pm}$ is a rather simple procedure,
since factorized form of the Eq.~(\ref{part-func})
and the low density of the instantons ($\pi^{2}\left(\frac{\rho}{R}\right)^{4}\sim0.1$).
These one allows us to average over positions and orientations of the instantons independently.
Light quark partition function at $N_f=1$ and $N_\pm=N/2$ is exactly given by
\begin{eqnarray}
&&Z[\xi,\xi^+]=\exp\left[{-\xi^+\left(\hat p \,+\, i(m+M(p))\right)^{-1}\xi}\right]\label{Z}\\\nonumber
&&\times\exp\left[{\rm Tr}\ln\frac{\hat p+im+iM(p)}{\hat p+im }+N\ln\frac{N/2}{\lambda}-N\right]
\\
&&N={\rm tr}\frac{iM(p)}{\hat p \,+\, i(m+M(p))},\, M(p)=\frac{\lambda}{N_c}(2\pi\rho F(p))^2.
\label{M}
\end{eqnarray}
Here the form-factor $F(p)$ is given by Fourier-transform of the zero-mode. The coupling $\lambda$ and the dynamical quark mass $M(p)$ are defined by the Eq. (\ref{M}).
At $N_f=2$, $N_\pm=N/2$ and saddle-point approximation (no meson loops contribution)
\begin{eqnarray}
&&Z[\xi_f,\xi_f^+]=\exp\left[-\sum_f\xi_f^+\left(\hat p+im_f+iM_f(p)\right)^{-1}\xi_f\right]
\label{ZNf=2}\\\nonumber
&&\times\exp\left[N\ln\frac{N/2}{\lambda}-N
- \frac{V\sigma^2}{2}+ \sum_f{\rm Tr}\ln\frac{\hat p+im_f+iM_f(p)}{\hat p+im_f }\right].
\end{eqnarray}
Here $\lambda,\sigma$ and dynamical quark mass
$$M(p)=\frac{\lambda^{0.5}}{2g}( 2\pi\rho)^2F^2(p)\sigma,\,\,\,g^2=\frac{(N^2_c-1)2N_c}{2N_c-1}$$
are defined from the Eqs.
\begin{eqnarray}
N=\frac{1}{2}{\rm Tr}\frac{iM_f(p)}{\hat p+im_f+iM_f(p)}=\frac{1}{2}\sigma^2.
\end{eqnarray}
In general, at $N_f >2$, and in the saddle-point approximation (no meson loops contribution) $Z[\xi_f,\xi_f^+]$ has a similar form as the Eqs. (\ref{Z}, \ref{ZNf=2}).
\section{Light quark propagator }
\label{Light quark propagator}
The propagator is defined as
\begin{eqnarray}
S=\int DA \,\,{\rm Det}\left(\hat P+im\right)\frac{1}{\hat P+im},\,\,\, \hat P=\hat p+\hat A.
\end{eqnarray}
In the instanton vacuum model $A\approx\sum_i A_i$, where $A_i$ are instantons and $DA\approx D\zeta .$
Then, accordingly Eq.~(\ref{M}) the light quark propagator is:
\begin{eqnarray}
S=\frac{1}{\hat p \,+\, i(m+M(p))}.
\end{eqnarray}
Pobylitsa~\cite{Pobylitsa:1989uq} neglected by the quark determinant:
\begin{eqnarray}
S_{Pob}=\int D\zeta \frac{1}{\hat P+im}
\end{eqnarray}
and derived the Eq.:
\begin{eqnarray}
S^{-1}_{Pob}= S_0^{-1} +\int D\zeta \sum_i(S_{Pob}-\hat A^{-1}_i)^{-1},
\label{PobEq}
\end{eqnarray}
where it was applied large $N_c$ argumentation. Representing $S^{-1}_{Pob}- S_0^{-1} = \Sigma$,
it was found the Eq.:
\begin{eqnarray}
&&\Sigma
=\frac{N}{2VN_c}{\rm tr_c} \sum_\pm \int dz_\pm \frac{ \hat p|\Phi_{0,\pm}> <\Phi_{0,\pm}| \hat p}{\Sigma_0}+O\left[\left(\frac{N}{VN_c}\right)^2\right],
\nonumber\\
&&\Sigma_0=<\Phi_{0,\pm}|\Sigma|\Phi_{0,\pm}>
\end{eqnarray}
and
finally ($m=0$ case) the solution for the dynamical quark mass:
\begin{eqnarray}\label{MPob}
M^2_{Pob}(k)= \frac{N}{4VN_c}\frac{(2\pi\rho)^4 F^4(k)}{ \int\frac{d^4q}{(2\pi)^4}\frac{(2\pi\rho)^4 F^4(q)}{q^2}}
\end{eqnarray}
corresponding to the Eq.
\begin{eqnarray}\label{MPob1}
4N_c \int\frac{d^4q}{(2\pi)^4}\frac{M^2_{Pob}(q)}{q^2}=\frac{N}{V}
\end{eqnarray}
The Eq. (\ref{M}) for the $\lambda$ ($m=0$ case) has explicit form
\begin{eqnarray}\label{Mdet}
4N_c \int\frac{d^4q}{(2\pi)^4}\frac{M^2(q)}{q^2+M^2(q)}=\frac{N}{V}.
\end{eqnarray}
As we see, the difference between Eqs. (\ref{MPob1}) and (\ref{Mdet}) is only in the denominators.
This one is due to the account of the quark determinant in the derivation of the Eq.~(\ref{Mdet}) (and Eq.~(\ref{M})).
In the following we will use the Eq. (\ref{M}) for the dynamical quark mass $M(k)$.
Any $N_f$ case in the saddle-point approximation has no essential difference with the present case $N_f=1$.
\section{Heavy quark propagator.}
\label{Heavy quark}
At the ref.~\cite{Diakonov:1989un} it was considered the Eq. for the heavy quark propagator in the line similar the Eq.~(\ref{PobEq}).
Our aim here is to extend the approach~\cite{Diakonov:1989un} taking in-to account the light quarks contribution at $N_f=1$ case.
So, define the heavy quark propagator as:
\begin{eqnarray}
&&S_H=\frac{1}{Z}\int D\psi D\psi^{\dagger}
\prod_{\pm}^{N_{\pm}}\bar V_{\pm}[\psi^{\dagger} ,\psi ]\,e^{\int\psi^{\dagger}(\hat p+im )\psi} w[\psi,\psi^\dagger],
\nonumber\\ \nonumber
&&w[\psi,\psi^\dagger]
=\int \frac{D\zeta}{\prod_{\pm}^{N_{\pm}}\bar V_{\pm}[\psi^{\dagger} ,\psi ]}
\prod_{\pm}^{N_{\pm}}V_{\pm}[\psi^{\dagger} ,\psi ]\frac{1}{\theta^{-1}-\sum_i a_i},
\nonumber\\
&& w_\pm=\frac{1}{\theta^{-1}-a_\pm},\,\, <t|\theta|t'>=\theta(t-t'), \\\nonumber
&& <t|\theta^{-1}|t'>=-\frac{d}{dt}\delta(t-t'),
a_i(t)=iA_{i,\mu}(x(t))\frac{d}{dt}x_\mu(t).
\end{eqnarray}
In the $w[\psi,\psi^\dagger]$ the measure of the integration has a factorized form
$ \prod_{\pm}^{N_{\pm}}\frac{d\zeta_\pm}{\bar V_{\pm}[\psi^{\dagger} ,\psi ]}$ as in the Eq.~(\ref{PobEq}). It provide the way for the
extension of this Eq.. Extended Eq. with the account of the light quarks has a form:
\begin{eqnarray}\label{Eqw}
&&w{-1}[\psi,\psi^\dagger]=\\\nonumber
&&= \theta^{-1} +\int \prod_{\pm}^{N_{\pm}}\frac{d\zeta_\pm}{\bar V_{\pm}[\psi^{\dagger} ,\psi ]} \sum_i\left(w[\psi,\psi^\dagger]-\hat A^{-1}_i\right)^{-1}.
\end{eqnarray}
Again, we have the approximate solution of this Eq. as:
\begin{eqnarray}
&&w^{-1}[\psi,\psi^\dagger]-\theta^{-1}=
\\
&&= \frac{N}{2}\sum_\pm \int\frac{ d\zeta_\pm}{\bar V_{\pm}[\psi^{\dagger} ,\psi ]} V_{\pm}[\psi^{\dagger} ,\psi ]\left( \theta-a_\pm^{-1}\right)^{-1}+ O(N^2/V^2)
\nonumber
\\
&&=-\frac{N}{2}\sum_{\pm}\int\frac{ d\zeta_\pm}{\bar V_{\pm}[\psi^{\dagger} ,\psi ]}V_{\pm}[\psi^{\dagger} ,\psi ]\frac{1}{\theta}
(w_\pm-\theta)\frac{1}{\theta}+ O(N^2/V^2)
\nonumber
\\
\nonumber
&&\equiv - \frac{N}{2}\sum_\pm \frac{1}{\bar V_{\pm}[\psi^{\dagger} ,\psi ]}\Delta_{H,\pm}[\psi^{\dagger},\psi ] + O(N^2/V^2)
\end{eqnarray}
and finally we get
\begin{eqnarray}
\label{SH1}
S_H=\left[\frac{1}{\theta^{-1} - \lambda\sum_\pm\Delta_{H,\pm}[\frac{\delta}{\delta\xi} ,\frac{\delta}{\delta\xi^+}] }
e^{-\xi^+\left(\hat p + i(m+M(p))\right)^{-1}\xi}\right]_{|_{\xi=\xi^+=0}}.
\end{eqnarray}
If to neglect by overlapping quark loops, then
\begin{eqnarray}
&&S_H^{-1}\approx\left[\left(\theta^{-1} - \lambda\sum_\pm\Delta_{H,\pm}[\frac{\delta}{\delta\xi} ,\frac{\delta}{\delta\xi^+}] \right)
e^{\xi^+\left(\hat p + i(m+M(p))\right)^{-1}\xi}\right]_{|_{\xi=\xi^+=0}}
\nonumber\\
&&=\theta^{-1} - \frac{N}{2VN_c}\sum_\pm\int d^4z_\pm {\rm tr}_c\left(\theta^{-1}(w_\pm-\theta)\theta^{-1}\right).
\label{SH3}
\end{eqnarray}
The Eq. (\ref{SH3}) exactly coincide with the similar one from~\cite{Diakonov:1989un}.
Now re-write the Eq. (\ref{SH1}) introducing heavy quark fiels $Q,Q^\dagger$:
\begin{eqnarray}
&& S_H=e^{\left[-{\rm tr}\ln\left(\hat p \,+\, i(m+M(p))\right)\right]}\int D\psi D\psi^{\dagger} D Q D Q^\dagger \,\,Q \, Q^\dagger\\\nonumber
&&\times\exp\left[\psi^{\dagger}(\hat p +i(m+M(p)))\psi+ Q^\dagger\left(\theta^{-1} - \lambda\sum_\pm\Delta_{H,\pm}[\psi^{\dagger},\psi ] \right)Q\right]
\\\nonumber
&&\times\exp\left[-{\rm tr}\ln\left(\theta^{-1} - \lambda\sum_\pm\Delta_{H,\pm}[\psi^{\dagger},\psi ]\right)\right],
\end{eqnarray}
where last exponent represent the (negligible) contribution of the heavy quark loops, while
the second one has the heavy and light quarks interaction action, explicitly represented by
\begin{eqnarray}
&& - \lambda\sum_\pm Q^\dagger\Delta_{H,\pm}[\psi^{\dagger},\psi ]Q=
\nonumber\\
&&= - i\lambda\sum_\pm\int d^4z_\pm \frac{d^4 k_1}{(2\pi)^4} \frac{d^4 k_2}{(2\pi)^4} e^{(i(k_2-k_1)z_\pm)} (2\pi\rho )^2 F(k_1 )F(k_2 )
\nonumber \\ \nonumber
&&\times\left[ \frac{1}{N_c^2}\psi^+(k_1)\frac{1\pm\gamma_5}{2}\psi(k_2)Q^+ {\rm tr}_c\left(\theta^{-1}(w_\pm-\theta)\theta^{-1}\right)Q\right.
\\\nonumber
&&\left.+\frac{1}{32(N_c^2-1)}\psi^+(k_1) (\gamma_\mu\gamma_\nu \frac{1\pm\gamma_5}{2})\lambda^i \psi(k_2){\rm tr}(\tau^{\mp}_{\mu}\tau^{\pm}_{\nu}\lambda^j)\right.
\\
&&\times\left. Q^+ {\rm tr}_c\left(\theta^{-1}(w_\pm-\theta)\theta^{-1}\lambda^j \right)\lambda^i Q\right].
\end{eqnarray}
We see that the heavy-light quarks interactions terms has a form of the product of the colorless currents of a heavy and light quarks together with similar term of the colorful currents product. The structure of these currents are defined by the instanton color orientation integration,
while the instanton position integration provide energy-momentum conservation in the interaction vertex.
At the $N_f>1$ case we have an interaction vertex with $N_f$ pairs of a light quark legs and the pair of a heavy quark legs. The specific structure of the interaction is defined again by instanton color orientation and will be much more reach then at $N_f=1$ case. We expect that the action generated by the instantons will have reach symmetry properties related to light and heavy quarks sectors both. Namely, it appear the light-heavy quarks interaction terms leading to the specific traces of the light quarks chiral symmetry in light-heavy quarks systems.
\section{Heavy quark anti-quark system.}
\label{Heavy quark antiquark}
Now it is considered the correlator for this system again with the account of ($N_f=1$ case) light quark contribution:
\begin{eqnarray}
&&<T|C(L_1,L_2)|0>=\frac{1}{Z}\int D\psi D\psi^{\dagger}
\left\{\prod_{\pm}^{N_{\pm}}\bar V_{\pm}[\psi^{\dagger} ,\psi ]\right\}\\ \nonumber
&&\times\exp\int\left(\psi^{\dagger}(\hat p+im )\psi\right)<T|W[\psi,\psi^\dagger]|0>,
\\ \nonumber
&&<T|W[\psi,\psi^\dagger]|0>
=\int\frac{ D\zeta}{\left\{\prod_{\pm}^{N_{\pm}}\bar V_{\pm}[\psi^{\dagger} ,\psi ]\right\}}
\left\{\prod_{\pm}^{N_{\pm}}V_{\pm}[\psi^{\dagger} ,\psi ]\right\}
\\ \nonumber
&&{\rm Tr}<T|\left(\theta^{-1}-\sum_i a^{(1)}_i\right)^{-1}|0>
<0|\left(\theta^{-1}-\sum_i a^{(2)}_i\right)^{-1}|T>.
\end{eqnarray}
Here the correlator is a Wilson loop along the rectangular contour $L\times r$, where the sides $L_1,L_2$ are parallel to $x_4$ axes and
separated by the distance $r$. The $a^{(1)},a^{(2)}$ are the projections of the instantons onto the lines $L_1,L_2.$
In the ref.~\cite{Diakonov:1989un} this correlator was considered within the approach similar to the Eq.~(\ref{PobEq}) of the ref.~\cite{Pobylitsa:1989uq} but without a light quarks.
The argumentation, which provided the derivation of the Eq~(\ref{Eqw}), is applicable to the present case and leads to the similar Eq..
\begin{eqnarray}
&&W^{-1}[\psi,\psi^\dagger]=
\\\nonumber
&&= w_1^{-1}[\psi,\psi^\dagger]\otimes w_2^{-1,T}[\psi,\psi^\dagger]
-\frac{N}{2}\sum_\pm\int\frac{ d\zeta_\pm}{ \bar V_{\pm}[\psi^{\dagger} ,\psi ]}
\\ \nonumber
&&\times V_{\pm}[\psi^{\dagger} ,\psi ] \left(w_1[\psi,\psi^\dagger]-a^{(1)-1}_\pm\right)^{-1}\otimes\left(w_2[\psi,\psi^\dagger]-a^{(2)-1}_\pm\right)^{-1,T} ,
\end{eqnarray}
where, superscript $T$ means the transposition, $\otimes$ -- tensor product.
This Eq. has an approximate solution:
\begin{eqnarray}
&&W^{-1}[\psi,\psi^\dagger]= w_1^{-1}[\psi,\psi^\dagger]\otimes w_2^{-1,T}[\psi,\psi^\dagger]
\\ \nonumber
&&
-\frac{N}{2}\sum_\pm\int\frac{ d\zeta_\pm}{ \bar V_{\pm}[\psi^{\dagger} ,\psi ]}
V_{\pm}[\psi^{\dagger} ,\psi ]
\\ \nonumber
&&\times\left(\theta^{-1}\left(w^{(1)}_\pm-\theta\right)\theta^{-1}\right)
\otimes\left(\theta^{-1}\left(w^{(2)}_\pm-\theta\right)\theta^{-1}\right)^{T}+ O(N^2/V^2).
\end{eqnarray}
and
\begin{eqnarray}
&&w_1^{-1}[\psi,\psi^\dagger]=\theta^{-1}-
\\\nonumber
&&-\frac{N}{2}\sum_{\pm}\frac{ d\zeta_\pm}{ \bar V_{\pm}[\psi^{\dagger} ,\psi ]} V_{\pm}[\psi^{\dagger} ,\psi ]\theta^{-1}(w^{(1)}_\pm-\theta)\theta^{-1}+ O(N^2/V^2)
\\\nonumber
&&=\theta^{-1} - \frac{N}{2}\sum_\pm \frac{1}{\bar V_{\pm}[\psi^{\dagger} ,\psi ]}\Delta^{(1)}_{H,\pm}[\psi^{\dagger},\psi ] + O(N^2/V^2)
\end{eqnarray}
and similar for the $w_2^{-1}[\psi,\psi^\dagger].$
From previous calculations we see that the lowest orders on $\frac{N}{N_cV}$ in $C(L_1,L_2)$ are given by the integration over $\psi,\psi^\dagger$ of the $W^{-1}[\psi,\psi^\dagger]. $ Here it was neglected by overlapping quark loops.
Then, we have the new interaction term between heavy quarks located on the lines $L_1$ and $L_2$ due to exchange of
the light quarks between them.
Explicitly the integration of the first term in $W^{-1}[\psi,\psi^\dagger]$ over $\psi,\psi^\dagger$ leads to:
\begin{eqnarray}
&&\frac{1}{Z}\int D\psi D\psi^{\dagger}
\left\{\prod_{\pm}^{N_{\pm}}\bar V_{\pm}[\psi^{\dagger} ,\psi ]\right\}\exp\int\psi^{\dagger}(\hat p+im )\psi
\\\nonumber
&&\times\, w_1^{-1}[\psi,\psi^\dagger]\otimes w_2^{-1,T}[\psi,\psi^\dagger]
=\left(\theta^{-1}-\lambda\sum_\pm\Delta^{(1)}_{H,\pm}[\frac{\delta}{\delta\xi} ,\frac{\delta}{\delta\xi^+}]\right)
\\\nonumber
&& \otimes
\left(\theta^{-1}-\lambda\sum_\pm\Delta^{(2)}_{H,\pm}[\frac{\delta}{\delta\xi} ,\frac{\delta}{\delta\xi^+}]\right)^{T} {e^{-\xi^+\left(\hat p \,+\, i(m+M(p))\right)^{-1}\xi}}_{|_{\xi=\xi^+=0}}.
\end{eqnarray}
Light quarks generated potential is given by
\begin{eqnarray}
&&V_{lq}=\left(\lambda\sum_\pm\Delta^{(1)}_{H,\pm}[\frac{\delta}{\delta\xi_1} ,\frac{\delta}{\delta\xi_1^+}]\right) \otimes
\left(\lambda\sum_\pm\Delta^{(2)}_{H,\pm}[\frac{\delta}{\delta\xi_2} ,\frac{\delta}{\delta\xi_2^+}]\right)^{T}
\nonumber\\
&&\times {e^{\left[-\xi_2^+\left(\hat p \,+\, i(m+M(p))\right)^{-1}\xi_1-\xi_1^+\left(\hat p \,+\, i(m+M(p))\right)^{-1}\xi_2\right]}}|_{\xi=\xi^+=0}.
\end{eqnarray}
The range of this potential is controlled by dynamical light quark mass $M\sim 350$~MeV and might be important for the heavy quarkonium states properties.
\section{ Conclusion.}
\label{ Conclusion}
Approximating the gluon field by the instanton configurations
it was derived the low-frequency part of the light quark determinant in
the presence of quark sources.
It was provided the calculation of the instanon generated light-heavy quarks interaction terms and
the heavy quark propagator with the account of the light
quark determinant together with the QCD instanton vacuum properties at the $N_f=1$ case.
With these knowledge it was calculated the light quark contribution to
the interaction between heavy quarks.
The extension of this approach to $N_f>1$ case is obvious and provide the possibility for the detailed investigation
of the role of the light quarks chiral symmetry and its spontaneous breaking for the heavy and heavy-light quarks systems.
The estimations of the light quark contributions to their properties are on the way.
|
\section{Introduction}
\newcommand{(\rho_i)_{i\in I}}{(\rho_i)_{i\in I}}
\newcommand{{\ell\in{\mathbb{L}}}}{{\ell\in{\mathbb{L}}}}
Let $\Gamma$ be a profinite group and $(\Gamma_i)_{i\in I}$ a family of
groups. For every $i$ let $\rho_i\colon \Gamma\to \Gamma_i$ be a
homomorphism. Following Serre (cf. \cite[p. 1]{bible}),
we shall say that the family $(\rho_i)_{i\in I}$ is
{\em independent}, provided the homomorphism
$$\Gamma\buildrel \rho \over\longrightarrow \prod_{i\in I} \rho_i(\Gamma)$$
induced by the $\rho_i$ is surjective.
Let $\Gamma'\subset \Gamma$ be a closed subgroup. We
call the family $(\rho_i)_{i\in I}$ {\em independent over $\Gamma'$}, if
$\rho(\Gamma')=\prod_{i\in I}\rho_i(\Gamma')$. Finally we call the
family $(\rho_i)_{i\in I}$ {\em almost independent}, if there exists an open subgroup
$\Gamma'\subset \Gamma$, such that $(\rho_i)_{i\in I}$ is independent over $\Gamma'$.
Of particular interest is the special case where $\Gamma={\mathrm{Gal}}_K$ is the absolute Galois
group of a field $K$, and $(\rho_\ell)_{\ell\in{\mathbb{L}}}$ is a family of $\ell$-adic
representations of ${\mathrm{Gal}}_K$, indexed by the set ${\mathbb{L}}$ of all prime numbers.
Important examples of such families of representations arise as follows:
Let $K$ be a field of characteristic zero and let $X/K$ be a separated $K$-scheme of finite type.
Denote by $\widetilde{K}$ an algebraic closure of $K$. For every $\ell\in{\mathbb{L}}$ and every $q\ge 0$ we consider
the representation of the absolute Galois group ${\mathrm{Gal}}(\widetilde{K}/K)$
$$\begin{xy}
\xymatrix{
\rho_{\ell, X}^{(q)}\colon {\mathrm{Gal}}(\widetilde{K}/K)\ar[r] &
\mathrm{Aut}_{{\mathbb{Q}}_\ell}(\mathrm{H}^q(X_{\widetilde{K}}, {\mathbb{Q}}_\ell))
}
\end{xy}$$
afforded by the \'etale cohomology group $\mathrm{H}^q(X_{\widetilde{K}}, {\mathbb{Q}}_\ell)$, and also the
representation
$$\begin{xy}
\xymatrix{
\rho_{\ell, X, c}^{(q)}\colon {\mathrm{Gal}}(\widetilde{K}/K)\ar[r] &
\mathrm{Aut}_{{\mathbb{Q}}_\ell}(\mathrm{H}^q_\mathrm{c}(X_{\widetilde{K}}, {\mathbb{Q}}_\ell))
}
\end{xy}$$
afforded by the \'etale cohomology group with compact support $\mathrm{H}^q_\mathrm{c}(X_{\widetilde{K}}, {\mathbb{Q}}_\ell)$.
One can wonder in which circumstances the families $(\rho_{\ell, X}^{(q)})_{\ell\in{\mathbb{L}}}$
and $(\rho_{\ell, X, c}^{(q)})_{\ell\in{\mathbb{L}}}$ are almost independent.
In the recent paper \cite{bible} Serre considered the special case where $K$ is a {number field}. He
proved a general independence criterion for certain families of $\ell$-adic representations over a number field
(cf. \cite[Section 2, Th\'eor\`em 1]{bible}),
and used this criterion together with results of Katz-Laumon and of Berthelot (cf. \cite{illusie})
in order to prove the following
Theorem (cf. \cite[Section 3]{bible}).
{\em Let $K$ be a number field and $X/K$ a
separated scheme of finite type. Then the families of representations $(\rho_{\ell, X}^{(q)})_{\ell\in{\mathbb{L}}}$
and $(\rho_{\ell, X, c}^{(q)})_{\ell\in{\mathbb{L}}}$ are almost independent.}
The special case of an abelian variety $X$ over a number field $K$ had been dealt with earlier in a letter from Serre to Ribet (cf. \cite{serretoribet2}).
In \cite[p. 4]{bible} Serre asks the following question.
{\em Does this theorem remain true, if one replaces the number field $K$ by a
finitely generated transcendental extension $K$ of ${\mathbb{Q}}$?}
This kind of problem also shows up in Serre's article \cite[10.1]{serre1994} and in
Illusie's manuscript \cite{illusie}.
The aim of our paper is to answer this question affirmatively.
In order to do this we prove an independence criterion for families of
$\ell$-adic representations of the \'etale fundamental group $\pi_1(S)$ of a normal ${\mathbb{Q}}$-variety $S$
(cf. Theorem \ref{crit} below). This criterion allows us to reduce the proof of the following
Theorem \ref{main2} to the
number field case, where it is known to hold true thanks to the theorem
of Serre (cf. \cite{bible}) mentioned above. We do take Tate twists into account.
For every $\ell\in {\mathbb{L}}$ we
denote by $\varepsilon_\ell\colon {\mathrm{Gal}}_K\to \mathrm{Aut}_{{\mathbb{Q}}_\ell}((\mathop{\varprojlim}\limits_{i\in {\mathbb{N}}} \mu_{\ell^i})\otimes {\mathbb{Q}}_\ell)\subset{\mathbb{Q}}_\ell^\times$
the cyclotomic character, by $\varepsilon_\ell^{\otimes-1}$ its contragredient and define for every $d\in {\mathbb{Z}}$
$$\rho^{(q)}_{\ell, X}(d):=\rho^{(q)}_{\ell, X}\otimes \varepsilon_\ell^{\otimes d}\ \ \mbox{and}\ \
\rho^{(q)}_{\ell, X, \mathrm{c}}(d):=\rho^{(q)}_{\ell, X, \mathrm{c}}\otimes \varepsilon_\ell^{\otimes d}.$$
\begin{thm} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$. Let $X/K$ be a separated scheme of
finite type. Then for every $q\in{\mathbb{N}}$ and every $d\in{\mathbb{Z}}$
the families $(\rho^{(q)}_{\ell, X}(d))_{\ell\in{\mathbb{L}}}$ and
$(\rho^{(q)}_{\ell, X, c}(d))_{\ell\in{\mathbb{L}}}$ of representations
of ${\mathrm{Gal}}_K$ are almost independent. \label{main2}
\end{thm}
Note that outside certain special cases it is not known whether
the representations occuring in Theorem \ref{main2} are semisimple.
Hence we cannot
use techniques like the semisimple approximation
of monodromy groups in the proof of Theorem \ref{main2}.
Theorem \ref{main2} has an important consequence for the arithmetic of abelian varieties.
Let $A/K$ be an abelian variety.
For every $\ell\in{\mathbb{L}}$ consider the Tate module $T_\ell(A):=\mathop{\varprojlim}\limits_{i} A(\widetilde{K})[\ell^i]$,
define $V_\ell(A):=T_\ell(A)\otimes_{{\mathbb{Z}}_\ell} {\mathbb{Q}}_\ell$ and let
$$\begin{xy}
\xymatrix{
\eta_{\ell, A}\colon {\mathrm{Gal}}(\widetilde{K}/K)\ar[r] & \mathrm{Aut}_{{\mathbb{Q}}_\ell}(V_\ell(A))
}
\end{xy}$$
be the $\ell$-adic representation attached to $A$. Then the ${\mathbb{Q}}_\ell[{\mathrm{Gal}}_K]$-modules $V_\ell(A)$ and
$\mathrm{H}^1(A^\vee_{\widetilde{K}}, {\mathbb{Q}}_\ell(1))$ are isomorphic, i.e. the representation
$\eta_{\ell, A}$ is isomorphic to $\rho_{\ell, A^\vee}(1)$.
Hence Theorem \ref{main2} implies that
the family $(\eta_{\ell, A})_{\ell\in{\mathbb{L}}}$ is almost independent. Denote by
$K(A[\ell^\infty])$ the fixed field in $\widetilde{K}$ of the kernel of $\eta_{\ell, A}$. Then
$K(A[\ell^\infty])$ is the field obtained from $K$ by adjoining the coordinates of the
$\ell$-power division points in $A(\widetilde{K})$. Using Remark \ref{indeprem} below we
see that Theorem \ref{main2} has the following Corollary.
\begin{coro} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$ and $A/K$ an abelian variety.
Then there is a finite extension $E/K$ such that
the family $(EK(A[\ell^\infty]))_{\ell\in{\mathbb{L}}}$ is linearly disjoint over $E$.\label{main1}
\end{coro}
This paper carries an appendix with a more elementary proof of this Corollary, which is based on
our Theorem \ref{crit} below, but avoiding use of \'etale cohomology.
\begin{center} {\bf Notation and Preliminaries} \end{center}
For a field $K$ fix an algebraic closure $\widetilde{K}$
and denote by ${\mathrm{Gal}}_K$ the absolute
Galois group of $K$.
We denote by ${\mathbb{L}}$ the set of all prime numbers.
Let $S$ be a scheme and $s\in S$ a point (in the underlying topological space). Then
$k(s)$ denotes the residue field at $s$. A {\em geometric point} of $S$ is a
morphism $\ol{s}\colon \mathrm{Spec}(\Omega)\to S$ where $\Omega$ is an algebraically closed field. To
give such a geometric point $\ol{s}$ is equivalent to giving a pair $(s, i)$ consisting
of a usual point $s\in S$ and an embedding $i\colon k(s)\to \Omega$. We then let $k(\ol{s})$ be
the algebraic closure of $i(k(s))$ in $\Omega$. Now assume $S$ is an integral scheme and
let $K$ be its function field. Then we view $S$ as equipped with the geometric
generic point $\mathrm{Spec}(\widetilde{K})\to S$ and denote by $\pi_1(S)$ the \'etale fundamental
group of $S$ with respect to this geometric point.
A {\em variety} $S$ over a field $F$ is an integral separated
$F$-scheme of finite type.
Now let $S$ be a connected normal scheme with function field $K$. Assume for
simplicity that $\mathrm{char}(K)=0$.
If $E/K$ is a an algebraic field extension, then $S^{(E)}$
denotes the normalization of $S$ in $E$ (cf. \cite[6.3]{EGAII}). This notation is used throughout this
manuscript. The canonical morphism $S^{(E)}\to S$ is universally closed and
surjective. (This follows from the going-up theorem, cf. \cite[6.1.10]{EGAII}.)
If $E/K$ is a finite extension,
then $S^{(E)}\to S$ is a finite morphism (cf. \cite[Proposition I.1.1]{milne}).
We shall say that an algebraic extension $E/K$ is {\em unramified along $S$}, provided
the morphism $S^{(E')}\to S$ is
\'etale for every {\em finite} extension $E'/K$ contained in $E$. We denote by
$K_{S, \mathrm{nr}}$ the maximal extension of $K$ inside $\widetilde{K}$ which is unramified
along $S$, and by $S_{\mathrm{nr}}$ the normalization of $S$ in $K_{S, \mathrm{nr}}$. One can then
identify $\pi_1(S)$ with ${\mathrm{Gal}}(K_{S, \mathrm{nr}}/K)$.
Let $E/K$ be a Galois extension.
If $P\in S$ is a closed point and $\hat{P}$ is a point in $S^{(E)}$ above $P$, then
we define $D_{E/K}(\hat{P})\subset {\mathrm{Gal}}(E/K)$ to be the decomposition group
of $\hat{P}$, i.e. the stabilizer of $\hat{P}$ under the action of ${\mathrm{Gal}}(E/K)$.
Then $k(\hat{P})/k(P)$ is Galois and the restriction map
$$\begin{xy}
\xymatrix{
r_{E/K, \hat{P}}\colon D_{E/K}(\hat{P})\ar[r] & {\mathrm{Gal}}(k(\hat{P})/k(P))
}
\end{xy}$$
is an epimorphism. To see this apply \cite[Proposition 1.1, p. 106]{SGA1} for the case $[E:K]<\infty$ and use
a limit argument. If $E/K$ is unramified along $S$, then $r_{E/K, \hat{P}}$ is bijective.
Let ${\cal G}/S$ be a finite \'etale group scheme and choose a closed
point $\hat{P}\in S_{\mathrm{nr}}$. Then there is a finite
extension $E/K$ in $K_{S, \mathrm{nr}}$ such that ${\cal G}\times_S S^{(E)}$ is
a constant group scheme over $S^{(E)}$. In particular, the action of
${\mathrm{Gal}}_K$ on ${\cal G}(\widetilde{K})$ factors through ${\mathrm{Gal}}(K_{S, \mathrm{nr}}/K)$, and the canonical
evaluation maps
$$\begin{xy}
\xymatrix{
{\cal G}(S^{(E)})\ar[r] & {\cal G}(E)={\cal G}(\widetilde{K}) &\mbox{and}& {\cal G}(S^{(E)})\ar[r] & {\cal G}(k(\hat{P}))
}
\end{xy}$$
are bijective. The composite isomorphism
$$sp_{{\cal G}, \hat{P}}\colon {\cal G}(\widetilde{K})\cong {\cal G}(k(\hat{P}))$$
is called the {\em cospecialization map}. This map is equivariant in the sense that
$sp_{{\cal G}, \hat{P}}(\sigma (x))=r_{E/K, P}(\sigma)(sp_{{\cal G}, \hat{P}}(x))$ for all
$x\in{\cal G}(E)$ and all $\sigma\in D_{E/K}(\hat{P})$.
\label{preliminary}
\section{Finiteness properties of Jordan extensions}
Let $E/K$ be an algebraic field extension and $d\in {\mathbb{N}}$. We will say that
$E/K$ is {\em $d$-flat}, if $E$ is a compositum of (finitely or infinitely many)
Galois extensions of $K$, each of degree $\le d$. In particular every $d$-flat extension is Galois.
We call the extension $E/K$
{\em $d$-Jordanian}, if $E/K$ is a (possibly infinite) abelian extension of a $d$-flat extension.
The $1$-Jordanian extensions of $K$ are hence just the abelian extensions of $K$.
If $K$ is a number field and $E/K$ is a $d$-Jordanian extension of $K$ which is
everywhere unramified, then $E/K$ is finite. This has been shown by Serre in \cite[Th\'eor\`eme 2]{bible},
making use of the Hermite-Minkowski theorem and the finiteness of the Hilbert class field.
The aim of this section is to derive a similar finiteness property for $d$-Jordanian
extensions of function fields over ${\mathbb{Q}}$. In Lemmata \ref{mwlemm}, \ref{curves} and \ref{katzlanglemm}
we follow closely the paper \cite{katzlang} of Katz and Lang on geometric class field theory, giving
complete details for the convenience of the reader.
If $E$ is any extension field of ${\mathbb{Q}}$, then we denote by $\kappa_E$ the
algebraic closure of ${\mathbb{Q}}$ in $E$,
$$\kappa_E:=\{x\in E\,: x\ \mbox{is algebraic over}\ {\mathbb{Q}}\},$$
and we call $\kappa_E$ the {\em constant field} of $E$.
\begin{rema} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$.
Let $E/K$ be an algebraic extension. Then there is a diagram of fields:
$$\begin{xy}
\xymatrix{
& E\ar@{-}[r]\ar@{-}[d] & \widetilde{{\mathbb{Q}}}E \ar@{-}[d]\\
\kappa_E\ar@{-}[r]\ar@{-}[d] & \kappa_EK \ar@{-}[r] \ar@{-}[d] & \widetilde{{\mathbb{Q}}}K \\
\kappa_K\ar@{-}[r] & K &
}
\end{xy}$$
The field $\kappa_K$ is a number field and $\kappa_E/\kappa_K$
is an algebraic extension. We say that $E/K$ is a {\em constant field extension}, if
$\kappa_E K=E$. If $E/K$ is Galois, then $\kappa_E/\kappa_K$,
$\kappa_EK/K$ and $\widetilde{{\mathbb{Q}}}E/\widetilde{{\mathbb{Q}}}K$ are Galois as well, and
the restriction maps ${\mathrm{Gal}}(\widetilde{{\mathbb{Q}}}E/\widetilde{{\mathbb{Q}}}K)\to {\mathrm{Gal}}(E/\kappa_E K)$ and
${\mathrm{Gal}}(\kappa_E K/K)\to {\mathrm{Gal}}(\kappa_E/\kappa_K)$ are both bijective.\label{elrema}
\end{rema}
The aim of this section is to prove the following Proposition.
\begin{prop} Let $S/{\mathbb{Q}}$ be a normal variety with function field $K$. Let $d\in {\mathbb{N}}$.
Let $E/K$ be a $d$-Jordanian
extension which is unramified along $S$. Then $E/\kappa_E K$ is a finite extension. \label{jordan}
\end{prop}
Note that in the situation of Proposition \ref{jordan} the extension $\kappa_E/\kappa_K$
may well be infinite algebraic.
The proof occupies the rest of this section.
\begin{lemm} Let $S/{\mathbb{Q}}$ be a normal variety with function field $K$.
Let $d\in {\mathbb{N}}$. Let $E/K$ be a $d$-flat
extension which is unramified along $S$. Then $E/\kappa_E K$ is finite and
${\mathrm{Gal}}(\kappa_E/\kappa_K)$ is a (possibly infinite) group of exponent $\le d$. \label{fin1}
\end{lemm}
{\em Proof.} There is a sequence $(K_i)_{i\in I}$ of intermediate fields of $E/K$
such that each $K_i/K$ is Galois with $[K_i:K]\le d$ and $E=\prod_{i\in I} K_i$.
Hence ${\mathrm{Gal}}(E/K)$ is a closed subgroup of $\prod_{i\in I} {\mathrm{Gal}}(K_i/K)$. By Remark \ref{elrema}
${\mathrm{Gal}}(\kappa_E/\kappa_K)$ is a quotient of ${\mathrm{Gal}}(E/K)$, hence
${\mathrm{Gal}}(\kappa_E/\kappa_K)$ has exponent $\le d$.
Again by Remark \ref{elrema} it is now enough to show that $\widetilde{{\mathbb{Q}}}E/\widetilde{{\mathbb{Q}}}K$
is finite. The Galois group ${\mathrm{Gal}}(\widetilde{{\mathbb{Q}}}E/\widetilde{{\mathbb{Q}}}K)$
is a quotient of $\pi_1(S_{\widetilde{{\mathbb{Q}}}})$, and
$\pi_1(S_{\widetilde{{\mathbb{Q}}}})$ is topologically finitely generated (cf. \cite[II.2.3.1]{SGA7}). Hence there
are only finitely many intermediate fields $L$ of $\widetilde{{\mathbb{Q}}}E/\widetilde{{\mathbb{Q}}}K$ with
$[L:\widetilde{{\mathbb{Q}}}K]\le d$ (cf. \cite[16.10.2]{friedjarden}). This implies that $\widetilde{{\mathbb{Q}}}E/\widetilde{{\mathbb{Q}}}K$
is finite.\hfill $\Box$
\begin{lemm} \label{exprime} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$.
Let $E/K$ be a (possibly infinite) Galois extension. Assume that ${\mathrm{Gal}}(E/K)$ has
finite exponent. Let $X=(X_1,\cdots, X_n)$ be a transcendence base
of $K/{\mathbb{Q}}$ and $R$ the integral closure of ${\mathbb{Z}}[X]$ in $E$.
\begin{enumerate}
\item[a)] The residue field $k(\mathfrak m)=R/\mathfrak m$ is {\em finite} for
every maximal ideal $\mathfrak m$ of $R$.
\item[b)] For every non-zero element $f\in R$ there exist two
maximal ideals $\mathfrak m_1$ and $\mathfrak m_2$ of $R$ such that
$f\notin \mathfrak m_1$ and $f\notin \mathfrak m_2$ and $\mathrm{char}(k(\mathfrak m_1))\neq
\mathrm{char}(k(\mathfrak m_2))$.
\end{enumerate}
\end{lemm}
{\em Proof.} Let $R'$ be the integral closure of ${\mathbb{Z}}[X]$ in $K$.
Let $\mathfrak m$ be a maximal ideal of $R$. Define $\mathfrak m':=\mathfrak m\cap R'$ and
$\mathfrak p=\mathfrak m\cap {\mathbb{Z}}[X]$. There are diagrams of fields and residue fields
$$\begin{xy}
\xymatrix{
{\mathbb{Q}}(X)\ar@{-}[r] & K \ar@{-}[r] & E & \mbox{and}& k(\mathfrak p)\ar@{-}[r] & k(\mathfrak m') \ar@{-}[r] & k(\mathfrak m).
}
\end{xy}$$
By the going-up theorem $\mathfrak p$ is a
maximal ideal of ${\mathbb{Z}}[X]$, and $k(\mathfrak p)={\mathbb{Z}}[X]/\mathfrak p$ is a {\em finite}
field. Furthermore $R'$ is a finitely generated ${\mathbb{Z}}[X]$-module
(cf. \cite[Prop. I.1.1]{milne}). This implies that
$k(\mathfrak m')$ is a finite field. The extension $k(\mathfrak m)/k(\mathfrak m')$ is
Galois and the Galois group $G:={\mathrm{Gal}}(k(\mathfrak m)/k(\mathfrak m'))$ is a subquotient of
${\mathrm{Gal}}(E/K)$. Hence $G$ is of finite exponent. On the other hand $G$ must be
procyclic, because it is a quotient of the Galois group $\hat{{\mathbb{Z}}}$ of the
finite field $k(\mathfrak m')$. It follows that $G$ is finite and that
$k(\mathfrak m)$ is a finite field. This finishes the proof of part a).
Now let $f\in R$ be a nonzero element. The canonical
morphism $p: \mathrm{Spec}(R)\to\mathrm{Spec}({\mathbb{Z}}[X])$ is closed (cf. \cite[6.1.10]{EGAII}),
hence $p(V(f))$ is a closed subset of $\mathrm{Spec}({\mathbb{Z}}[X])$. It is also a proper subset
of $\mathrm{Spec}({\mathbb{Z}}[X])$.
It follows that there is a non-zero polynomial $g\in {\mathbb{Z}}[X]$ such that
$D(g)\cap p(V(f))=\emptyset$. Choose $a\in
{\mathbb{Z}}^n$ with $g(a)\neq 0$. Then choose distinct prime numbers $p_1\neq p_2$
not dividing $g(a)$. For $i\in\{1, 2\}$ consider
the maximal ideal $\mathfrak p_i=(p_i, X-a_1,\cdots, X-a_n)$ of ${\mathbb{Z}}[X]$.
Then $\mathfrak p_1, \mathfrak p_2\in D(g)$. Finally let
$\mathfrak m_1$ and $\mathfrak m_2$ be prime ideals of $R$ such that $p(\mathfrak m_i)=\mathfrak p_i$ for
$i\in\{1, 2\}$. Then $\mathfrak m_1$ and $\mathfrak m_2$ have the desired
properties.\hfill $\Box$
We now show that a weak form of the Mordell-Weil theorem holds true over finitely generated
extensions of fields like the field $\kappa_E$ occuring in Lemma \ref{fin1}. If $B$ is a semiabelian variety
over a field $K$, then we define $T(B)=\prod_{\ell\in {\mathbb{L}}} T_\ell(B)$
and $T(B)_{\neq p}:=\prod_{\ell\in {\mathbb{L}}\setminus\{p\}} T_\ell(B)$ (for $p\in{\mathbb{L}}$),
where $T_\ell(B)=\mathop{\varprojlim}\limits_{i\in{\mathbb{N}}} B(\widetilde{K})[\ell^i]$ is the Tate module of $B$ for every $\ell\in {\mathbb{L}}$.
If $M$ is a compact topological ${\mathrm{Gal}}_K$-module, then we define the {\em module of coinvariants}
$M_{{\mathrm{Gal}}_K}$ of $M$ to be the largest Hausdorff quotient of $M$ on which ${\mathrm{Gal}}_K$ acts trivially.
\begin{lemm} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$. Let
$E/K$ be a Galois extension. Assume that ${\mathrm{Gal}}(E/K)$ has
finite exponent. Let $B/E$ be a semiabelian variety. Then
$T(B)_{{\mathrm{Gal}}_E}$ is finite. \label{mwlemm}
\end{lemm}
{\em Proof.} Let $E'/E$ be a finite extension over which the torus part of $B$
splits. Then there exists a finite
Galois extension $L/K$ such that $LE\supset E'$, and ${\mathrm{Gal}}(LE/K)$ has
finite exponent again. The group $T(B)_{{\mathrm{Gal}}_{E}}$ is a quotient of $T(B)_{{\mathrm{Gal}}_{LE}}$.
Hence we may assume right from
the beginning that $B$ is an extension of an abelian variety $A$ by
a split torus ${\mathbb{G}}_{m, E}^d$.
Then there is an exact sequence of ${\mathrm{Gal}}_E$-modules
$$\begin{xy}
\xymatrix{
0\ar[r] & T({\mathbb{G}}_m)^d\ar[r] & T(B)\ar[r] & T(A)\ar[r] & 0
}
\end{xy}$$
As the functor $-_{{\mathrm{Gal}}_E}$
is right exact, it is enough to prove that $T(A)_{{\mathrm{Gal}}_E}$ and
$T({\mathbb{G}}_m)_{{\mathrm{Gal}}_E}$ are both finite. We may thus assume
that either $B$ is an abelian variety over $E$ (case 1) or
$B={\mathbb{G}}_{m, E}$ (case 2). We shall prove the finiteness of
$T(B)_{{\mathrm{Gal}}_E}$ in both cases.
Choose a transcendence base $X=(X_1,\cdots, X_n)$ of $K/{\mathbb{Q}}$ and let
$R$ be the integral closure of ${\mathbb{Z}}[X]$ in
$E$. In case 1 there is a nonempty open subscheme $U\subset \mathrm{Spec}(R)$
such that $B$ extends to an abelian scheme ${\cal B}$ over $U$. In case 2 we
define $U=\mathrm{Spec}(R)$ and put ${\cal B}:={\mathbb{G}}_{m, U}$. Let $\mathfrak m$ be a
maximal ideal of $R$ contained in $U$, define $p=\mathrm{char}(R/\mathfrak m)$, and
denote by $\ol{B}={\cal B}\times_U \mathrm{Spec}(k(\mathfrak m))$ the special fibre at $\mathfrak m$.
Let $n$ be a positive integer which is coprime to $p$. Then the
restriction of ${\cal B}[n]$ to $S:=U[1/n]$ is a finite \'etale
group scheme over $S$ and $\mathfrak m\in S$. Let ${\mathfrak m}_\mathrm{nr}$ be a closed point of $S_{\mathrm{nr}}$ over $\mathfrak m$.
Taking a projective limit over the cospecialization maps (cf. Section \ref{preliminary})
$B[n](\widetilde{E})\cong \ol{B}[n](k({\mathfrak m}_\mathrm{nr}))$, we obtain an isomorphism
$$T(B)_{\neq p}\cong T(\ol{B})_{\neq p},$$
which induces a surjection
$T(\ol{B})_{\neq p, {\mathrm{Gal}}_{{\mathbb{F}}}}\to T({B})_{\neq p, {\mathrm{Gal}}_E}$, where
we have put ${\mathbb{F}}=k(\mathfrak m)$. The field ${\mathbb{F}}$ is {\em finite} by Lemma
\ref{exprime} and $\ol{B}$ is either an abelian variety over ${\mathbb{F}}$ (case 1)
or the multiplicative group scheme over ${\mathbb{F}}$ (case 2). In both cases it is
known that $T(\ol{B})_{\neq p, {\mathrm{Gal}}_{{\mathbb{F}}}}$ is finite (cf. \cite[Theorem 1 (ter), p. 299]{katzlang}).
This shows that
$T(B)_{\neq p, {\mathrm{Gal}}_E}$ is finite, whenever there exists a maximal ideal $\mathfrak m$ of
$R$ contained in $U$ with $\mathrm{char}(k(\mathfrak m))=p$. Now it follows by part b) of Lemma \ref{exprime} that
there are two different prime numbers $p_1\neq p_2$ such that
$T(B)_{\neq p_1, {\mathrm{Gal}}_E}$ and $T(B)_{\neq p_2, {\mathrm{Gal}}_E}$ are finite, and the
assertion follows from that.\hfill $\Box$
Let $K_0$ be a field of characteristic zero and $S/K_0$ a normal geometrically irreducible
variety with function field $K$. There is a canonical epimorphism
$p\colon \pi_1(S)\to {\mathrm{Gal}}_{K_0}$ (with kernel $\pi_1(S_{\widetilde{K_0}})$) and, following
Katz-Lang (\cite[p. 285]{katzlang}), we define ${\cal K}(S/K_0)$ to be the
kernel of the map $\pi_1(S)_{\mathrm{ab}}\to {\mathrm{Gal}}_{K_0, \mathrm{ab}}$ induced by $p$ on the abelianizations.
If we denote by $K_{S, \mathrm{nr}, \mathrm{ab}}$ the maximal abelian extension of $K$ which is
unramifield along $S$, then there is a diagram of fields
$$\begin{xy}
\xymatrix{
& K_{S, \mathrm{nr}, \mathrm{ab}}\ar@{-}[r]\ar@{-}[d] & \widetilde{K_0}K_{S, \mathrm{nr}, \mathrm{ab}} \ar@{-}[d]\\
K_{0, \mathrm{ab}}\ar@{-}[r]\ar@{-}[d] & K_{0, \mathrm{ab}}K \ar@{-}[r] \ar@{-}[d] & \widetilde{K_0}K \\
K_0\ar@{-}[r] & K &
}
\end{xy}$$
(cf. \cite[p. 286]{katzlang}) and the groups ${\mathrm{Gal}}(K_{S, \mathrm{nr}, \mathrm{ab}}/K_{0, \mathrm{ab}}K)$
and ${\mathrm{Gal}}(\widetilde{K_0}K_{S, \mathrm{nr}, \mathrm{ab}}/\widetilde{K_0}K)$ are both isomorphic to ${\cal K}(S/K_0)$.
The main result in the paper \cite{katzlang} of Katz and Lang is: If $K_0$ is finitely
generated and $S/K_0$ a smooth geometrically irreducible variety, then ${\cal K}(S/K_0)$
is finite. On the other hand, if $K_0$ is algebraically closed and $S/K_0$ is a
smooth proper geometrically irreducible curve of genus $g$,
then ${\cal K}(S/K_0)\cong \hat{{\mathbb{Z}}}^{2g}$ is infinite, unless $g=0$. In order to finish
up the proof of Proposition \ref{jordan} we have to prove the
finiteness of ${\cal K}(S/K_0)$ in the case of certain algebraic extensions $K_0/{\mathbb{Q}}$ (like
the field $\kappa_E$ in Lemma \ref{fin1})
which are {\em not} finitely generated but much smaller than $\widetilde{{\mathbb{Q}}}$.
\begin{lemm} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$.
Let $E/K$ be a (possibly infinite) Galois extension. Assume that ${\mathrm{Gal}}(E/K)$ has
finite exponent. Let $C/E$ be a smooth proper geometrically irreducible
curve and $S$ the complement of a divisor $D$ in $C$. Then
${\cal K}(S/E)$ is finite.\label{curves}
\end{lemm}
{\em Proof.} There is a finite extension $E'/E$ such that $S$ has an
$E'$-rational point and $D$ is $E'$-rational. There is a finite extension
$E''/E'$ which is Galois over $K$. Then ${\mathrm{Gal}}(E''/K)$ must have finite
exponent (because ${\mathrm{Gal}}(E/K)$ and ${\mathrm{Gal}}(E''/E)$ do). Furthermore
${\cal K}(S_{E''}/E'')$ surjects onto ${\cal K}(S/E)$ (cf. \cite[Lemma 1, p. 291]{katzlang}).
Hence we may assume from
the beginning that $S$ has an $E$-rational point and $D$ is $E$-rational.
The generalized Jacobian $J$ of $C$ with respect to the modulus
$D$ is a semiabelian variety.
(If $S=C$, then $J$ is just the usual Jacobian variety of $C$.)
Furthermore there is an isomorphism
$$\pi_1(S_{\widetilde{E}})_{\mathrm{ab}}\cong T(J).$$
On the other hand $\pi_1(S_{\widetilde{E}})_{\mathrm{ab}, {\mathrm{Gal}}_E}$ is isomorphic to ${\cal K}(S/E)$ (cf. \cite[Lemma 1, p. 291]{katzlang}).
Hence it is enough to prove that $T(J)_{{\mathrm{Gal}}_E}$ is finite. But this
has already been done in Lemma \ref{mwlemm}.\hfill $\Box$
\begin{lemm} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$. Let
$E/K$ be a (possibly infinite) Galois extension. Assume that ${\mathrm{Gal}}(E/K)$ has
finite exponent. Let $S/E$ be a normal geometrically irreducible variety.
Then ${\cal K}(S/E)$ is finite.\label{katzlanglemm}
\end{lemm}
{\em Proof.} There is a finite extension $L/E$ and a sequence of elementary fibrations
in the sense of M. Artin (cf. \cite[Expos\'e XI, 3.1-3.3]{SGA4})
$$\begin{xy}
\xymatrix{
\mathrm{Spec}(L)=U_0\ar@{<-}[r]^{f_1} & U_1 \ar@{<-}[r]^{f_2} & U_2\ar@{<-}[r]^{f_3} &\cdots\ar@{<-}[r]^{f_n} & U_n\subset S_L
}
\end{xy}
$$
where $U_n$ is a non-empty open subscheme of $S_L$. Then $\dim(U_i)=i$ for all $i$. We may assume that $L/K$ is
Galois and then ${\mathrm{Gal}}(L/K)$ is of finite exponent. Let $L_i$ be the
function field of $U_i$. Then the generic fibre $S_{i+1}:=U_{i+1}\times_{U_i} \mathrm{Spec}(L_i)$
of $f_{i+1}$ is a curve over $L_i$ which is the complement of a divisor in a smooth proper
geometrically irreducible curve $C_{i+1}/L_i$. The extension $L_i/L$ is finitely generated
(of transcendence degree $i$). Hence $L_i=L(u_1,\cdots, u_s)$ for certain elements
$u_1,\cdots, u_s\in L_i$. Let us define $K_i:=K(u_1,\cdots, u_s)$. Then there
is a diagram of fields
$$\begin{xy}
\xymatrix{
K_i\ar@{-}[rr]\ar@{-}[d] & & L_i\ar@{-}[d]\\
K\ar@{-}[r]\ar@{-}[d] & E\ar@{-}[r] & L\\
{\mathbb{Q}} &&
}
\end{xy}
$$
such that the vertical extensions are all finitely generated and $L_i=K_iL$.
The extension $L_i/K_i$ is Galois because $L/K$ is Galois, and the restriction
map ${\mathrm{Gal}}(L_i/K_i)\to{\mathrm{Gal}}(L/K)$ is injective. Hence ${\mathrm{Gal}}(L_i/K_i)$ is a group of
finite exponent and $K_i$ is finitely generated.
Lemma \ref{curves} implies that
${\cal K}(S_{i+1}/L_i)$ is finite for every $i\in\{0,\cdots, n-1\}$. By
\cite[Lemma 2]{katzlang} and \cite[(1.4)]{katzlang} it follows that ${\cal K}(U_n/L)$ is finite.
Then \cite[Lemma 3]{katzlang} implies that ${\cal K}(S_L/L)$ is finite, and \cite[Lemma 1]{katzlang}
shows that ${\cal K}(S/E)$ is finite, as desired.\hfill $\Box$.
{\em Proof of Proposition \ref{jordan}.} Let $S/{\mathbb{Q}}$ be a normal variety
with the function field $K$. Let $E/K$ be a $d$-Jordan extension contained in
the extension $K_{S, \mathrm{nr}}/K$
There is a $d$-flat extension $L/K$ in $E$ such that $E/L$ is abelian.
By Lemma \ref{fin1} $L/\kappa_LK$ is a finite extension. We have the following diagram of fields.
$$\begin{xy}
\xymatrix{
\kappa_E\ar@{-}[r]\ar@{-}[d] & \kappa_E K\ar@{-}[r]\ar@{-}[d] & \kappa_E L\ar@{-}[r]\ar@{-}[d] & E\ar@{-}[r] & K_{S, \mathrm{nr}} \\
\kappa_L\ar@{-}[r]\ar@{-}[d] & \kappa_L K\ar@{-}[r]\ar@{-}[d] & L & & \\
\kappa_K\ar@{-}[r] & K
}
\end{xy}$$
Now $S^{(L)}$ is the normalization of the geometrically irreducible
$\kappa_L$-variety $$S^{(\kappa_LK)}=S\times_{\kappa_K}\mathrm{Spec}(\kappa_L)$$ in the
{\em finite} extension $L/\kappa_LK$. Hence $S^{(L)}$ is a geometrically
irreducible variety over $\kappa_L$. (The crucial point is that $S^{(L)}$ is
of finite type over $\kappa_L$.) The extension $E/L$ is abelian and unramified along $S^{(L)}$.
Hence ${\mathrm{Gal}}(E/\kappa_EL)$ is a quotient of ${\cal K}(S^{(L)}/\kappa_L)$.
The field $\kappa_K$ is a {\em number field} and ${\mathrm{Gal}}(\kappa_L/\kappa_K)$ is a
group of exponent $d<\infty$, because it is a quotient of ${\mathrm{Gal}}(L/K)$ (cf. Remark \ref{elrema}). Hence
Lemma \ref{katzlanglemm} implies that ${\cal K}(S^{(L)}/\kappa_L)$ is finite.
It follows that $E/\kappa_EL$ is a finite extension.
Now $\kappa_EL/\kappa_EK$ is finite, because $L/\kappa_L K$ is finite. It follows
that $E/\kappa_E K$ is finite, as desired.\hfill $\Box$
\section{Representations of the fundamental group}
We start this section with two
remarks and a lemma about families of representations of certain profinite groups.
Then we prove an independence criterion for families
of representations of the \'etale fundamental group $\pi_1(S)$ of a normal ${\mathbb{Q}}$-variety
$S$ (cf. Theorem \ref{crit}). This criterion is the technical heart of the paper.
\begin{rema} Let $K$ be a field, $\Omega/K$ a Galois extension and $I\subset {\mathbb{N}}$. Let $(\Gamma_i)_{i\in I}$ be a family of profinite
groups. For every $i\in I$ let $\rho_i: {\mathrm{Gal}}(\Omega/K)\to \Gamma_i$ be a continuous homomorphism.
Let $K_i$ be the fixed field of $\ker(\rho_i)$ in $\Omega$.
Then the following conditions are equivalent.
\begin{enumerate}
\item[(i)] The family $(\rho_i)_{i\in I}$ is independent.
\item[(ii)] The family $(K_i)_{i\in I}$ of
fields is linearly disjoint over $K$.
\item[(iii)] If $s\ge 1$ and $i_1<i_2<\cdots < i_{s+1}$ are elements of $I$, then
$$K_{i_1}\cdots K_{i_s}\cap K_{i_{s+1}}=K.$$
\label{indeprem}
\end{enumerate}
\end{rema}
{\em Proof.} As the homomorphisms $\rho_i$ induce
isomorphisms ${\mathrm{Gal}}(K_i/K)\cong \mathrm{im}(\rho_i)$, (i) is satisfied
if and only if the natural map
${\mathrm{Gal}}(\Omega/K)\to \prod_{i\in I} {\mathrm{Gal}}(K_i/K)$ is surjective, and
this is in turn equivalent to (ii) (cf. \cite[2.5.6]{friedjarden}).
It is well-known that (ii) is equivalent
to (iii) (cf. \cite[p. 36]{friedjarden}).\hfill $\Box$
\begin{rema} Let $\Gamma$ be a profinite group and $n\in {\mathbb{N}}$.
For every $\ell\in{\mathbb{L}}$ let $\Gamma_\ell$ be a profinite group and $\rho_\ell\colon \Gamma\to \Gamma_\ell$
a continuous homomorphism. Assume that for every $\ell\in {\mathbb{L}}$ there is an integer $n\in {\mathbb{N}}$ such that $\Gamma_\ell$
is isomorphic to a subquotient of $\mathrm{GL}_n({\mathbb{Z}}_\ell)$.
\begin{enumerate}
\item[a)] Let $\Gamma'\subset \Gamma$ be an open subgroup.
If the family $(\rho_\ell)_{\ell\in {\mathbb{L}}}$ is independent,
then there is a finite subset $I\subset {\mathbb{L}}$ such that
the family $(\rho_\ell)_{\ell\in {\mathbb{L}}\smallsetminus I}$ is independent
over $\Gamma'$.
\item[b)] The following conditions (i) and (ii) are equivalent.
\begin{enumerate}
\item[(i)] The family $(\rho_\ell)_{\ell\in {\mathbb{L}}}$ is almost
independent.
\item[(ii)] There exists a finite subset $I\subset {\mathbb{L}}$ such
that $(\rho_\ell)_{\ell\in {\mathbb{L}}\smallsetminus I}$ is almost independent.\label{indeprema}
\end{enumerate}
\end{enumerate}
\end{rema}
{\em Proof.} Let $\rho\colon \Gamma\to \prod_{\ell\in{\mathbb{L}}} \Gamma_\ell$ be
the homomorphism induced by the $\rho_\ell$.
To prove a) assume that $\rho(\Gamma)=\prod_{\ell\in{\mathbb{L}}} \rho_\ell(\Gamma)$. The subgroup
$\rho(\Gamma')$ is open in $\prod_{\ell\in{\mathbb{L}}} \rho_\ell(\Gamma)$, because a
surjective homomorphism of profinite groups is open (cf. \cite[p. 5]{friedjarden}).
It follows from the definition of the product topology that
there is a finite subset $I\subset {\mathbb{L}}$ such that
$\rho(\Gamma')\supset \prod_{\ell\in I}\{1\}\times \prod_{\ell\in{\mathbb{L}}\smallsetminus I} \rho_\ell(\Gamma)$. This
implies that $(\rho_\ell)_{\ell\in{\mathbb{L}}\smallsetminus I}$ is independent over $\Gamma'$ and finishes the proof of part a).
For part b) see \cite[Lemme 3]{bible}.\hfill $\Box$
Let $K$ be a field, $n\in{\mathbb{N}}$ and $\Omega/K$ a fixed Galois extension.
For every $\ell\in{\mathbb{L}}$ let $\Gamma_\ell$ be a profinite group and $\rho_\ell\colon {\mathrm{Gal}}(\Omega/K)\to \Gamma_\ell$
a continuous homomorphism. Assume that $\Gamma_\ell$ is isomorphic to a subquotient of
$\mathrm{GL}_n({\mathbb{Z}}_\ell)$ for every $\ell\in {\mathbb{L}}$. Denote by $K_\ell$ the fixed field
in $\Omega$ of the kernel of $\rho_\ell$. Then $K_\ell$ is a Galois
extension of $K$ and $\rho_\ell$ induces an isomorphism
${\mathrm{Gal}}(K_\ell/K)\cong \rho_\ell({\mathrm{Gal}}(\Omega/K))$.
For every extension $E/K$ contained in $\Omega$ and
every $\ell\in{\mathbb{L}}$ we
define $G_{\ell, E}:=\rho_\ell({\mathrm{Gal}}(\Omega/E))$ and $E_\ell:=EK_\ell$. Then
$G_{\ell, E}$ is isomorphic to a subquotient of $\mathrm{GL}_n({\mathbb{Z}}_\ell)$ and
$\rho_\ell$ induces an isomorphism
$${\mathrm{Gal}}(E_\ell/E)\cong G_{\ell, E}.$$ Furthermore we define
$G_{\ell, E}^+$ to be the subgroup of $G_{\ell, E}$
generated by its $\ell$-Sylow subgroups. Then $G_{\ell, E}^+$
is normal in $G_{\ell, E}$. Finally we
let $E_\ell^+$ be the fixed field of $\rho_\ell^{-1}(G_{\ell, E}^+)\cap{\mathrm{Gal}}(\Omega/E)$.
Then $E_\ell^+$ is an intermediate field of $E_\ell/E$ which is
Galois over $E$, the group ${\mathrm{Gal}}(E_\ell/E_\ell^+)$ is isomorphic to $G_{\ell, E}^+$
and ${\mathrm{Gal}}(E_\ell^+/E)$ is isomorphic to
$G_{\ell, E}/G_{\ell, E}^+$.
\begin{lemm} Let $E/K$ be a Galois extension contained in $\widetilde{K}$ and
let $\ell\in{\mathbb{L}}$.\label{lemm1}
\begin{enumerate}
\item[a)] The extension $E_\ell^+/E$ is a {\em finite} Galois extension, and
${\mathrm{Gal}}(E_\ell^+/E)$ is isomorphic to a subquotient of
$\mathrm{GL}_n({\mathbb{F}}_\ell)$.
\item[b)] If $E/K$ is finite and $[E:K]$ is not divisible by $\ell$, then
$G_{\ell, E}^+=G_{\ell, K}^+$ and $EK_\ell^+=E_\ell^+$.
\end{enumerate}
\end{lemm}
{\em Proof.} The profinite group $G_{\ell, E}$ is a closed normal subgroup
of $G_{\ell, K}$ and $G_{\ell, K}$ is isomorphic to a subquotient of $\mathrm{GL}_n({\mathbb{Z}}_\ell)$.
Hence there is a closed subgroup $U_\ell$ of $\mathrm{GL}_n({\mathbb{Z}}_\ell)$ and a closed normal
subgroup $V_\ell$ of $U_\ell$ such that there is an isomorphism $i: G_{\ell, E}\to U_\ell/V_\ell$. Furthermore there is a closed normal subgroup $U_\ell^+$ of $U_\ell$ containing
$V_\ell$ such that $i(G_{\ell, K}^+)=U_\ell^+/V_\ell$. The group $U_\ell/U_\ell^+$ is isomorphic to
$G_{\ell, E}/G_{\ell, E}^+$. Its order is coprime to $\ell$.
The kernel of the restriction map
$r\colon \mathrm{GL}_n({\mathbb{Z}}_\ell)\to \mathrm{GL}_n({\mathbb{F}}_\ell)$ is a pro-$\ell$ group; hence the
intersection of this
kernel with $U_\ell$ is contained in $U_\ell^+$. This shows that $r$ induces an
isomorphism $U_\ell/U_\ell^+\to r(U_\ell)/r(U_\ell^+)$. Altogether we see that
$${\mathrm{Gal}}(E_\ell^+/E)\cong G_{\ell, E}/G_{\ell, E}^+\cong U_\ell/U_\ell^+\cong r(U_\ell)/r(U_\ell^+) $$
and Part a) follows,
because $r(U_\ell)/r(U_\ell^+) $ is obviously a subquotient of
$\mathrm{GL}_n({\mathbb{F}}_\ell)$.
Every $\ell$-Sylow subgroup of $G_{\ell, E}$ lies in an $\ell$-Sylow subgroup
of $G_{\ell, K}$, hence $G_{\ell, E}^+\subset G_{\ell, K}^+$. Assume from
now on that $[E:K]$ is finite and not divisible by $\ell$. Then
every $\ell$-Sylow subgroup of $G_{\ell, K}$ must map to the trivial group
under the projection $G_{\ell, K}\to G_{\ell, K}/G_{\ell, E}$, because
the order of the quotient group is coprime to $\ell$. Hence every
$\ell$-Sylow subgroup of $G_{\ell, K}$ lies in $G_{\ell, E}$. This shows
that $G_{\ell, K}^+=G_{\ell, E}^+$. The Galois group
${\mathrm{Gal}}(E_\ell/EK_\ell^+)$ is $G_{\ell, K}^+\cap G_{\ell, E}$ and the
Galois group ${\mathrm{Gal}}(E_\ell/EK_\ell^+)$ is $G_{\ell, E}^+$. As $G_{\ell, K}^+=G_{\ell, E}^+$ it follows
that ${\mathrm{Gal}}(E_\ell/EK_\ell^+)={\mathrm{Gal}}(E_\ell/E_\ell^+)$, hence
$EK_\ell^+=E_\ell^+$.\hfill $\Box$
Let $S$ be a normal ${\mathbb{Q}}$-variety with function field $K$.
We shall now study families of representations of the fundamental group
$\pi_1(S)$ (viewing $S$ as a scheme equipped with the generic
geometric point $\mathrm{Spec}(\widetilde{K})\to K$).
Recall that we may identify $\pi_1(S)$ with
${\mathrm{Gal}}(K_{S, \mathrm{nr}}/K)$
\begin{thm} Let $S/{\mathbb{Q}}$ be a normal variety with function field $K$.
Let $P_\mathrm{nr}\in S_{\mathrm{nr}}$ be a closed point.
For every $\ell\in{\mathbb{L}}$ let $\Gamma_\ell$ be a profinite group and
$\rho_\ell\colon \pi_1(S)\to \Gamma_\ell$ a
continuous homomorphism. We make two assumptions.
\begin{enumerate}
\item[a)] Assume there is an integer $n\in {\mathbb{N}}$ such that for every $\ell\in{\mathbb{L}}$ the profinite group $\Gamma_\ell$
is isomorphic to a subquotient of $\mathrm{GL}_n({\mathbb{Z}}_\ell)$.
\item[b)]
Assume that there exists an open subgroup $D'$ of the decomposition group $D_{K_{S, \mathrm{nr}}/K}(P_{\mathrm{nr}})$ such that the
family $(\rho_\ell)_{\ell\in{\mathbb{L}}}$ is independent over $D'$.
\end{enumerate}
Then
the family $(\rho_\ell)_{\ell\in{\mathbb{L}}}$ is almost independent.\label{crit}
\end{thm}
This theorem may seem surprising at the first glance, since $D_{K_{S, \mathrm{nr}}/K}(P_{\mathrm{nr}})$ is usually far from being
open in $\pi_1(S)$. The proof of Theorem \ref{crit} occupies the rest of this section. From now
on all the assumptions of Theorem \ref{crit} are in force, until the proof is finished.
For every algebraic extension $E/K$ contained in $K_{S, \mathrm{nr}}$ we define
$G_{\ell, E}=\rho_\ell({\mathrm{Gal}}(K_{S, \mathrm{nr}}/E))$, $G_{\ell, E}^+$, $E_\ell$ and $E_\ell^+$ exactly
as before. Furthermore we shall write $P_E$ for the point in
$S^{(E)}$ below $P_{\mathrm{nr}}$.
We tacitly assume in the sequal that $\widetilde{{\mathbb{Q}}}$ denotes the algebraic closure
of $K$ {\em inside} $\widetilde{K}$. Then already $K_{S, \mathrm{nr}}$ contains
$\widetilde{{\mathbb{Q}}}$, because the constant field extensions of $K$ are unramified along $S$.
The structure morphism $S_{\mathrm{nr}}\to\mathrm{Spec}({\mathbb{Q}})$ factors through $\mathrm{Spec}(\widetilde{{\mathbb{Q}}})$,
because $S_{\mathrm{nr}}$ is normal. It follows in particular that $k(P_{\mathrm{nr}})=
\widetilde{{\mathbb{Q}}}$.
\begin{lemm} There is a finite Galois extension
$E/K$ contained in $K_{S, \mathrm{nr}}$ and a finite subset
$I\subset {\mathbb{L}}$ such that the following
statements about $E$ and $I$ hold true:
\begin{enumerate}
\item[a)] For all $\ell\in{\mathbb{L}}\smallsetminus I$ the
extension $E_\ell^+/E$ is a constant field extension, that is:
$\kappa_{E_\ell^+}E=E_\ell^+$.
\item[b)] The point $P_E$ is a $\kappa_E$-rational point of $S^{(E)}$.
\item[c)] The family $(\rho_\ell)_{\ell\in {\mathbb{L}}\smallsetminus I}$ is independent
over $D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})$.\label{lemm2}
\end{enumerate}
\end{lemm}
{\em Proof.} Let $L:=\prod_{\ell\in{\mathbb{L}}} K_\ell^+$ be the composite field
of all the $K_\ell^+$. By Lemma \ref{lemm1}, for each $\ell\in{\mathbb{L}}$, the group
${\mathrm{Gal}}(K_\ell^+/K)$ is isomorphic to a subquotient
of $\mathrm{GL}_n({\mathbb{F}}_\ell)$, and $|{\mathrm{Gal}}(K_\ell^+/K)|$ is not divisible by $\ell$.
By \cite[Th\'eor\`eme 3']{bible} (which is a generalization
due to Serre of the classical theorem of Jordan) it follows that there
is an integer $d$ (independent of $\ell$) such that for every
$\ell\in {\mathbb{L}}$ the group ${\mathrm{Gal}}(K_\ell^+/K)$ has an abelian normal subgroup
$A_\ell$ of index $[{\mathrm{Gal}}(K_\ell^+/K):A_\ell]\le d$. Let $K_\ell'$ be the
fixed field of $A_\ell$ in $K_\ell^+$. Then $K':=\prod_\ell K_\ell'$ is
a $d$-flat extension of $K$ and $K'K_\ell^+/K'$ is abelian for every $\ell\in{\mathbb{L}}$.
It follows that $L/K$ is a $d$-Jordanian extension. Furthermore
$L/K$ is contained in $K_{S, \mathrm{nr}}$. By Proposition \ref{jordan},
$L$ is a {\em finite} extension of $\kappa_LK$. Note that $\kappa_L/{\mathbb{Q}}$ may well be
an infinite extension. Hence there is an element $\omega\in L$ such that $L=\kappa_LK(\omega)$.
Let $E_1$ be the Galois closure of $K(\omega)/K$ in $L$. Then $E_1/K$ is a finite
Galois extension and $\kappa_L E_1=L$. Hence we have a diagram of fields
$$\begin{xy}
\xymatrix{
\kappa_L K\ar@{-}[r]\ar@{-}[d] & L\ar@{-}[d]\\
K\ar@{-}[r] & E_1
}
\end{xy}$$
in which the vertical extensions are constant field extensions and in which
the horizontal extensions are finite. Furthermore $L$ contains $K_\ell^+$
for every $\ell\in{\mathbb{L}}$.
Now consider the canonical isomorphism
$$r\colon D_{K_{S, \mathrm{nr}}/K}(P_{\mathrm{nr}})\cong {\mathrm{Gal}}(k(P_{\mathrm{nr}})/{k(P_K)}).$$
Let $\lambda_1$ be the fixed field of $r(D')$ in $k(P_{\mathrm{nr}})=\widetilde{{\mathbb{Q}}}$.
Since $D'$ is open in $D_{K_{S, \mathrm{nr}}/K}(P_{\mathrm{nr}})$, the field $\lambda_1$ is a
finite extension of $k(P_K)$, so $\lambda_1$ is a finite extension of ${\mathbb{Q}}$.
Choose a finite Galois extension $\lambda/\kappa_K$ containing
$\lambda_1$ and $k(P_{E_1})$, and define $E:=\lambda E_1$. Then $S^{(E)}=S^{(E_1)}
\times_{\kappa_{E_1}} \mathrm{Spec}(\lambda)$ and $\kappa_E=\lambda$. There is the following
diagram of number fields:
$$\begin{xy}
\xymatrix{
\lambda_1\ar@{-}[r]\ar@{-}[d] & \lambda\ar@{-}[d]\ar@{=}[r] & \kappa_E\ar@{-}[r] &k(P_E)\\
k(P_K)\ar@{-}[r]\ar@{-}[d] & k(P_{E_1})\ar@{-}[d]&&\\
\kappa_K\ar@{-}[r]& \kappa_{E_1}&&\\
}
\end{xy}$$
The fibre of $P_{E_1}$
under the projection $S^{(E)}\to S^{(E_1)}$ is $\mathrm{Spec}(\kappa_E\otimes_{\kappa_{E_1}} k(P_{E_1}))$, and this fibre splits up into the coproduct of $[k(P_{E_1}):\kappa_{E_1}]$
many copies of $\mathrm{Spec}(\kappa_E)=\mathrm{Spec}(\lambda)$, because $\lambda/\kappa_E$ is Galois
and $\lambda\supset k(P_{E_1})$. Thus all points in $S^{(E)}$ over $P_{E_1}$ are
$\kappa_E$-rational. In particular $P_E$ is $\kappa_E$-rational.
It follows that $$r(D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}}))=
{\mathrm{Gal}}(k(P_{\mathrm{nr}})/{k(P_E)})={\mathrm{Gal}}(k(P_{\mathrm{nr}})/\kappa_E),$$
and this group is an open subgroup of
$r(D')={\mathrm{Gal}}(k(P_{\mathrm{nr}})/\lambda_1)$ because
$\kappa_E$ is a finite extension of $\lambda_1$. Hence $D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})$
is an open subgroup of $D'$.
As $(\rho_\ell)_{\ell\in{\mathbb{L}}}$ is independent over $D'$ by one
of our assumptions, it follows from part a) of Remark
\ref{indeprema} that there is a finite subset $I'\subset {\mathbb{L}}$ such
that the family $(\rho_\ell)_{\ell\in{\mathbb{L}}\smallsetminus I'}$ is independent
over $D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})$. Finally $K_\ell^+E/E$ is a
constant field extension,
because $K_\ell^+E$ is an intermediate field of
$LE/E$ and $LE=\kappa_LE$ is a constant field extension of $E$
due to our construction. By Lemma \ref{lemm1} we see that
$E_\ell^+=K_\ell^+E$ for all $\ell\in {\mathbb{L}}$ which do not divide
the index $[E:K]$. Hence assertions a), b) and c) follow, if we
put $I:=I'\cup\{\ell\in{\mathbb{L}}: \ell\ \mbox{divides}\ [E:K]\}$.\hfill $\Box$
\begin{lemm} Let $E$ and $I$ be as in Lemma \ref{lemm2}. Let
$s\ge 1$. Let $\ell_1<\cdots<\ell_{s+1}$ be some elements of
${\mathbb{L}}\smallsetminus I$. Then
$E_{\ell_1}\cdots E_{\ell_s}\cap E_{\ell_{s+1}}$ is a regular
extension of $\kappa_E$ (i.e. the algebraic closure of ${\mathbb{Q}}$ in
$E_{\ell_1}\cdots E_{\ell_s}\cap E_{\ell_{s+1}}$ is $\kappa_E$).\label{lemm3}
\end{lemm}
{\em Proof.} The canonical isomorphism
$$r\colon D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})\cong {\mathrm{Gal}}(k(P_{\mathrm{nr}})/k(P_E))$$
induces by restriction an isomorphism $$D_{K_{S, \mathrm{nr}}/E_\ell}(P_{\mathrm{nr}})=
D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})\cap {\mathrm{Gal}}(K_{S, \mathrm{nr}}/E_\ell)\cong {\mathrm{Gal}}(k(P_{\mathrm{nr}})/k(P_{E_\ell}))$$
for every $\ell\in{\mathbb{L}}$. Hence $k(P_{E_\ell})$ is the fixed field in
$k(P_{\mathrm{nr}})$ of the kernel of $\rho_\ell\circ r^{-1}$. The
family $(\rho_\ell)_{\ell\in {\mathbb{L}}\smallsetminus I}$ is independent over
$D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})$ by Lemma \ref{lemm2}. Hence Remark \ref{indeprem} shows that
$(k(P_{E_\ell}))_{\ell\in{\mathbb{L}}\smallsetminus I}$ is linearly disjoint over
$k(P_E)$. Define $F:=E_{\ell_1}\cdots E_{\ell_s}$.
There is a diagram of
residue fields:
$$\begin{xy}
\xymatrix{
& & k(P_{\mathrm{nr}})\ar@{-}[d]\\
& & k(P_F)\ar@{-}[lld]\ar@{-}[ld]\ar@{-}[d]\ar@{-}[rrd] & & \\
k(P_{E_{\ell_1}}) &k(P_{E_{\ell_2}}) & k(P_{E_{\ell_3}}) & \cdots & k(P_{E_{\ell_{s}}}) \\
& & k(P_E)\ar@{-}[llu]\ar@{-}[lu]\ar@{-}[u]\ar@{-}[rru] & &
}\end{xy}$$
We have $k(P_F)=k(P_{E_{\ell_1}})\cdots k(P_{E_{\ell_s}})$,
because
$$\begin{array}{rcl}
{\mathrm{Gal}}(k(P_{\mathrm{nr}})/k(P_{E_{\ell_1}})\cdots k(P_{E_{\ell_s}}))&=&\bigcap_{i=1}^s G(k(P_{\mathrm{nr}})/k(P_{E_{\ell_i}}))=\\
&=& r(\bigcap_{i=1}^s D_{K_{S, \mathrm{nr}}/E_{\ell_i}}(P_{\mathrm{nr}}))=\\
&=&r(D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})\cap \bigcap_{i=1}^s {\mathrm{Gal}}(K_{S, \mathrm{nr}}/E_{\ell_i}))=\\
&=&r(D_{K_{S, \mathrm{nr}}/E}(P_{\mathrm{nr}})\cap {\mathrm{Gal}}(K_{S, {\mathrm{nr}}}/F))\\
&=& r(D_{K_{S, \mathrm{nr}}/F}(P_{\mathrm{nr}}))=G(k(P_{\mathrm{nr}})/k(P_F)).\end{array}$$
Furthermore there is a diagram
$$\begin{xy}
\xymatrix{
k(P_F)\ar@{-}[dr] & & k(P_{E_{\ell_{s+1}}})\ar@{-}[dl]\\
& k(P_{F\cap E_{\ell_{s+1}}})\ar@{-}[d] &\\
& k(P_E) &
}
\end{xy}$$
and $k(P_F)\cap k(P_{E_{\ell_{s+1}}})=k(P_E)$ due to the fact that $(k(P_{E_\ell}))_{\ell\in{\mathbb{L}}\smallsetminus I}$ is
linearly disjoint over $k(P_E)$.
It follows that $k(P_{F\cap E_{\ell_{s+1}}})=k(P_E)$. Finally $k(P_E)=\kappa_E$, because $P_E$ is a $\kappa_E$-rational point
of $S^{(E)}$. This shows that the normalization of $S^{(E)}$ in
$F\cap E_{\ell_{s+1}}$ has a $\kappa_E$-rational point and thus its function field $F\cap E_{\ell_{s+1}}$ must
be regular over $\kappa_E$. \hfill $\Box$
Let $\ell\ge 5$ be a prime number. We denote by $\Sigma_\ell$ the set of isomorphism classes
of groups which are either the cyclic group ${\mathbb{Z}}/\ell$, or the quotient of $\underline{H}(F)$ modulo
its center, where $F$ is a finite field of characteristic $\ell$ and $\underline{H}$ is a connected smooth
algebraic group over $F$ which is geometrically simple and simply connected. These are the simple groups of Lie
type in characteristic $\ell$. It is known (cf. \cite[Th\'eor\`eme 5]{bible}), that
$\Sigma_\ell\cap\Sigma_{\ell'}=\emptyset$ for all primes $5\le \ell<\ell'$.
(As Serre points out in \cite{bible}, the proof of this theorem is essentially due to
E. Artin \cite{artin}. It was completed in \cite{KLMS}.) In the following proof we shall strongly use
this result.
{\em End of Proof of Theorem \ref{crit}.} Let $E$ and $I$ be as in Lemma \ref{lemm2}.
In order to finish up the proof of Theorem
\ref{crit} it suffices to prove the following
{\bf Claim.} {\em There is a finite subset $I'\subset {\mathbb{L}}$ containing $I$, such that
$(E_\ell)_{\ell\in{\mathbb{L}}\smallsetminus I'}$ is linearly disjoint over $E$.}
In fact, once this claim is proven, it follows that $(\rho_\ell)_{\ell\in {\mathbb{L}}\smallsetminus I'}$ is
independent over ${\mathrm{Gal}}(K_{S, \mathrm{nr}}/E)$ by Remark \ref{indeprem}, and Remark \ref{indeprema} implies
that the whole family $(\rho_\ell)_{\ell\in {\mathbb{L}}}$ must be almost independent, as desired.
In \cite[Th\'eor\`eme 4]{bible} Serre proves: {\em There is
a constant $C$ such that for every prime number $\ell>C$ every
finite simple subquotient of $GL_n({\mathbb{Z}}_\ell)$ of order divisible by $\ell$ lies in $\Sigma_\ell$.}
This is a generalization of a well-known result of Nori (cf. \cite[Theorem B]{nori1987}).
Let us define $I':=I\cup \{2, 3\}\cup \{\ell\in {\mathbb{L}}: \ell\le C\}$. For $\ell\in{\mathbb{L}}$ every
non-trivial quotient of $G_{E, \ell}^+$ has order divisible by $\ell$: In fact, if
$h: G_{E, \ell}^+\to Q$ is an epimorphism onto a non-trivial group $Q$,
then the image of some $\ell$-Sylow subgroup of $G_{E, \ell}$ under $h$ must be non-trivial.
Hence, for every $\ell\in {\mathbb{L}}\smallsetminus I'$ every finite
simple quotient of $G_{E, \ell}^+$ lies in $\Sigma_\ell$.
We shall now prove the Claim. Let $s\ge 1$ and $\ell_1<\cdots<\ell_{s+1}$ be elements
of ${\mathbb{L}}\smallsetminus I'$. It suffices to show that $E_{\ell_1}\cdots E_{\ell_s}\cap E_{\ell_{s+1}}=E$,
assuming by induction that the sequence $(E_{\ell_1},\cdots, E_{\ell_s})$ is already linearly disjoint over $E$.
This assumption implies
$${\mathrm{Gal}}(E_{\ell_1}\cdots E_{\ell_s}/E)\cong G_{\ell_1, E}\times \cdots \times G_{\ell_s, E}$$
and ${\mathrm{Gal}}(E_{\ell_1}\cdots E_{\ell_s}/E_{\ell_1}^+\cdots E_{\ell_s}^+)\cong G_{\ell_1, E}^+\times \cdots\times G_{\ell_s, E}^+$.
Suppose that $E_{\ell_1}\cdots E_{\ell_s}\cap E_{\ell_{s+1}}\neq E$. Then there would be an intermediate field $L$
of that extension such that $Q:={\mathrm{Gal}}(L/E)$ is a finite simple group. We would have the following
diagram of fields:
$$\begin{xy}
\xymatrix{
E_{\ell_1}\cdots E_{\ell_s}\ar@{-}[d]_{G_{\ell_1, E}^+\times \cdots \times G_{\ell_s, E}^+}\ar@{-}[dr] & & E_{\ell_{s+1}}\ar@{-}[dl]\ar@{-}[d]^{G_{\ell_{s+1}, E}^+}\\
E_{\ell_1}^+\cdots E_{\ell_s}^+\ar@{-}[dr] & L\ar@{-}[d]^Q & E_{\ell_{s+1}}^+\ar@{-}[dl]\\
& E &
}\end{xy}$$
But $L/\kappa_E$ is a regular extension (cf. Lemma \ref{lemm3}), hence $\kappa_L=\kappa_E$.
On the other hand $E_{\ell_i}^+/E$ is a constant field extension for every $i=1,\cdots, s+1$ (cf. Lemma \ref{lemm2}).
It follows that
${\mathrm{Gal}}(LE_{\ell_{s+1}}^+/E_{\ell_{s+1}}^+)\cong Q$ and
${\mathrm{Gal}}(LE_{\ell_1}^+\cdots E_{\ell_s}^+/E_{\ell_1}^+\cdots E_{\ell_s}^+)\cong Q$. Hence $Q$ is simultaneously a quotient group
of $G_{\ell_1, E}^+\times\cdots\times G_{\ell_s, E}^+$ and of $G_{\ell_{s+1, E}}^+$.
It follows that $$Q\in \left(\Sigma_{\ell_1}\cup\cdots\cup\Sigma_{\ell_s}\right)\cap\Sigma_{\ell_{s+1}},$$
which contradicts Artin's theorem that $\Sigma_\ell\cap\Sigma_{\ell'}=\emptyset$ for all primes
$5\le \ell<\ell'$.\hfill $\Box$
\section{Proof of the main theorem}
{\em Proof of Theorem \ref{main2}.} Let $K$ be a finitely generated extension of ${\mathbb{Q}}$. Let $X/K$ be
a separated scheme of finite type.
Let $T=(T_1,\cdots, T_r)$ be a transcendence base of $K/{\mathbb{Q}}$ and $S_0$ be
the normalization of $\mathrm{Spec}({\mathbb{Q}}[T])$ in $K$. Then $S_0$ is a normal ${\mathbb{Q}}$-variety with
function field $K$.
The spreading-out principles in \cite{EGAIV3} (cf. in particular \cite{EGAIV3}[8.8.2],
\cite{EGAIV3}[8.10.5],
\cite{EGAIV3}[8.9.4]), allow
us to construct a dense open subscheme $S\subset S_0$ and a flat separated morphism of finite type
$f: {\cal X}\to S$ with generic fibre $X$.
We choose a closed point $P\in S$ and a closed point
$P_{\mathrm{nr}}\in S_{\mathrm{nr}}$ over $P$ and denote by
$\ol{P}: \mathrm{Spec}(k(P_{\mathrm{nr}}))\to S_{\mathrm{nr}}\to S$ the corresponding
geometric point of $S$. Note that $k(P_{\mathrm{nr}})$ is algebraically closed (cf. the second paragraph after
Theorem \ref{crit}). We define $\widetilde{k}:=k(P_{\mathrm{nr}})$.
Furthermore we denote by $\ol{\xi}\colon \mathrm{Spec}(\widetilde{K})\to S$
the generic geometric point of $S$ afforded
by the choice of $\widetilde{K}$. We let $X_P:={\cal X}\times_S k(P)$, $X_{\ol{P}}={\cal X}\times_S \mathrm{Spec}(k(P_{\mathrm{nr}}))$ and
$X_{\ol{\xi}}={\cal X}\times_S \mathrm{Spec}(\widetilde{K})$ be the corresponding fibres of
${\cal X}$. Note that $X_{\ol{\xi}}=X_{\widetilde{K}}$ and $X_{\ol{P}}=X_{P, \widetilde{k}}$.
Let $q\in{\mathbb{N}}$. From now on we shall consider two cases. For the first case we define
$\rho_\ell:=\rho_{\ell, X}^{(q)}$, $T_\ell:=\mathrm{H}^q(X_{\ol{\xi}}, {\mathbb{Z}}_\ell)$, $V_\ell:=\mathrm{H}^q(X_{\ol{\xi}}, {\mathbb{Q}}_\ell)$,
$T_{\ell, P}:=\mathrm{H}^q(X_{\ol{P}}, {\mathbb{Z}}_\ell)$, $V_{\ell, P}:=\mathrm{H}^q(X_{\ol{P}}, {\mathbb{Q}}_\ell)$ and $\mathfrak{F}_\ell:=\mathrm{R}^q f_*({\mathbb{Z}}_\ell)$ for every $\ell\in{\mathbb{L}}$. For the second case we define
$\rho_\ell:=\rho_{\ell, X, \mathrm{c}}^{(q)}$, $T_\ell:=\mathrm{H}^q_\mathrm{c}(X_{\ol{\xi}}, {\mathbb{Z}}_\ell)$, $V_\ell:=\mathrm{H}^q_\mathrm{c}(X_{\ol{\xi}}, {\mathbb{Q}}_\ell)$,
$T_{\ell, P}:=\mathrm{H}^q_\mathrm{c}(X_{\ol{P}}, {\mathbb{Z}}_\ell)$, $V_{\ell, P}:=\mathrm{H}^q_\mathrm{c}(X_{\ol{P}}, {\mathbb{Q}}_\ell)$ and $\mathfrak{F}_\ell:=\mathrm{R}^q f_!({\mathbb{Z}}_\ell)$ for every $\ell\in{\mathbb{L}}$.
In both cases $\rho_{\ell, P}$ will stand for the representation of ${\mathrm{Gal}}(\widetilde{k}/k(P))$ on $V_{\ell, P}$.
All residue characteristics of $S$ are zero. Hence
there is a dense open subscheme $U\subset S$ such that for every $\ell\in{\mathbb{L}}$ the
${\mathbb{Z}}_\ell$-sheaves $\mathrm{R}^q f_*({\mathbb{Z}}_\ell)|U$ and $\mathrm{R}^q f_!({\mathbb{Z}}_\ell)|U$ are
lisse and of formation compatible with any base change $U'\to U$ (cf. \cite[Corollaire 2.6]{illusie}, \cite[Th\'eore\`me 3.1.2]{katzlaumon} and \cite[Th\'eore\`me 3.3.2]{katzlaumon}). Considering the cartesian diagrams
$$\begin{xy}
\xymatrix{
X_{\ol{P}}\ar[r]\ar[d] & \tilde{k}\ar[d] & X_{\ol{\xi}}\ar[r]\ar[d] & \widetilde{K}\ar[d] \\
f^{-1}(U)\ar[r] & U & f^{-1}(U)\ar[r] & U
}\end{xy}$$
we can for every $\ell\in {\mathbb{L}}$ identify the stalks of $\mathfrak{F}_\ell$ by the following
base change isomorphisms
$$\mathfrak{F}_{\ell, \ol{P}}\cong T_{\ell, P}\ \ \mbox{and}\ \
\mathfrak{F}_{\ell, \ol{\xi}}\cong T_{\ell}.$$
The fact that the ${\mathbb{Z}}_\ell$-sheaves $\mathfrak{F}_\ell|U$ are lisse
implies that for every $\ell\in{\mathbb{L}}$ the representation $\rho_{\ell}$
factors through $\pi_1(U)$ and that
there is a cospecialization isomorphism $\mathfrak{F}_{\ell, \ol{\xi}}\cong
\mathfrak{F}_{\ell, \ol{P}}$. Putting these isomorphisms together and tensoring with
${\mathbb{Q}}_\ell$ we obtain a cospecialization isomorphism
$sp_{\ell}: V_\ell\cong V_{\ell, P}$
for every $\ell\in{\mathbb{L}}$. In order to take the Tate twists into account let
$\varepsilon_\ell: {\mathrm{Gal}}(\widetilde{K}/K)\to {\mathbb{Q}}_\ell^\times$ be
the cyclotomic character of ${\mathrm{Gal}}_K$ and by $\varepsilon_{\ell, P}: {\mathrm{Gal}}(\widetilde{k}/k(P))\to {\mathbb{Q}}_\ell^\times$
the cyclotomic character of ${\mathrm{Gal}}(\widetilde{k}/k(P))$. Let $d\in {\mathbb{Z}}$ and define $\rho_\ell(d):=\rho_\ell\otimes
\varepsilon_{\ell}^{\otimes d}$ and $\rho_{\ell, P}(d):=\rho_\ell\otimes
\varepsilon_{\ell, P}^{\otimes d}$.
The cospecialization isomorphism $sp_\ell$ fits into a commutative
diagram
$$\begin{xy}
\xymatrix{
{\mathrm{Gal}}(K_{S, \mathrm{nr}}/K)\ar[r]^{\rho_\ell(d)} & \mathrm{Aut}_{{\mathbb{Q}}_\ell}(V_\ell)\ar[dd] \\
D_{K_{S, \mathrm{nr}}/K}(P_{\mathrm{nr}})\ar[u]\ar[d] & \\
{\mathrm{Gal}}(\widetilde{k}/k(P))\ar[r]^{\rho_{\ell, P}(d)} & \mathrm{Aut}_{{\mathbb{Q}}_\ell}(V_{\ell, P})
}\end{xy}$$
for every $\ell\in{\mathbb{L}}$.
There is a constant $b\in {\mathbb{N}}$ such that for every $\ell\in{\mathbb{L}}$
the inequality $\dim(V_\ell)\le b$ holds true (cf. \cite[Corollaire 1.3]{illusie}).
Furthermore, if we donote the
torsion part of the finitely generated ${\mathbb{Z}}_\ell$-module
$T_\ell$ by $T_\ell'$, then $T_\ell/T_\ell'$ injects into $V_\ell$ and the representation $\rho_\ell(d)$
factors through $\mathrm{Aut}_{{\mathbb{Z}}_\ell}(T_\ell/T_\ell')$. Hence $\mathrm{im}(\rho_\ell(d))$ (and also
$\mathrm{im}(\rho_{\ell, P}(d))$) is isomorphic to
a closed subgroup of $\mathrm{GL}_b({\mathbb{Z}}_\ell)$ for every $\ell\in{\mathbb{L}}$. Hence the families
$(\rho_\ell(d))_{\ell\in{\mathbb{L}}}$ and
$(\rho_{\ell, P}(d))_{\ell\in{\mathbb{L}}}$ of representations of $\pi_1(U)$ satisfy assumption a) of
Theorem \ref{crit} (and condition (B) of \cite[p.~3]{bible}).
Now note that $X_P$ is a separated scheme of finite type over the number field $k:=k(P)$. For a place
$v$ of a number field we denote by $p_v$ its residue characteristic. There is a finite
extension $k'/k$ and a finite set $T$ of places of $k'$ such that the following holds true:
\begin{enumerate}
\item[(1)] For every place $v$ of $k'$ with $v\notin T$ and every $\ell\in {\mathbb{L}}\smallsetminus\{p_v\}$ the representation $\rho_{\ell, P}(d)$ is
unramified at $v$.
\item[(2)] For every $v\in T$, every extension $\hat{v}$ of $v$ to $\widetilde{k}$ and every $\ell\in {\mathbb{L}}\smallsetminus\{p_v\}$ the image of the inertia group $I_{\hat{v}}$ under the
representation $\rho_{\ell, P}(d)$ is a pro-$\ell$ group.
\end{enumerate}
This is shown for $d=0$ in \cite[Th\'eore\`me 4.3]{illusie}, and the case $d\neq 0$ follows
as well, because the cyclotomic character $\varepsilon_{\ell, P}$ is unramified at every place $v$ of $k'$
with $p_v\neq \ell$. Because the family $(\rho_{\ell, P}(d))_{\ell\in{\mathbb{L}}}$ satisfies
the condition (B) of \cite[p.~3]{bible} and conditions (1) and (2) and because $k$ is a number field,
Serre's theorem \cite[Th\'eore\`me 1]{bible} implies that the
family $(\rho_{\ell, P}(d))_{\ell\in{\mathbb{L}}}$ is almost independent.
Now the above diagram shows that there is an open subgroup $D'$ of $D_{K_{S, \mathrm{nr}}}(P_{\mathrm{un}})$ such that the restricted family $(\rho_{\ell}(d)|D')_{\ell\in{\mathbb{L}}}$ is independent,
and our Theorem \ref{crit} implies that $(\rho_{\ell}(d))_{\ell\in{\mathbb{L}}}$ is almost independent as desired.\hfill $\Box$
\label{cohomology}
|
\section{\label{sec:level1}Introduction}
Physical theories prescribe probability calculi for computing measurement outcome expectations and state assignments for physical systems.
The probability calculi prescribed by classical theories of physics are fundamentally different from the probability calculus prescribed by quantum mechanics. For instance, Birkhoff and von Neumann pointed out that classical experimental propositions regarding physical systems form Boolean algebras; whereas, quantum experimental propositions comprise nondistributive orthomodular lattices \cite*{vonNeumann36}. Moreover, Feynman emphasized that the classical Markovian law of probability composition fails to hold in the description of general quantum mechanical phenomena \cite*{Feynman48}\cite*{Adler95}. Instead, quantum \textit{probability amplitudes} superimpose. These essential features are not unique to the calculus prescribed by usual quantum mechanics as a theory over the complex field -- they are enjoyed in quantum theories formulated over any of the associative normed division rings $\mathbb{R}$, $\mathbb{C}$, or $\mathbb{H}$.\\
\indent What, then, \textit{does} distinguish quantum theories formulated over $\mathbb{R}$ or $\mathbb{H}$ from usual complex quantum mechanics (\textsc{cqm})? In the case of real quantum theory (\textsc{rqt}) \cite*{Stueckelberg60}, multipartite systems are endowed with some rather unusual properties. For example, in \textsc{rqt}, there exist states associated with $n$-partite systems for which every subsystem is maximally entangled with each of the other subsystems, where $n$ can be arbitrarily large \cite*{Wootters10}. Furthermore, \textsc{rqt} is not a locally tomographic theory -- it is instead a bilocally tomographic theory \cite*{Hardy10}. These observations point to aspects of \textsc{rqt} that cannot be realized within the usual \textsc{cqm} framework. However, the evolution and measurement of a multipartite complex quantum state under discrete or continuous evolution in \textsc{cqm} \textit{can} be simulated using states and operators in \textsc{rqt} \cite*{McKague09a}.\\
\indent In the case of quaternionic quantum theory (\textsc{qqt}) \cite*{Finkelstein59}\cite*{Finkelstein62}\cite*{Finkelstein63}, the very notion of `independent subsystems' is ill defined. In fact, quaternion-linear tensor products of quaternionic modules \textit{do not exist} \cite*{Razon91}. This constitutes a significant obstacle for the development of a consistent definition of local quaternionic operations, and it has been argued that one is actually prevented from speaking of absolutely independent systems in \textsc{qqt} \cite*{Finkelstein62}. These peculiar features may set \textsc{qqt} apart from usual \textsc{cqm}. Nevertheless, in the context of $1$-dimensional quantum wave mechanics, it has been shown that \textsc{cqm} is consistent with \textsc{qqt}. Specifically, one can recreate the entire structure of $1$-dimensional complex quantum wave mechanics inside $1$-dimensional quaternionic quantum wave mechanics \cite*{Nash92}. Conversely, the experimental propositions in \textsc{qqt} that commute with a fixed anti-Hermitian unitary operator are isomorphic to the experimental propositions of \textsc{cqm} \cite*{Finkelstein62}. In the context of quantum information processing involving unitary transformations and projective measurements, it has been shown that circuits acting on $n$ $2$-dimensional quaternionic systems can be simulated by circuits acting on $n+1$ qubits \cite*{Fernandez03}.\\
\indent In this paper, we consider a \textit{generalized} formulation of dynamics in \textsc{qqt}, rather than only considering the restricted class of quantum processes treated in \cite*{Fernandez03}. We treat generalized quaternionic quantum measurements as positive operator valued measures on quaternionic modules, and we treat quaternionic quantum channels as completely positive trace preserving quaternionic maps. Given an arbitrary $d$-dimensional quaternionic quantum state $\rho$ for a physical system $\mathfrak{S}$, we show that a generalized quaternionic quantum measurement $\mathfrak{M}_{\mathbb{H}}$ on $\mathfrak{S}$ can be simulated by a complex quantum measurement $\mathfrak{M_{\mathbb{C}}}$ on $\mathfrak{S}$ with an associated $2d$-dimensional complex quantum state $\sigma$. We also show that any quaternionic quantum channel can be simulated by a completely positive trace preserving map in \textsc{cqm}.\\
\indent The remainder of this paper is structured as follows. In section II, we review, for the reader's convenience, prerequisite material concerning quaternions and quaternionic linear algebra. In section III, we introduce a quaternionic quantum formalism, and we prove a Gleason-type theorem dictating the quaternionic Born rule for calculating probabilities for outcomes of generalized measurements in \textsc{qqt}. In section IV, we exhibit quaternionic quantum dynamics as complex quantum dynamics on complex Hilbert spaces. Finally, we conclude in Section V.
\section{\label{sec:level2}Quaternionic Algebra}
\subsection{\label{sec:level3}Quaternions}
The quaternions were first discovered by Hamilton \cite*{Hamilton44}. We express $h\in\mathbb{H}$ as $h=1h_{0}+ih_{1}+jh_{2}+kh_{3}$ in terms of its constituents $h_{r}\in\mathbb{R}\;\forall r\in\{0,1,2,3\}$, and the quaternion basis elements $\{1,i,j,k\}$, which obey
\begin{equation}
i^{2}=j^{2}=k^{2}=ijk=-1 \text{.}
\label{basisElements}
\end{equation}
$\mathbb{H}$ is an abelian group with respect to addition defined via $h+h^{'}=1(h_{0}+h^{'}_{0})+i(h_{1}+h^{'}_{1})+j(h_{2}+h^{'}_{2})+k(h_{3}+h^{'}_{3})$, and a monoid with respect to noncommutative multiplication defined via
\begin{eqnarray}
hh^{'} = & 1(h_{0}h^{'}_{0}-h_{1}h^{'}_{1}-h_{2}h^{'}_{2}-h_{3}h^{'}_{3})+ \nonumber \\
& i(h_{0}h^{'}_{1}+h_{1}h^{'}_{0}+h_{2}h^{'}_{3}-h_{3}h^{'}_{2})+ \nonumber \\
& j(h_{0}h^{'}_{2}+h_{2}h^{'}_{0}-h_{1}h^{'}_{3}+h_{3}h^{'}_{1})+ \nonumber \\
& k(h_{0}h^{'}_{3}+h_{3}h^{'}_{0}+h_{1}h^{'}_{2}-h_{2}h^{'}_{1})\text{,}\;\;\;
\end{eqnarray}
$\forall h,h^{'}\in\mathbb{H}$. Quaternion addition and multiplication are distributive in the sense that $h(h^{'}+h^{''})=hh^{'}+hh^{''}$, and $(h+h^{'})h^{''}=hh^{''}+h^{'}h^{''}$ $\forall h,h^{'},h^{''}\in\mathbb{H}$. The quaternionic conjugation operation $h \rightarrow \overline{h}$ taking $\{1,i,j,k\}\rightarrow\{1,-i,-j,-k\}$ is an involutory anti-automorphism inducing a multiplicative norm $|h|=(h\overline{h})^{\frac{1}{2}}$ on $\mathbb{H}$.
The compact symplectic group Sp(1) of unit-norm quaternions is isomorphic to SU(2), which can be seen from viewing $\varphi\in$ Sp(1) as $\varphi=\gamma_{1}+\gamma_{2}j$ in terms of \begin{eqnarray}
\gamma_{1}=1\varphi_{0}+i\varphi_{1}\in\mathbb{C} \label{symCorRep1} \text{,}\\
\gamma_{2}=1\varphi_{2}+i\varphi_{3}\in\mathbb{C} \text{,}
\label{symCorRep2}
\end{eqnarray}
so that $\varphi\overline{\varphi}=1\implies |\gamma_{1}|^{2}+|\gamma_{2}|^{2}=1$. Next, by defining $f:$ Sp(1) $\rightarrow$ SU(2) such that
\begin{equation}
f(\varphi)=\begin{bmatrix}\;\;\;\gamma_{1} &\gamma_{2} \vspace{0.5ex}\\ -\overline{\gamma_{2}} & \overline{\gamma_{1}}\end{bmatrix}\text{,}
\label{symRep}
\end{equation}
it is clear that $f$ is a bijection and that $f(\varphi_{1}\varphi_{2})=f(\varphi_{1})f(\varphi_{2})$, establishing that Sp(1) $\cong$ SU(2). More generally, one has that Sp(d) $\cong$ U(2d, $\mathbb{C}$) $\cap$ Sp(2d, $\mathbb{C}$), where Sp(d) $\cong$ U(d, $\mathbb{H}$) is the group of $d\times d$ unitary quaternionic matrices \cite*{Fulton91}. There are, however, subtle distinctions between quaternionic and complex matrix algebras due to the noncommutativity of quaternion multiplication.
\subsection{\label{sec:level4}Quaternionic Modules and Matrices}
For an excellent review of quaternionic linear algebra, we refer the reader to \cite*{Zhang97}. In this paper, we adopt the convention wherein the Cartesian product $\mathbb{H}^{d}$ is taken as a \textit{right} quaternionic module. We equip $\mathbb{H}^{d}$ with the standard symplectic inner product $\langle \cdot|\cdot\rangle:\mathbb{H}^{d}\rightarrow\mathbb{H}$ defined via $\langle \phi|\chi\rangle=\sum_{r=1}^{d}\overline{\phi_{r}}\chi_{r}$ $\forall \phi, \chi\in\mathbb{H}^{d}$, where $\phi_{r}$ and $\chi_{r}$ denote the projections of $\phi$ and $\chi$ onto elements of a basis for $\mathbb{H}^{d}$. Furthermore, we adopt the convention wherein linear operators on $\mathbb{H}^{d}$ act as elements of $M_{p,d}(\mathbb{H})$ -- the set of $p\times d$ quaternionic matrices -- from the \textit{left}, as usual. Stated explicitly, if $A\in M_{p,d}(\mathbb{H})$ with entries $[A]_{rs}$ and $\phi\in\mathbb{H}^{d}$, then our conventions imply that $A\phi$ is computed as
\begin{equation}
A\phi=\sum_{r=1}^{p}\sum_{s=1}^{d}\sum_{t=1}^{d}|r\rangle A_{rs} \langle s|t\rangle\phi_{t}=\sum_{r=1}^{p}\sum_{s=1}^{d}|r\rangle A_{rs}\phi_{s}\text{,}
\label{matrixMult}
\end{equation}
where the last equality in \eqref{matrixMult} follows in general if and only if $s$ and $t$ are elements of an orthonormal basis.
Following Finkelstein et al.\ \cite*{Finkelstein59}, we define the \textit{trace} of $A\in M_{d,d}(\mathbb{H})$ with respect to a basis $\Omega=\left\{\omega_{1},\omega_{2},\dots,\omega_{d}\right\}$ for $\mathbb{H}^{d}$ as $\mathrm{tr}(A)=\mathrm{Re}\left(\sum_{r=1}^{d}\langle\omega_{r}|A\omega_{r}\rangle\right)$. It follows that the trace of $A$ is independent of $\Omega$. It also follows that the cyclic property of the trace holds: $\mathrm{tr}(ABC)=\mathrm{tr}(CAB)$ $\forall A,B,C\in M_{d,d}(\mathbb{H})$. $M_{p,d}(\mathbb{H})$ admits an involution $^{*}$ defined such that $[A^{*}]_{rs}=\overline{[A]_{sr}}$. When $p=d$, $\langle\phi|A\chi\rangle=\langle A^{*}\phi|\chi\rangle$, and we denote the set of self-adjoint quaternionic matrices satisfying $A=A^{*}$ by $M_{d,d}(\mathbb{H})_{sa}$. Given our conventions, the spectral theorem holds for self-adjoint quaternionic matrices \cite*{Adler95}. As usual, we say that $A\in M_{d,d}(\mathbb{H})$ is \textit{unitary} when $AA^{*}=\mathds{1}_{\mathbb{H}^{d}}$, and we say that $A\in M_{d,d}(\mathbb{H})$ is \textit{positive semi-definite} when $\langle\phi|A\phi\rangle\ge0$ $\forall\phi\in\mathbb{H}^{d}$. We state without proof that positive semi-definiteness implies self-adjointness for elements of $M_{d,d}(\mathbb{H})$. We equip $M_{d,d}(\mathbb{H})_{sa}$ with the symmetric positive-definite $\mathbb{R}-$bilinear form $(\cdot,\cdot): M_{d,d}(\mathbb{H})_{sa}\times M_{d,d}(\mathbb{H})_{sa} \rightarrow \mathbb{R}$ defined via $(A,B)=\mathrm{tr}(AB)$. For the remainder of this paper we shall view $M_{d,d}(\mathbb{H})_{sa}$ as a real vector space. On that view, $\mathrm{tr}(AB)$ is an \textit{inner product} on the real vector space $M_{d,d}(\mathbb{H})_{sa}$ inducing the norm $|A|=\sqrt{\mathrm{tr}(A^{2})}$.
\subsection{\label{sec:level5}\texorpdfstring{Embedding $M_{p,d}(\mathbb{H})$ into $M_{2p,2d}(\mathbb{C})$}{Embedding Mp,d(H) into M2p,2d(C)}}
Let $A\in M_{p,d}(\mathbb{H})$ be a $p\times d$ quaternionic matrix with $A=\Gamma_{1}+\Gamma_{2}j$, where $\Gamma_{1},\Gamma_{2}\in M_{p,d}(\mathbb{C})$ are obtained by decomposing the matrix elements $[A]_{rs}$ according to \eqref{symCorRep1} and \eqref{symCorRep2}. In analogy with \eqref{symRep}, we define the embedding $\psi_{p,d}:M_{p,d}(\mathbb{H})\rightarrow M_{2p,2d}(\mathbb{C})$ via
\begin{equation}
\psi_{p,d}\left(A\right)=
\begin{bmatrix}
\;\;\;\Gamma_{1} & \Gamma_{2}\; \vspace{0.5ex}\\
-\overline{\Gamma_{2}} & \overline{\Gamma_{1}}\;
\end{bmatrix}
\text{.}
\label{embed}
\end{equation}
If $a,a'\in\mathbb{R}$, $A,A'\in M_{p,d}(\mathbb{H})$, and $B\in M_{d,q}(\mathbb{H})$, then it follows from \eqref{embed} that
\begin{equation}
\psi_{p,d}(aA'+a'A')=a\psi_{p,d}\left(A\right)+a'\psi_{p,d}\left(A'\right)\text{,}
\label{rLin}
\end{equation}
\begin{equation}
\psi_{p,d}\left(A\right)\psi_{d,q}\left(B\right)=\psi_{p,q}\left(AB\right) \text{,}
\label{homo}
\end{equation}
\begin{equation}
\psi_{d,p}\left(A^{*}\right)=\psi_{p,d}\left(A\right)^{*} \text{.}
\label{star}
\end{equation}
It is also readily verified that $\psi_{p,d}$ is an injection. Furthermore, when $p=d=q$, $\psi_{d,d}$ is the usual injective $^{*}$-homomorphism from $M_{d,d}(\mathbb{H})$ into $M_{2d,2d}(\mathbb{C})$ pointed out by Farenick and Pidkowich in \cite*{Farenick03}.
\section{\label{sec:level6}Quaternionic Quantum Formalism}
\subsection{\label{sec:level7}States, Evolution, and Measurement}
In \textsc{qqt}, a quantum state for a $d$-dimensional physical system is associated with a unit-trace positive semi-definite matrix $\rho\in M_{d,d}(\mathbb{H})_{sa}$. We will assume that the time-evolution of a quantum state $\rho$ is governed by a quantum channel $\Phi$ whose action is defined by a completely positive trace preserving quaternionic map \cite*{Kossakowski00}. On that assumption, we take $\Phi(\rho)=\sum_{r=1}^{n}A_{r}\rho A_{r}^{*}$, where $A_{r}\in M_{p,d}(\mathbb{H})$ are such that $\sum_{r=1}^{n}A_{r}A_{r}^{*}=\mathds{1}_{\mathbb{H}^{p}}$, and where $n\in \mathbb{Z}_{+}$. We will associate a quaternionic quantum measurement device $\mathfrak{M}_{\mathbb{H}}$ with a positive operator valued measure on $\mathbb{H}^{d}$ whose values are $\{E_{1},\dots,E_{m}\}$, $m\in\mathbb{Z}_{+}$, such that $E_{r}\in M_{d,d}(\mathbb{H})$ are positive semi-definite and $\sum_{r=1}^{m}E_{r}=\mathds{1}_{\mathbb{H}^{d}}$. Each $E_{r}$ corresponds to a measurement outcome that may occur with a probability given by the Born rule.
\subsection{\label{sec:level8}The Born Rule in QQT}
For dimension $d\ge3$, the Born rule for calculating probabilities for outcomes of projection valued measurements in usual \textsc{cqm} was derived by Gleason \cite*{Gleason57}. Gleason's result carries over to \textsc{rqt} and \textsc{qqt} \cite*{Buhagiar09}. Caves et al.\ extended Gleason's result in a noncontextual setting to cover quantum measurements associated with positive operator valued measures on complex Hilbert spaces for all dimensions $d\ge2$ \cite*{Caves04}.
We will now proceed to show that the result given by Caves et al.\ carries over to \textsc{qqt}. Let us denote by $\mathcal{E}(\mathbb{H}^{d})$ the set of all \textit{quaternionic quantum effects} -- that is, the set of all positive semi-definite linear operators $E$ on $\mathbb{H}^{d}$ admitting $\mathrm{tr}(E^{2})\le d$. We define a \textit{quaternionic frame function} as any map $f:\mathcal{E}(\mathbb{H}^{d})\rightarrow[0,1]$ satisfying
\begin{equation}
\sum_{E_{r}\in X}f(E_{r})=1\text{,}\;\forall X=\Big\{E_{r}\in\mathcal{E}(\mathbb{H}^{d})\;\Big|\;\sum_{r}E_{r}=\mathds{1}_{\mathbb{H}^{d}}\Big\}.
\end{equation}
For every frame function $f$, there exists a unique unit-trace positive semi-definite $\rho\in M_{d,d}(\mathbb{H})_{sa}$ such that
\begin{equation}
f(E)=(E,\rho)=\mathrm{tr}(E\rho).
\label{bornRule}
\end{equation}
This is the Born rule for calculating probabilities for outcomes of generalized measurements in \textsc{qqt}. For the proof, note that quaternionic quantum effects admit a spectral resolution in terms of real eigenvalues and orthogonal eigenprojectors. As a result, the proof given by Caves et al.\ in the complex case can almost literally be transfered to the quaternionic case, and we encourage the reader to consult \cite*{Caves04} for details. In particular, one can establish $\mathbb{R}-$linearity of $f$ on $M_{d,d}(\mathbb{H})_{sa}$. Now, let $\{\Upsilon_{1},\dots,\Upsilon_{d(2d-1)}\}$ be an orthonormal basis for $M_{d,d}(\mathbb{H})_{sa}$. We can expand any effect $E$ as a linear combination of the $\Upsilon_{r}$ in terms of coefficients $(\Upsilon_{r},E)$. Also, there exists a unique operator $\rho$ that we can expand as a linear combination of the $\Upsilon_{r}$ in terms of coefficients $f(\Upsilon_{r})$. It follows that $f(E)=(E,\rho)$. The operator $\rho$ is positive semi-definite, which is verified by letting $E=|\phi\rangle\langle\phi|$ for arbitrary $\phi\in\mathbb{H}^{d}$. We also have that $\rho$ is unit-trace, which follows from the observation that $\mathrm{tr}(\rho)=(\rho,\mathds{1}_{\mathbb{H}^{d}})=\left(\rho,\sum_{E_{r}\in X}E_{r}\right)=\sum_{E_{r}\in X}f(E_{r})=1$, finishing the proof. It is worth mentioning that these arguments would fail to hold if we had used the standard Hilbert-Schmidt inner product on $M_{d,d}(\mathbb{H})_{sa}$, which is not real-valued in general.
\section{\label{sec:level9}Complex Simulations of Quaternionic Quantum Dynamics}
\subsection{\label{sec:level0}Inner Product Correspondence}
Before we show that quaternionic quantum dynamics can be simulated by complex quantum dynamics, it will be useful to establish the following correspondence between our inner product on $M_{d,d}(\mathbb{H})_{sa}$ and the usual Hilbert-Schmidt inner product on $M_{2d,2d}(\mathbb{C})_{sa}$:
\begin{equation}
\mathrm{tr}(AB) = \textstyle{\frac{1}{2}}\mathrm{tr}\Big(\psi_{d,d}(A)\psi_{d,d}(B)\Big)\;\;\forall A,B \in M_{d,d}(\mathbb{H})_{sa}\text{.}
\label{claim}
\end{equation}
For the proof, we expand $A=\Gamma_{1}+\Gamma_{2}j$ and $B=\Lambda_{1}+\Lambda_{2}j$ in terms of complex self-adjoint $\Gamma_{1}=\Gamma_{1}^{*}$ and $\Lambda_{1}=\Lambda_{1}^{*}$, and complex antisymmetric $\Gamma_{2}=-\Gamma_{2}^{\mathrm{T}}$ and $\Lambda_{2}=-\Lambda_{2}^{\mathrm{T}}$. Expanding the LHS of \eqref{claim} we get
\begin{eqnarray}
& \underbrace{\textstyle{\frac{1}{2}}\mathrm{tr}\big(\Gamma_{1}\Lambda_{1}+\Lambda_{1}\Gamma_{1}\big)}_{\mathtt{\alpha}}+\underbrace{\textstyle{\frac{1}{2}}\mathrm{tr}\big(\Gamma_{2}j\Lambda_{2}j+\Lambda_{2}j\Gamma_{2}j\big)}_{\mathtt{\beta}} \nonumber \\
& + \underbrace{\textstyle{\frac{1}{2}}\mathrm{tr}\big(\Gamma_{1}\Lambda_{2}j+\Lambda_{1}\Gamma_{2}j+\Gamma_{2}j\Lambda_{1}+\Lambda_{2}j\Gamma_{1}\big)}_{\mathtt{\delta}}\text{,}
\label{LHS}
\end{eqnarray}
whereas expanding the RHS of \eqref{claim} we get
\begin{equation}
\underbrace{\textstyle{\frac{1}{2}}\mathrm{tr}\left(\Gamma_{1}\Lambda_{1}+\overline{\Gamma_{1}}\;\overline{\Lambda_{1}}\right)}_{\mathtt{\alpha}'}+\underbrace{\textstyle{\frac{1}{2}}\mathrm{tr}\left(-\Gamma_{2}\overline{\Lambda_{2}}-\overline{\Gamma_{2}}\Lambda_{2}\right)}_{\mathtt{\beta}'}\text{.}
\label{RHS}
\end{equation}
It is not hard to see that $\mathtt{\alpha}=\mathtt{\alpha}'$, $\mathtt{\beta}=\mathtt{\beta}'$, and $\mathtt{\delta}=0$. We have defined the trace operation so that it is basis-independent, and so, for simplicity, we can compute $\mathtt{\alpha}$, $\mathtt{\beta}$, and $\mathtt{\delta}$ in terms of the standard orthonormal basis $\{e_{1},\dots,e_{d}\}$ admitting $e_{r}-\overline{e_{r}}=0$ $\forall r\in\{1,\dots,d\}$. On that view, one immediately sees that $\mathtt{\alpha}=\mathtt{\alpha}'$. Next, we observe that $j\Lambda_{2}j=-\overline{\Lambda_{2}}$ and $j\Gamma_{2}j=-\overline{\Gamma_{2}}$. Therefore $\mathtt{\beta}=\mathtt{\beta}'$. Finally, we observe that $j\Lambda_{1}=\overline{\Lambda_{1}}j$ and $j\Gamma_{1}=\overline{\Gamma_{1}}j$, and after some algebra one finds that $\mathtt{\delta}=0$, finishing the proof.
\subsection{\label{sec:level11}Simulating Generalized Measurements}
Equipped with \eqref{claim}, we are now ready to prove that generalized measurements in \textsc{qqt} can be simulated by usual quantum measurements in \textsc{cqm} with positive operator valued measures on complex Hilbert spaces. Let $\rho\in M_{d,d}(\mathbb{H})_{sa}$ be a quaternionic quantum state for a physical system $\mathfrak{S}$, and let $\mathfrak{M}_{\mathbb{H}}=\{E_{1},\dots ,E_{m}\}\subseteq\mathcal{E}(\mathbb{H}^{d})$ define a generalized quaternionic quantum measurement with outcome probabilities $p(r)=\mathrm{tr}(E_{r}\rho)$. Then, there exists a complex quantum state $\sigma(\rho)=\textstyle{\frac{1}{2}}\psi_{d,d}(\rho)\in M_{2d,2d}(\mathbb{C})_{sa}$ and a positive operator valued measure $\mathfrak{M}_{\mathbb{C}}=\{\psi_{d,d}(E_{1}),\dots ,\psi_{d,d}(E_{m})\}\subseteq\mathcal{E}(\mathbb{C}^{2d})$ on complex Hilbert space with outcome probabilities $q(r)=\mathrm{tr}\big(\psi_{d,d}(E_{r})\sigma(\rho)\big)$, such that $\forall r$: $q(r)=p(r)$.
For the proof, we begin by showing that $\psi_{d,d}$ preserves positive semi-definiteness. The spectral decomposition of positive semi-definite $\rho\in M_{d,d}(\mathbb{H})_{sa}$ is given by $\rho=\sum_{r=1}^{d}|\xi_{r}\rangle\lambda_{r}\langle\xi_{r}|$ in terms of $\lambda_{r}\in\mathbb{R}_{+}$ and eigenprojectors $\Xi_{r}=|\xi_{r}\rangle\langle\xi_{r}|$. We have that $\psi_{d,d}$ is $\mathbb{R}-$linear from \eqref{rLin}, and from \eqref{homo} it is clear that $\psi_{d,d}$ maps projections on $\mathbb{H}^{d}$ to projections on $\mathbb{C}^{2d}$. Thus, $\psi_{d,d}(\rho)$ is a positive semi-definite operator on complex Hilbert space. Next, we define positive semi-definite $\sigma(\rho)=\textstyle{\frac{1}{2}}\psi_{d,d}(\rho)$, and by \eqref{claim} we have that $\sigma(\rho)$ is unit-trace. Therefore, $\sigma(\rho)$ is a valid complex quantum state. Also, from the definition of $\psi_{d,d}$ it follows that $\psi_{d,d}(\mathds{1}_{\mathbb{H}^{d}})=\mathds{1}_{\mathbb{C}^{2d}}$, and applying $\mathbb{R}-$linearity of $\psi_{d,d}$ once again, it follows that $\mathfrak{M}_{\mathbb{C}}=\{\psi_{d,d}(E_{1}),\dots,\psi_{d,d}(E_{m})\}$ is a valid positive operator valued measure on complex Hilbert space. Finally, applying the quaternionic Born rule \eqref{bornRule} and using \eqref{claim} we see that $\forall r$:
\begin{equation}
p(r)=\mathrm{tr}(E_{r}\rho)=\mathrm{tr}\big(\psi_{d,d}(E_{r})\sigma(\rho)\big)=q(r)\text{,}
\end{equation}
finishing the proof.
\subsection{\label{sec:level12}Simulating Quantum Channels}
In this section, we prove that quaternionic quantum channels can be simulated by completely positive trace preserving maps in usual \textsc{cqm}. Let $\rho\in M_{d,d}(\mathbb{H})_{sa}$ be a quaternionic quantum state for a physical system $\mathfrak{S}$, and let $\Phi:M_{d,d}(\mathbb{H})_{sa}\rightarrow M_{p,p}(\mathbb{H})_{sa}$ be a quaternionic quantum channel whose action is defined via
\begin{equation}
\Phi(\rho)=\sum_{r=1}^{n}A_{r}\rho A_{r}^{*}\text{,}
\end{equation}
where $A_{r}\in M_{p,d}(\mathbb{H})$ and $\sum_{r=1}^{n}A_{r}A_{r}^{*}=\mathds{1}_{\mathbb{H}^{p}}$. Then, there exists a complex quantum channel $\Theta$ whose action on $\sigma(\rho)$ is defined via
\begin{equation}
\Theta(\sigma(\rho))=\sum_{r=1}^{n}\psi_{p,d}(A_{r})\sigma(\rho)\psi_{p,d}(A_{r})^{*}\text{,}
\end{equation}
and given an arbitrary quaternionic quantum measurement defined by $\mathfrak{M}_{\mathbb{H}}=\{E_{1},\dots ,E_{m}\}\subseteq\mathcal{E}(\mathbb{H}^{d})$ one has that $\forall r$:
\begin{equation}
\mathrm{tr}\Big(E_{r}\Phi\big(\rho\big)\Big)=\mathrm{tr}\Big(\psi_{p,p}\big(E_{r}\big)\Theta\big(\sigma(\rho)\big)\Big)\text{,}
\label{equiv}
\end{equation}
\\
where $\mathfrak{M}_{\mathbb{C}}=\{\psi_{p,p}(E_{1}),\dots,\psi_{p,p}(E_{m})\}$ is a positive operator valued measure on complex Hilbert space. Put otherwise, any generalized preparation $\rightarrow$ transformation $\rightarrow$ measurement process in \textsc{qqt} corresponds to an algorithm in usual complex quantum information theory.
For the proof, note that \eqref{rLin}, \eqref{homo}, and \eqref{star} imply that $\sum_{r}\psi_{p,d}(A_{r})\psi_{p,d}(A_{r})^{*}=\mathds{1}_{\mathbb{C}^{2p}}$, so $\Theta$ is a valid complex quantum channel. We have already established that $\sigma(\rho)$ is a valid complex quantum state, and so it follows that $\Theta(\sigma(\rho))$ is a valid complex quantum state. Now, again using \eqref{rLin}, \eqref{homo}, and \eqref{star} we have that
\begin{equation}
\Theta(\sigma(\rho))=\textstyle{\frac{1}{2}}\psi_{p,p}\Big(\sum_{r}A_{r}\rho A_{r}^{*}\Big)\text{,}
\end{equation}
and so by \eqref{claim} we see that \eqref{equiv} holds, finishing the proof.
\section{\label{sec:level13}Conclusion}
Ultimately, one would like to use the developing technologies of quantum information science to test the validity of usual \textsc{cqm} versus \textsc{qqt} \cite*{Peres79}. Before one can perform such tests, however, one must have a clear conception of the relations and contrasts between these two theories with respect to the full apparatus of quantum information theory, not just the projective measurements and unitary operations considered in \cite*{Fernandez03}. This paper fills that gap in the literature. We have shown that \textit{all} generalized quantum dynamics in \textsc{qqt} can be realized as usual quantum dynamics in \textsc{cqm}. In particular, we have shown that generalized measurements associated with positive operator valued measures on quaternionic modules can be simulated by usual quantum measurements in \textsc{cqm}. Furthermore, we have shown that quaternionic quantum channels can be simulated by completely positive trace preserving maps in \textsc{cqm}. These results offer a new vantage point to view quaternionic quantum algorithms from \textit{inside} usual complex quantum information theory.
\section{\label{sec:level14}Acknowledgments}
The author thanks Chris Fuchs for his guidance and support. The author also thanks Howard Barnum for discussions on Jordan algebras, and \AA sa Ericsson for feedback on the manuscript. This work was supported in part by the U. S. Office of Naval Research (Grant No. N00014-09-1-0247), and by the province of Ontario through OGS.
|
\section*{Introduction}
The theory of $\SU(2)$ representations and re-coupling is known for quite a long time. The main motivation was the quantization of angular momenta, with obvious applications in spectroscopy, atomic/molecular physics. Nevertheless, the study of $\SU(2)$ re-coupling coefficients remains an active field of research in modern physics \cite{spinnets-marzuoli}. One reason is the appearance of a new notion to describe some phases of matter: topological order \cite{wen-top-orders}. Its emergence has been described as a condensation of spin networks \cite{levin-wen-condensation} (possibly with a quantum group deformation). It also gives a new way to think of fault-tolerant quantum computations \cite{dennis-top-memory}.
In the same time, a similar formalism has led to interests in re-coupling coefficients to capture aspects of quantum geometry and possibly quantum gravity. The main idea in that approach is that there is no fixed background geometry, but geometry is instead a fully dynamical object. Spin foam models have emerged as direct applications of that idea (see e.g. \cite{SFreview} for a review). A useful model to explore the spin foam program is the Ponzano-Regge model \cite{PR, PR1} which aims at defining the path integral for three-dimensional pure gravity. The latter is a topological field theory: there are no local degrees of freedom due to a large set of specific gauge symmetries. Thus some interests in that model also comes from topological invariants, and especially quantum invariants since a quantum deformation of the Ponzano-Regge model gives the well-defined Turaev-Viro model \cite{TV}.
These models are constructed by triangulating the space-time manifold and defining the corresponding amplitude as a product of local amplitudes attached to each tetrahedron of the triangulation. The basic building block of the Ponzano-Regge model associated to each tetrahedron is the Wigner 6j-symbol of the $\SU(2)$ re-coupling theory.
The 6j-symbol is certainly the most studied Wigner coefficient, and we refer to \cite{qm6j, quantum-tet-marzuoli} for modern reviews describing numerous aspects and to \cite{Varshalovich} for textbook results. An important regime of such a quantum object is the semi-classical limit, when angular momenta become large. The asymptotics of the 6j-symbol has been studied and proved in \cite{SG,roberts, integral,razvan6j}, and next-to-leading orders have been obtained in \cite{NLO_maite, 6jnlo, recursion_maite}. The asymptotics of the Ponzano-Regge model on handlebodies (with more than one tetrahedron) is presented in \cite{dowdall-handlebodies}. Recent works have also focused on more complicated objects, like the 15j-symbol and EPR amplitude in \cite{barrett-asym15j}. An extended Born-Oppenheimer approximation has been developped and applied to the asymptotics of the 9j, 12j and 15j-symbols with some large and small angular momenta \cite{Yu_and_Littlejohn}. One of the author has also studied this regime for arbitrary 3nj-symbols in \cite{3nj-small} with a simpler method.
An efficient way to study such asymptotics as well as algebraic properties of re-coupling coefficients is the use of recursion relations on the angular momenta. It is actually a way to define the 6j-symbol \cite{SG, Varshalovich}, and some insights into asymptotics and recursions on arbitrary 3nj-symbols are given in \cite{3nj-marzuoli}. From the point of view of topological field theory, recursion relations have been understood as a way to encode (some features of) the dynamics, i.e. characterizing the invariance under gauge symmetries \cite{recursion_simone, valentin2}. All these considerations actually also apply to the four-dimensional case \cite{valentin1} and spin foam models for 4d quantum gravity. In this case, we do not yet have the symmetries and the dynamics under full control and we strongly think that the development of recursion relations for the spin foam amplitudes will be a very useful tool to understand further the structure and invariance of 4d spin foams.
There exist several ways to derive recursion relations for the 6j-symbol. The original method is to start from the Biedenharn-Elliott (BE) pentagon identity. The latter encodes at the algebraic level a topological invariance of the Ponzano-Regge model under a special move. It is an equality between the product of two 6j-symbols (the weight for two tetrahedra glued along a triangle) and a sum of products of three 6j-symbols (which correspond to three tetrahedra sharing an edge)\footnote{This states the invariance under the 3-2 Pachner move. Topological invariance also requires invariance under the 4-1 move which is actually induced from the Biedenharn-Elliott identity together with the orthogonality relation of 6j-symbols. But it is only formal due to divergences \cite{PR_john, twisted-homology}.}. Recursion relations are obtained by specializing the BE identity with well-chosen boundary angular momenta \cite{SG}
Another path towards recursion relations is to think of Wigner symbols as evaluations of spin network states. Then, one can interpret recursion relations as Wheeler-DeWitt equations in the spin network basis, generated by some Hamiltonian operators which annihilate the evaluations. Such an operator is expected to be the generator of a gauge symmetry \cite{recursion_simone,valentin1,valentin2}. That is a way to see how recursion relations encode and provide the spin network states of Loop quantum gravity with their dynamics. The standard recursion on the 6j-symbol was derived in that way in \cite{valentin2}, where the Hamiltonian operators form a first-class algebra classically, which means that they really generate a gauge symmetry.
The final method which we would like to report on here is to start with the expression of the 6j-symbol as an integral over four copies of $\SU(2)$ (see e.g. \cite{integral}). That group integral approach was already proposed in an earlier study \cite{recursion_simone}. In this previous work, the method could only be applied to a special class of isosceles 6j-symbols (and to the 10j-symbol of the Barrett-Crane model for 4d Euclidean quantum gravity). The idea is that measures in such group integrals have some specific properties which can be used to get recursion formulae. Further, those measures have a geometric nature which eases the interpretation in the asymptotics.
However, the case of generic 6j-symbols has a more complicated measure. Still, it is possible to extract recursion formulae using similar techniques, as we show in the present paper. The organization is the following. In the Section \ref{sec:groupint}, we present the integral which is our starting point. In the Section \ref{sec:newrec}, we prove a new recursion, acting on the square of the 6j-symbol, whose asymptotics and semi-classical interpretation are discussed in the Section \ref{sec:semiclass}. Finally we open in the Section \ref{sec:towards} the interesting possibility of deriving recursion relations in a systematic fashion for arbitrary spin foam amplitudes in the Ponzano-Regge model, as opposed to a single tetrahedron.
\section{6j-Symbol as a Group Integral} \label{sec:groupint}
Let us consider a tetrahedron with its six edges labeled by $i=1,\dotsc,6$ and its four vertices labeled by $a=1,\dotsc,4$. We attach an irreducible representation of $\SU(2)$ to each edge, i.e we associate a spin $j_i\in \N/2$ to the edge $i$. Equivalently, we will refer to these spins as $j_{(ab)}$ where the (symmetric) pair of vertices $a,b$ uniquely defines the edge $i$. At each vertex of the tetrahedron graph, we associate the unique (up to normalization) 3-valent intertwiner between the representations living on the three edges attached to that vertex. The intertwiner is the invariant vector in the tensor product of the three representations meeting at the vertex. Intertwiners are given explicitly by the Clebsch-Gordan coefficients, or more precisely by the Wigner 3jm-symbols. Finally, tensoring these four intertwiners living on the four vertices of the graph and taking the trace on magnetic indices, we obtain the 6j-symbol.
The square of the 6j-symbol can be expressed as an integral over the group $\SU(2)$, using the integral expression of the products of 3jm-symbols \cite{Varshalovich},
\begin{equation}
\int dg\, \prod_{i=1}^3D^{(j_i)}_{a_ib_i}(g)
\,=\,\begin{pmatrix} j_1 &j_2 &j_3\\a_1 &a_2 &a_3\end{pmatrix} \begin{pmatrix} j_1 &j_2 &j_3\\b_1 &b_2 &b_3\end{pmatrix}\;,
\end{equation}
where $D^{(j_i)}(g)$ is the Wigner matrix of $g\in\SU(2)$ in the representation of spin $j_i$ and the quantities into brackets are the 3jm-symbols.
This way, we get one integral over the group for each vertex of the graph. Taking the trace leads to an expression for the square of the 6j-symbol as an integral over $\SU(2)^4$ (see e.g. \cite{integral} for more details),
\be \label{int6jsquare}
\{6j\}^2 \,\equiv\,\begin{Bmatrix} j_{12} &j_{13} &j_{14}\\ j_{34} &j_{24} &j_{23}\end{Bmatrix}^2
\,=\,
\int_{\SU(2)^4} [dg_{a}]^{4}\,
\prod_{(ab)} \chi_{j_{(ab)}}(g_ag_b^{-1})\;,
\ee
where the integral is taken with the Haar measure $[dg]$ over $\SU(2)$ and the character $\chi_j(g)=\tr_j D^{(j)}(g)$ is the trace of the Wigner matrix $D^{(j)}(g)$. This is the starting point of our analysis.
Since characters are invariant under conjugation by $\SU(2)$ group elements, we can write the above expression as an integral over the class angles $\theta_{ab}$ of the group elements $g_ag_b^{-1}$. The change of measure was computed in \cite{integral},
\be \label{angleint}
\{6j\}^2
\,=\,
\f2{\pi^4}\int_{D} [d\theta_{ab}]^{6}\,
\f1{\sqrt{\det_{4\times 4} (\cos\theta_{ab})}}\,
\prod_{(ab)} \sin (2j_{ab}+1)\theta_{ab}\;,
\ee
where we take the convention that $\theta_{aa}=0$ and thus $\cos\theta_{aa}=1$. The measure is given in term of the determinant of a $4\times 4$ Gram matrix $(\cos\theta_{ab})$. It has a simple geometrical interpretation: ${\sqrt{\det (\cos\theta_{ab})}}$ gives the volume of the (dual) tetrahedron on the sphere $\cS^3$, whose spherical edge lengths are the angles $\theta_{ab}$. The above formula can be used to extract the asymptotic behavior of the 6j-symbol at large spin through a saddle point analysis \cite{integral}.
The domain of integration $D$ is the set of all possible spherical tetrahedra, i.e. the subset of $[0,\pi]^6$ satisfying $\theta_{ac}\leq \theta_{ab}+\theta_{bc}$ and $\theta_{ab}+\theta_{bc}+\theta_{ca}\leq 2\pi$.
Since it will be our departure point, let us sketch the derivation of \eqref{angleint}. Among the four group variables in \eqref{int6jsquare}, we can absorb, say, $g_4$ in a re-definition of the others, $g_a\mapsto g_a g_4^{-1}$ for $a=1,2,3$. Clearly, that does not change the arguments of the characters. Moreover, due to the right invariance of the Haar measure, $dg_1, dg_2, dg_3$ are unchanged, and the integral over $g_4$ becomes trivial. Then we write each group element $g_a = \exp(i \theta_{a4}\, \hat{n}_{a}\cdot\vec{\sigma})$, where $\theta_{a4}$ is the class angle, and $\hat{n}_a\in\cS^2$ the axis of the rotation, and the measure reads
\be
dg_a=\frac{2}{\pi}\ \sin^2\theta_{a4} d\theta_{a4}\,d^2\hat{n}_a\;.
\ee
The class angles are kept as integration variables since they are the arguments of the characters with spins $j_{a4}$. As for the six variables $(\hat{n}_a)_{a=1,2,3}$, we split them into two sets. One consists in the dot products $u_{ab}\equiv \hat{n}_a\cdot\hat{n}_b$, and the other set form a global rotation (three components) which leaves the variables $u_{ab}$ invariant and trivially factorizes. Finally, we notice that the arguments of the characters with spins $(j_{ab})$ with $a,b\neq 4$ are class angles, which we denote $\theta_{ab}$, which are related to the variables $(u_{ab})$ by
\be
\cos \theta_{ab} = \cos\theta_{a4}\,\cos\theta_{b4} + \sin\theta_{a4}\,\sin\theta_{b4}\,u_{ab}\;.
\ee
The final step is to change the variables from $u_{ab}$ to $\theta_{ab}$. The Jacobian is straightforward to evaluate and leads to \eqref{angleint}.
\section{New Recursion} \label{sec:newrec}
Let us introduce the following notations,
\be
P[X_{ab}]\equiv \det(X_{ab}),\quad \text{and}\qquad T^\pm_{ab} f(j_{ab}) = f(j_{ab}\pm\tfrac12)\;.
\ee
$P$ is the determinant of a collection of numbers $(X_{ab})$, and is polynomial of degree four in the six variables $X_{ab}=X_{ba}$, $X_{aa}=1$. The operators $T^\pm_{ab}$ shift the spin $j_{ab}$ for a given pair $(ab)$. The new recursion we are going to present is
\be \label{rec6jsquare}
\left[4\,
P[\tfrac12(T^+_{ab}+T^-_{ab})]\,
\left(
(2j_{\alpha\beta}+2)\,T^+_{\alpha\beta}\,-\,2j_{\alpha\beta}\,T^-_{\alpha\beta}
\right)
+ \pp_{X_{\alpha\beta}}P[\tfrac12(T^+_{ab}+T^-_{ab})]\,(T^+_{\alpha\beta}-T^-_{\alpha\beta})^2
\right]\
\{6j\}^2=0 \;,
\ee
for each symmetric pair $(\alpha\beta)$ of vertices, with $\alpha,\beta=1,2,3,4$ and $\alpha\neq\beta$.
Our method is fairly simple and works exactly like Schwinger-Dyson equations in field theory. We multiply the integrand of \eqref{angleint} with a well chosen insertion and take a total derivative and integrate over all angles.
Singling out one edge $(\alpha\beta)$, we consider the following integral
\be
\int [d\theta_{ab}]^{6}\
\pp_{\theta_{\alpha\beta}}
\Bigl[
\,\sin\theta_{\alpha\beta}\,\sqrt{\det_{4\times 4} (\cos\theta_{ab})}\,
\prod_{(ab)} \sin (2j_{ab}+1)\theta_{ab}
\Bigr]
\,=\,
0\;,
\ee
which is indeed zero since the factor $\det(\cos\theta_{ab})$ vanishes on the boundary of the domain $D$ \cite{integral}.
To expand this integral, we use the notation $P[X_{ab}]= \det_{4\times 4} X_{ab}$. The measure term is given by its evaluation on $X_{ab}=\cos\theta_{ab}$. We compute explicitly the derivative in the previous expression and distinguish the two terms where we differentiate the Gram matrix determinant or the other terms,
\beq
0&=&
\int [d\theta_{ab}]^{6}\,\,
\Bigg{[}\f{\det (\cos\theta_{ab})}{\sqrt{\det (\cos\theta_{ab})}}\,
\Bigl(
\cos\theta_{\alpha\beta} \,\sin(2j_{\alpha\beta}+1)\theta_{\alpha\beta}
+(2j_{\alpha\beta}+1) \,\sin\theta_{\alpha\beta} \,\cos(2j_{\alpha\beta}+1)\theta_{\alpha\beta}
\Bigr) \nn\\
&&-\,\f{\pp_{X_{\alpha\beta}}P[X_{ab}]|_{X=\cos\theta}}{2\sqrt{\det (\cos\theta_{ab})}}
\,\sin^2\theta_{\alpha\beta} \,\sin(2j_{\alpha\beta}+1)\theta_{\alpha\beta}\,\Bigg{]}
\prod_{(ab)\ne(\alpha\beta)} \sin (2j_{ab}+1)\theta_{ab}\;,
\eeq
which can be re-written as
\beq
0&=&
\int [d\theta_{ab}]^{6}\,\,
\Bigg{[}\f{\det (\cos\theta_{ab})}{\sqrt{\det (\cos\theta_{ab})}}\,
\Bigl(
(2j_{\alpha\beta}+2)\,\sin(2j_{\alpha\beta}+2)\theta_{\alpha\beta}
-2j_{\alpha\beta}\,\sin 2j_{\alpha\beta}\theta_{\alpha\beta}
\Bigr) \nn\\
&&-\,\f{\pp_{X_{\alpha\beta}}P[X_{ab}]|_{X=\cos\theta}}{\sqrt{\det (\cos\theta_{ab})}}\,
(1-\cos^2\theta_{\alpha\beta}) \,\sin(2j_{\alpha\beta}+1)\theta_{\alpha\beta}\,\Bigg{]}
\prod_{(ab)\ne(\alpha\beta)} \sin (2j_{ab}+1)\theta_{ab}\;.
\eeq
Now all the $\cos\theta$-factors can be re-absorbed as shifts in the spin labels $j_{ab}$. Indeed, the multiplication by a factor $\cos\theta_{ab}$ corresponds to the operator $\f12(T^+_{ab}+T^-_{ab})$ due to the trigonometric identity
\be \label{trigo}
\cos\theta_{ab}\ \sin (2j_{ab}+1)\theta_{ab}
\,=\,
\frac{\sin (2j_{ab}+2)\theta_{ab}
\,+\,
\sin (2j_{ab})\theta_{ab}}2 \,=\, \frac{T^+_{ab}+T^-_{ab}}{2}\ \sin (2j_{ab}+1)\theta_{ab} \,.
\ee
That leads to the following recursion relation on the squared 6j-symbol
\be \nonumber
\left[4\,
P[\tfrac12(T^+_{ab}+T^-_{ab})]\,
\left(
(2j_{\alpha\beta}+2)\,T^+_{\alpha\beta}\,-\,2j_{\alpha\beta}\,T^-_{\alpha\beta}
\right)
+ \pp_{X_{\alpha\beta}}P[\tfrac12(T^+_{ab}+T^-_{ab})]\,(T^+_{\alpha\beta}-T^-_{\alpha\beta})^2
\right]\
\{6j\}^2=0 \;,
\ee
where we used that $T^+_{\alpha\beta}T^-_{\alpha\beta}=T^-_{\alpha\beta}T^+_{\alpha\beta}=\id$. This is \eqref{rec6jsquare}. The non-triviality of the relation is ensured by the presence of the factors $(2j+2)$ and $2j$ in the first line, which implies that the coefficients of this equation can not all vanish.
\medskip
To better understand the explicit structure of this relation, we need to expand the Gram determinant,
\be
P[X_{ab}]=\sum_{\sigma\in{\cal S}_4}\epsilon(\sigma)\,\prod_a^4X_{a\sigma(a)}\;,
\ee
where we sum over all permutations of 4 elements. We classify them by their number of fixed points. The only permutation that fixes all elements is the identity. Then we get the two-point cycles, which exchange two elements without touching the other two elements. These contribute as $-X_{ab}^2$ to the determinant, where $a$ and $b$ are the two elements which are interchanged. Then permutations with a single fixed point are given by the 3-cycles, e.g. $1\arr 2\arr 4\arr 1$ leaving $3$ invariant. These give terms of order 3. Finally, we have the six 4-cycles, which lead to term of order 4 in the polynomial. More explicitly, we can compute
\beq
P[X] &=& 1- X^2_{12}- X^2_{13}- X^2_{14}- X^2_{23}- X^2_{24}- X^2_{34} \nn\\
&& +2X_{23}X_{24}X_{34}+2X_{13}X_{34}X_{14}+2X_{12}X_{24}X_{14}+2X_{12}X_{23}X_{13}\nn\\
&&+X_{12}^2X_{34}^2+X_{13}^2X_{24}^2+X_{14}^2X_{23}^2\nn\\
&&-2X_{12}X_{23}X_{34}X_{14}-2X_{13}X_{34}X_{24}X_{12}-2X_{14}X_{24}X_{23}X_{13}\;,
\eeq
with $4!=24$ terms. We also compute its derivative with respect to, say, $X_{12}$,
\be
\pp_{X_{12}}P
\,=\,
2\left(X_{12}(X_{34}^2-1)+X_{23}X_{13}+X_{24}X_{14}-X_{23}X_{14}X_{34}-X_{24}X_{13}X_{34}
\right)\;.
\ee
In the end, we obtain a recursion relation which involves mixed shifts on the spins $j_{ab}$ up to five $\pm\f12$-shifts.
\section{Asymptotics and Semi-Classical Geometry} \label{sec:semiclass}
\subsection{The Recursion as a Closure Condition}
The most well-known (and useful) recursion relation on the 6j-symbol is a second order difference equation \cite{Varshalovich},
\be \label{rec6j}
A_{+1}[j_i]\,\begin{Bmatrix} j_1+1 &j_2 &j_3 \\ j_4 &j_5 &j_6 \end{Bmatrix} +
A_{0}[j_i]\,\begin{Bmatrix} j_1 &j_2 &j_3 \\ j_4 &j_5 &j_6 \end{Bmatrix} +
A_{-1}[j_i]\,\begin{Bmatrix} j_1-1 &j_2 &j_3 \\ j_4 &j_5 &j_6 \end{Bmatrix} = 0,
\ee
and there also exists a similar second order recursion with $\pm \f12$-shifts. The explicit coefficients can be found in \cite{Varshalovich,SG} for instance. This equation can be used to derive all 6j-symbols from an initial condition, and is thus used numerically to generate lists of values of the symbol. This means that in principle the recursion relation \eqref{rec6jsquare} could be obtained from this one. This seems however to be quite complicated.
Still it is interesting to compare both:
\begin{itemize}
\item the standard relation only has three terms, but complicated coefficients,
\item while our new recursion relation has many more terms (corresponding to the various combinations of shifts), but all coefficients are trivial (either $\pm1$ or $\pm(2j+2)$, or $\pm 2j$).
\end{itemize}
Hence it certainly can not be used in efficient numerical simulations, but in contrast its geometrical interpretation at leading order is straightforward, as we will now explain.
In geometric terms, the recursion \eqref{rec6j} has been interpreted in \cite{recursion_simone} as the statement of invariance under an elementary deformation on a tetrahedron whose result is a small displacement of a vertex. This simply comes out of interpreting some specialization of the Biedenharn-Elliott pentagon identity \cite{Varshalovich} in the 3-2 Pachner move. From this point of view, the recursion \eqref{rec6jsquare} should certainly be seen as a non-trivial combination of such elementary moves, successively performed. The question is then what geometric properties are encoded into such combination of moves. While the standard recursion \eqref{rec6j} encodes the full symmetry of the model (topological invariance), the new recursion should only probe some simpler geometric properties of the 6j-symbol. To extract this information, it is far easier and more fruitful to look at its asymptotic form. At the leading order, this is just the closure of the corresponding tetrahedron.
The asymptotics of the 6j-symbol when all spins are homogeneously scaled is\footnote{Note that it can be efficiently derived from \eqref{rec6jsquare}.} (see \cite{roberts,PR,SG,integral})
\be
\{6j\}\sim\frac{\cos\bigl(S_R[j_i]+\f\pi 4\bigr)}{\sqrt{12\pi V[j_i]}}\;.
\ee
Here, $V[j_i]$ is the volume of the tetrahedron with edge lengths $j_i+\f12$ and $S_R[j_i]=\sum_i (j_i+\f12)\phi_i[j_i]$ is the Regge action for the same tetrahedron. The angles $\phi_i[j_i]$ are the dihedral angles of the tetrahedron computed as functions of the edge lengths. In particular, they satisfy the following property,
\be
\det_{4\times 4}(\cos \phi_{ab}) = 0\;.
\ee
Mathematically, it means exactly that the angles $\phi_{ab}$ are dihedral angles for some tetrahedron. Equivalently, we will refer to that condition as the \emph{closure} of a tetrahedron. This is part of the information on the $\phi_{ab}[j_i]$. The other information that is needed to recover the asymptotics is that they are the dihedral angles for the tetrahedron whose lengths are $(j_{ab}+\frac12)$. Those two conditions are the roots of the first order formulation of Regge calculus, where lengths and angles are independent variables \cite{first-order-regge}. They can be both extracted from the recursion \eqref{rec6j}, \cite{SG}.
Squaring the asymptotic approximation, we easily get
\be
\{6j\}^2\sim\f1{24\pi V[j_i]}\,\left(1-\sin\bigl(2\,S_R[j_i]\bigr)\right),\quad \text{where}\qquad
2S_R[j_i]=\sum_i (2j_i+1)\phi_i[j_i]\;.
\ee
This leading order asymptotics goes as $j^{-3}$ as we scale the spins homogeneously.
Let us insert it in the new recursion relation and check that it solves it at leading order in the large spin limit. Ignoring the volume pre-factor (which only contributes to higher orders), we have two terms in the asymptotics: the constant term $1$ and the oscillatory term $\sin(2S_R[j_i])$.
The constant term trivially solves the recursion relation because of the perfect balance of plus and minus signs in the equation. More precisely, both the shifts due to the Gram determinant (since $\sum_\sigma \epsilon(\sigma)=0$) and due to the $(1-\cos^2\theta)$ factor (and actually also the shifts due to $\pp_X P$) vanish.
The non-trivial check thus comes from the oscillatory term. Let us look at the behavior of $\sin 2S_R$ when spins are shifted. By applying basic trigonometry, we easily see that
\beq
(T^+_{ab}+T^-_{ab})\ \sin (2S_R[j_i])
&=& 2\cos\phi_{ab}\,\sin (2S_R[j_i])\;,\nn\\
(T^+_{ab}-T^-_{ab})\ \sin (2S_R[j_i])
&=& 2\sin\phi_{ab}\,\cos(2S_R[j_i])\;.
\eeq
Keeping only the leading contribution to the recursion relation, i.e. only the terms with an explicit $j_{\alpha\beta}$, we see that it holds if and only if
\be
P[\tfrac12(T^+_{ab}+T^-_{ab})]\,\bigl(T^+_{\alpha\beta}-T^-_{\alpha\beta}\bigr)\,\sin(2S_R[j_i])\propto
\sin\phi_{\alpha\beta}\,\det_{4\times 4}\cos\phi_{ab}\ \cos(2S_R[j_i])
\,=\, 0\;.
\ee
This is obviously true due to the closure of the tetrahedron, which directly implies the vanishing of the Gram determinant of the geometrical dihedral angles, $\det_{4\times 4}\cos\phi_{ab}=0$.
Equivalently, if we try to find solutions to the new recursion \`a la WKB, we find that for any oscillations going like $\exp(i\sum j_{ab}\, \omega_{ab}[j_i])$, the angles $\omega_{ab}$ are constrained to be dihedral angles for some tetrahedron, whatever they are as functions of the spins.
If we want to push the analysis to higher orders, we have to include the next-to-leading order the 6j-symbol (e.g. \cite{recursion_maite,NLO_maite}) and the corrections due to the volume pre-factor. The geometrical interpretation of the higher order corrections to the 6j-symbol itself is not clear at all, so it might be enlightening to push this analysis and see if our new recursion relation can provide a natural geometrical meaning. The hope is that our recursion does not encode the full symmetries of the 6j-symbol but focuses on its closure which maybe partially characterize the next-to-leading orders of the 6j-symbol.
\subsection{Comparison with other Closure Equations}
\subsubsection{Asymptotic Equations} \label{sec:asym}
One of the key idea in loop quantum gravity and spin foam models is to make sense of quantum geometry using $\SU(2)$ re-coupling. From this view point, it is expected that the relevant $\SU(2)$ objects satisfy such closure equations.
So far, it has been mainly investigated in the context of models for four-dimensional quantum geometry. It has been recently shown in \cite{valentin1} that closure relations generically holds asymptotically in the large spin limit when the leading order oscillates with the Regge action of the 4-simplex.
Consider a $d$-simplex (with $d=3,4$) with its $(d-2)$-subsimplices labeled by spins $(j_{ab})$. Interpreting the latter as the volumes of the $(d-2)$-simplices, we assume we are in a regime where a classical geometry of the simplex with those volumes is available. This assumption is obviously satisfied for the 6j-symbol, but is more subtle in four dimensions since ten areas do not always determine a flat 4-simplex. It can be reached nevertheless by using an overcomplete basis of intertwiners which typically provides the normals to the triangles in addition to their areas. The assumption then implies that dihedral angles are well-defined and depend on the spins. They may also depend on the normals but we can neglect their variations to leading order.
If the amplitude goes like $W[j_{ab}]\sim A\,\exp(iS_R[j_{ab}])$, then it satisfies the following equation for each spin $j_{ab}$, \cite{SG, valentin1},
\be \label{regge-criterion}
\left((T^+_{ab})^2+(T^-_{ab})^2\right)\,W[j_{ab}] \,=\, 2\,\cos\phi_{ab}\,W[j_{ab}]\;.
\ee
It means that multiplication by the cosine of a dihedral angle can be compensated by shifts on the corresponding spin of $\pm 1$. Then taking the antisymmetrized sum over all permutations of products of that equation, we form on the right hand side an identically vanishing factor, $\det(\cos \phi_{ab}) =0$, so that
\be \label{asymclosure}
P[(T^+_{ab})^2+(T^-_{ab})^2]\ W[j_{ab}]\,=\,0\;.
\ee
That provides a definition of a closure operator.
\subsubsection{Exact Equations}
The above relation only holds asymptotically, for the 6j-symbol, the 15j-symbol and more complicated objects \cite{valentin1}. Hence getting an \emph{exact} closure relation for the 6j-symbol is an interesting achievement.
The technical difference between \eqref{asymclosure} and the asymptotic form of the exact equation \eqref{rec6jsquare} is the following. The operator $(T^+_{ab})^2+(T^-_{ab})^2$ which is the argument of $P$ in \eqref{asymclosure} generates shifts of $\pm1$ while in \eqref{rec6jsquare}, the argument of $P$ is $\frac12 (T^+_{ab}+T^-_{ab})$ which generates shifts of $\pm\frac12$. That difference is compensated since $P$ acts on the 6j-symbol itself in \eqref{asymclosure} but on its square in the exact recursion. Note also that the exact recursion picks up additional terms in the sub-leading orders, due to the derivatives of $P$ in \eqref{rec6jsquare}.
The new recursion \eqref{rec6jsquare} is actually the second known example of an exact relation whose leading asymptotic behavior reduces to the closure of a simplex. The first example we have found is the 10j-symbol. The similarities of the derivation with the present situation strengthens the potential generality of the method. Hence we briefly sketch a few steps of that derivation which was reported with details in \cite{recursion_simone}.
The 10j-symbol is a quantum weight associated to a 4-simplex in a model for four-dimensional geometry known as the Barrett-Crane spin foam model \cite{BC}. It is a function of ten spins $j_{(ab)}$ living on the 4-simplex boundary graph (the complete graph with 5 vertices, where each vertex is linked to the other four vertices by a single edge) and its asymptotics displays oscillations with the Regge action of the 4-simplex \cite{asym10j}. It is simply expressed as an integral over five copies of $\SU(2)$,
\be \label{10jgroup}
\{10j\}
\,=\,
\int_{\SU(2)^5} [dg_{a}]^{4}\,
\prod_{(ab)} \chi_{j_{ab}}(g_ag_b^{-1})\;.
\ee
Once again, we can write this as an integral over the class angles of the group elements $g_ag_b^{-1}$ and the structure of the 10j-symbol is actually simpler than that of the squared 6j-symbol \cite{integral}:
\be \label{10jangles}
\{10j\}
\,=\,
\f4{\pi^6}\int_0^{\pi} [d\theta_{ab}]^{10}\,
\delta(\det_{5\times 5} (\cos\theta_{ab}))\,
\prod_{(ab)} \sin (2j_{ab}+1)\theta_{ab}\;,
\ee
where we simply have to impose that the $5\times 5$ Gram matrix has a vanishing determinant. This constraint states that the ten angles $\theta_{ab}$ are the dihedral angles of a true geometric, flat 4-simplex.
To get a recursion, we insert the determinant of the Gram matrix in the integral, which then vanishes due to the constraint in the measure term,
\be
\int [d\theta_{ab}]^{10}\
\delta(\det_{5\times 5} (\cos\theta_{ab}))\,
\det_{5\times 5} (\cos\theta_{ab})\,
\prod_{(ab)} \sin (2j_{ab}+1)\theta_{ab}
\,=\,
0\;.
\ee
The new factor $\det_{5\times 5} (\cos\theta_{ab})$ is a polynomial in the variables $\cos\theta_{ab}$. Each cosine actually produces a $\pm\f12$-shift in the $j_{ab}$'s spins labeling the 10j-symbol, due to the trigonometric identity \eqref{trigo}.
In \cite{recursion_simone}, it was checked that amplitudes going like $\exp(i\sum_{(ab)} (2j_{ab}+1)\phi_{ab})$, i.e. which oscillates with the Regge action of the 4-simplex whose triangle areas have the values\footnote{The 10j-symbol has only ten quantum numbers which do not always determine a 4-simplex. But it is known that there is a regime where it does \cite{asym10j}.} $(j_{ab})$, do indeed satisfy the above equation. That is because, not very surprisingly, it reduces to $\det(\cos\phi_{ab})$ in the asymptotics.
The surprising fact about that equation on the 10j-symbol is that it is exact although the Regge oscillations are definitely not the dominant contributions to the asymptotics \cite{nloBC}. This is in contrast with the case of the 6j-symbol we have been studying here. The 10j-symbol is thus a good example to illustrate the difference between the reasoning of the Section \ref{sec:asym} in the large spin limit and the exact treatment (when available). In particular, we do not expect the 10j-symbol to satisfy the equation \eqref{regge-criterion} at the leading order in the asymptotics.
\section{Towards Recursion Relations for Spinfoam Amplitudes} \label{sec:towards}
The previous sections were dedicated to the new recursion from the point of view of the 6j-symbol itself. Here we take a different perspective and try to sketch some implications for the Ponzano-Regge model \cite{PR,PR1}.
An important difference between the standard recursion \eqref{rec6j} and the new one \eqref{rec6jsquare} that we want to emphasize is that the latter acts on the square of the 6j-symbol, as opposed to the 6j-symbol itself. While the 6j-symbol is the building block of the Ponzano-Regge model, which is attached to each tetrahedron of a triangulation, its square is the amplitude for the 3-sphere triangulated by two tetrahedra which are identified along their boundary,
\be
Z_{\rm PR}(\cS^3) = \int \prod_{a=1}^4 dg_a \ \prod_{a<b} \delta(g_a\,g_b^{-1}) = \sum_{\{j_{ab}\in\frac{N}{2}\}_{a<b}} \Bigl[\prod_{a<b} (2j_{ab}+1)\Bigr]\ \begin{Bmatrix} j_{12} &j_{13} &j_{14}\\ j_{34} &j_{24} &j_{23}\end{Bmatrix}^2\;.
\ee
This equation is only formal since it is actually divergent (there is one redundant Dirac delta in the product over all $a<b$). Making sense of that expression normally involves a gauge-fixing process, described in \cite{PR1}, which amounts to setting one spin in the summand to zero.
Our new recursion allows a different analysis (which we unfortunately cannot make complete). Note that the standard recursion on the 6j-symbol \eqref{rec6j} cannot be used in the above formula, since it requires to transform independently each symbol. Hence it is not a consistent symmetry of the amplitude. The new recursion, however, is directly a symmetry of the summand.
The usual recursion \eqref{rec6j} encodes the gauge invariance of the model. From the spacetime point of view, that comes from specializing the Biedenharn-Elliott identity which states the invariance under the 2-3 Pachner move. In the Hamiltonian framework, it is a quantization of a generator of a gauge symmetry which enables to solve for the wave-function on the boundary of a tetrahedron \cite{valentin2}. So it may be tempting to interpret the new recursion as being also related to that gauge symmetry. However, that is unclear to us, and we have looked at the algebra of the recursions acting on different spins without finding pieces of evidence.
Instead, the new recursion can be interpreted as a \emph{global} symmetry of the summand. By global, we mean that it acts on all spins, none of them being left unchanged. Together with the result of the previous section, that gives the following picture. The shifts on the squared 6j-symbol correspond to a deformation of the triangulation of the 3-sphere equipped with a discrete metric (given by the arguments of the symbol). Invariance under that deformation implies that the metric has to close the boundaries of the tetrahedra (which are identified). Thus, it provides with a consistent metric deformation of the triangulation.
\medskip
We would like to propose to develop this perspective and write recursion relations for more general triangulations. Let us start with a triangulation $\Delta$ of the 3-sphere. We consider a connected graph $\Gamma$ on $\Delta$, and fix all the spins $j_e$ on the edges $e\in\Gamma$ to fixed values. Looking at the Ponzano-Regge amplitude for that triangulation as a function of the spins $(j_e)_{e\in\Gamma}$, we would like to derive a recursion relation under shifts of the spins $(j_e)_{e\in\Gamma}$. To approach the problem, it has been shown in \cite{pr3} that we can write this amplitude as a group integral.
We start by choosing a framing for the graph $\Gamma$, i.e. we embed it faithfully in a closed surface in $\Delta$. More precisely we choose a set of triangles forming a closed surface $\Sigma$, such that the edges of $\Gamma$ are all in $\Sigma$. In general, $\Gamma$ is not a triangulation of the whole surface $\Sigma$, but still defines a cellular decomposition of $\Sigma$ whose faces are delimited by the edges of $\Gamma$. We require that the faces are all homeomorphic to the 2-disk. Then the result proved in \cite{pr3} is that for a trivial topology\footnote{Reference \cite{pr3} also deals with the more complicated case of a non-trivial surface topology. With a non-trivial topology defined by the genus $g$ of the surface $\Sigma$, we need to associate group elements $U_C\in\SU(2)$ to each cycle $C$ of the surface. Assuming that each edge $e\in\Gamma$ belongs to a single cycle $C(e)$ (or to none), we have
\be
Z_\Delta[j_{e\in\Gamma}]=\int [\prod_fdg_f]\,\int [\prod_C dU_C]\,
\prod_e \chi_{j_e}(g_{s(e)}U_{C(e)}^{\epsilon_C(e)}g_{t(e)}^{-1})\;,
\ee
where the sign $\epsilon_C(e)$ records the relative orientation of the edge $e$ with respect to the orientation of the cycle.} $\Sigma\sim\cS^2$, the amplitude has a simple expression, with group elements $(g_f)$ attached to the faces of $\Sigma$,
\be
Z_\Delta[j_{e\in\Gamma}]=\int [\prod_fdg_f]\,
\prod_e \chi_{j_e}(g_{s(e)}g_{t(e)}^{-1})\;.
\ee
Here $s(e)$ and $t(e)$ denote the two faces of $\Sigma$ sharing the edge $e$. Since characters are invariant under taking the inverse of the argument, it does not matter which face we choose as the source or the target for each edge.
Written as such, this amplitude is the generalization of the integral for the squared 6j-symbol \eqref{int6jsquare}. If we look at the graph dual to $\Gamma$ in $\Sigma$, we see that the group elements $(g_f)$ are actually attached to the vertices of the dual graph. Thus it is the square of the evaluation of the spin network state dual to the graph $\Gamma$ on the surface $\Sigma$.
The next step is to re-express the measure using the class angles $\theta_e$ of $g_{s(e)}g_{t(e)}^{-1}$ as basic variables,
\be
Z_\Delta[j_{e\in\Gamma}]=\int [d\theta_{e}]^{\vert \Gamma\vert}\ \mu(\theta_{e})\,\prod_{e\in\Gamma} \sin\left(2j_e+1\right)\,\theta_e\;,
\ee
where $\mu$ is the Jacobian of the change of variables. Its analysis has been started in \cite{Feyndiag} where it has been proven that $\mu$ involves a product of Gram determinants similarly to the case of the squared 6j-symbol. However, the squared 6j-symbol is a special case where all the angular variables $\theta_e$ are independent. Generically that is not true because the elements $g_f$ are unit vectors in $\R^4$, so that only four of them can be linearly independent. Typically, the integral \eqref{10jgroup} for the 10j-symbol contains five copies of $\SU(2)$ which implies that one among the ten angles is determined by the others. That is the reason which leads $\mu$ to be a constraint in \eqref{10jangles}.
In the absence of constraints, one can use the method developed in the present paper to get recursion formulae. In the presence of a constrained measure $\mu$, it is in principle possible to also apply the method which was used on the 10j-symbol, provided that the constraints in $\mu$ are polynomials in $(\cos\theta_e)$. However, the geometric content of $\mu$ seems to depend on the connectivity matrix of $\Gamma$, so that at the end of the day, such recursion formulae lack a global, unified geometric interpretation.
\section{Outlook}
In this short paper, we have adapted a method outlined in \cite{recursion_simone} to derive recursion relations for the square of the 6j-symbol using its expression as a multiple integral over $\SU(2)$. This method gives a very easy way to obtain recursion relations, which has a simple algebraic origin in the expression of the integral measure. Moreover, we have shown how it is simply related at leading order in the semi-classical limit to the closure of the tetrahedron whose quantum edge lengths are given by the spins labeling the 6j-symbol. Finally, we have explained how our method can be generalized to derive recursion relation for arbitrary spin foam amplitudes in the Ponzano-Regge model. We believe that this opens the door to a few issues and possibilities, which we hope to study in future work:
\begin{itemize}
\item Since the leading order of the recursion relation has a simple geometrical interpretation as the closure of the corresponding classical tetrahedron, we can hope that the next-to-leading order of our new recursion relation have a similar straightforward interpretation. It may also be that its analysis is technically less involved that when using the standard recursion as done in \cite{recursion_maite}.
\item It would be interesting to pursue this approach and further study the measure of the group integral for arbitrary spin foam amplitudes, in order to understand what geometric properties can emerge.
\item The existence of recursion relations in spin foam amplitudes is symptomatic of the existence of a symmetry. The standard recursion relation is usually related to the topological invariance of the Ponzano-Regge model under Pachner moves. Since our recursion relation can be interpreted as deformations of the space-time triangulation, it would be interesting to investigate its relation to discrete diffeomorphisms.
\item The specific invariance of 3d gravity and more generally BF theory in higher dimensions has been recently revisited \cite{PR_john, twisted-homology, matteo3}. In these works, one considers deformation properties of the full amplitude, using the notion of flat discrete connections, but not in the spin representation. That would be highly valuable to understand how the different objects used in these two different approaches actually probe the same geometric features.
\end{itemize}
\section*{Acknowledgments}
EL is partially supported by the ANR ``Programme Blanc" grants LQG-09.\\
Research at Perimeter Institute is supported by the Government of Canada through Industry Canada and by the Province of Ontario through the Ministry of Research and Innovation.
|
\section{Introduction}
In its present shape, our Universe appears to be a large ensemble of bound structures. These are the results of an
intricate net of chemical and physical processes of increasing complexity, the first of which is fragmentation
of the primordial gas clouds, see e.g. \cite{b15}.
One of the fundamental ingredient of this ensemble of processes
is the cooling of pristine plasma. Indeed, the birth of the first objects during the evolution of the primordial gas could not have
occured without significant cooling, as the gas temperatures involved in the
initial phase of the Universe are too high to allow the early clouds to gravitationally collapse. Among the
plasma constituents, molecules play the most significant role as coolers of the pristine gas.
Molecular cooling takes place mainly via rotational-vibrational (ro-vibrational) transitions between
the internal, nuclear motion degrees of freedom.
For temperatures below about 8000~K molecular cooling processes are much more efficient
than the corresponding atomic ones, see e.g. \cite{b18bis}.
In models, the standard way of capturing the effects of cooling by a given species is via the {\it cooling function}, which
gives the energy lost per second at a specified temperature.
Cooling functions are made up by two main contributions. First, molecules spontaneously emit radiation while making
ro-vibrational transitions; second, collision induced emission can occur; this is particularly important for high symmetry species where spontaneous emission can be very weak.
In a plasma, two key features render a molecule an efficient cooler at low temperature: its fractional abundance and its dipole moment.
In the chemistry of the primordial Universe H$_2$ represents the most abundant molecular species. Its role
as an efficient primordial coolant has been widely described, e.g. in
the formation of the first stars (e.g. Abel et al.~2002) and galaxies (e.g. Bromm et al.~2009, Benson~2010).
Both collisional and radiative H$_2$ cooling contributions have been considered in the literature. The role of
different collisional partners (H$_2$, He, H$^+$ and $e^-$) was considered by Glover \& Abel~(2008).
Together with H$_2$, the cooling properties of its cation H$_2^+$ due both to radiative and collisional
(H and $e^-$ impact excitation) pathways have been described (Suchkov \& Shchekinov~1978).
Both H$_2$ and H$_2^+$ are customarily included in molecular cooling models because of their high relative fractional abundance.
However, because of their lack of a permanent dipole moment,
only weak electric-quadrupole or magnetic-dipole transitions are allowed in these species, severly limiting their cooling efficiency, particularly at low density.
Considering the mole fractions of chemical species given by
chemical networks such as the one by Galli \& Palla (1998) (hereafter, GP98),
other molecular species should be added to the list of coolants.
Grouping the primordial molecules as deuterated or helium-containing or
lithium enriched, HD, HD$^+$, HeH$^+$, LiH and LiH$^+$ represent the candidates to focus on.
Their fractional abundances are expected to have the following freeze-out values ($z\approx 10$):
$f$(HD)~$\simeq 10^{-9}$; $f$(HD$^+$)~$\simeq 10^{-18}$; $f$(HeH$^+$)~$\simeq 10^{-12}$; $f$(LiH)~$\simeq 10^{-19}$;
$f$(LiH)$^+$~$\simeq 10^{-17}$. It is hard to form triatomic molecules
at high $z$, but the most likely species are H$_3^+$ and H$_2$D$^+$;
their fractional
abundances are thought to be approximately $f$(H$_3^+$)~$\simeq 10^{-18}$ and
$f$(H$_2$D$^+$)$^+$~$\simeq 10^{-19}$, respectively.
Collisional contributions to the cooling
have been evaluated in difLferent density regimes for
HD ad LiH (\cite{b3}, GP98).
In particular, the
effect of the most abundant collision partner H on this cooling pathway has been investigated.
However, in order to complete this description, the contribution of radiative cooling has to be taken into account as well.
The present work is organized as follows. In Section~\ref{conti} the equations used to compute the radiative
cooling function are presented, describing the methods adopted for each molecule. Some details on previous calculations on
collisional cooling functions are given. Section~\ref{conclusioni} summarizes our results and provides fits to the calculated data.
\section[]{Radiative cooling function: definition and calculations}\label{conti}
The total cooling function for the chemical species X is usually approximated as (e.g. Hollenbach \& McKee 1979, GP98):
\begin{equation}
\Lambda_\mathrm{X}=\Lambda_\mathrm{X,LTE}\left( 1+\frac{n^{\mathrm{cr}}}{n\mathrm{(collider)}}\right) ^{-1}
\label{generalcooling}
\end{equation}
where $\Lambda_\mathrm{X,LTE}$ represents the local thermodynamic equilibrium radiative contribution to the cooling function, $n^\mathrm{cr}$ and $n\mathrm{(collider)}$ are the critical
and collider density, respectively. The former is defined by:
\begin{equation}
n^{\mathrm{cr}}=\left(\frac{\Lambda_\mathrm{X,LTE}}{\Lambda_\mathrm{X}(n\mathrm{(collider)})}n\mathrm{(collider)}\right)_{n\mathrm{(collider)}\rightarrow 0}
\label{ncr}
\end{equation}
where $\Lambda_\mathrm{X}(n\mathrm{(collider)}\rightarrow 0)$ is the low-density limit of the cooling function.
Eq.~(\ref{generalcooling}) is valid for all density regimes of the collider particle.
In the high-density limit $\Lambda_\mathrm{X}$ reduces to $\Lambda_\mathrm{X,LTE}$,
implying that in this regime local thermal equilibrium is reached and that cooling is
dominated by ro-vibrational transitions of the molecular species X.
Denoting with the letters ``$u$'' and ``$l$'' repectively the upper and lower ro-vibrational
states between which a spontaneous transition may occur, the
radiative cooling function is defined as:
\begin{equation}
W(T)=\frac{1}{Z(T)}\sum_{u,l} A_{ul}(E_u-E_l)(2J_u-1) g_u n_u
\label{w}
\end{equation}
where $Z(T)$
represents the partition function of the molecule, $A_{ul}$ the Einstein coefficient
for the $u\rightarrow l$ transition, $J_u$ the rotational angular momentum
of the upper rotational state, $g_u$ the nuclear spin degeneracy (equal to 1
for all calculations below),
$E_u$ and $E_l$ the energies of the upper and lower state, and $n_u$ the population of the upper level.
As discussed above, under LTE conditions it holds:
\begin{equation} \begin{split}\label{wlte}
W(T) & \equiv \Lambda_\mathrm{LTE}(T)\\
n_u & = e^{-E_u/(k_B T)}\\
Z(T) & =\sum_{i} (2J_i+1) \, g_i \, e^{-E_i/(k_B T)}
\end{split}
\end{equation}
Radiative cooling functions have been computed in the present work using the equations above.
As $k_B^{-1} = 1.43878$~cm~K, the Boltzmann factor $n_u$
implies that at a temperature of $1000$~K the cooling function is mostly determined by
transitions with $E_u \lesssim 700$~cm$^{-1}$.
Data on the emission probabilities and energy levels of the molecular
species under analysis are therefore required in order to compute $\Lambda_{\mathrm{LTE}}$.
Depending on the available data, different strategies have been adopted for each molecular system considered.
For HD and HeH$^+$, the relevant data have been calculated respectively by Abgrall et al.~(1982) and Engel et al.~(2005).
To obtain the necessary data for the other diatomic molecules under investigation the program {\sc level} 8.0 by Le Roy (2007) was used.
This
program solves the radial one-dimensional Schr\"{o}dinger equation
\begin{equation}\begin{split}\label{EdS}
&-\frac{\hslash^2}{2\mu}\frac{\mathrm{d}^2\Psi(r)}{\mathrm{d}r^2}+V_J(r)\Psi(r)=E_{v,J}\Psi(r)\\
& V_J(r) =V(r) + \frac{\hbar^2 [J (J+1)-\Omega^2]}{2 \mu r^2}
\end{split}\end{equation}
where $\mu$ is the reduced mass of the system, $J$ the total rotational angular moment, $\Omega$
the projection of the electronic angular momentum along the internuclear axis ($\Omega=0$ for the states of our interest) and
$V(r)$ is the diatomic potential energy curve, given as input.
Once the ro-vibrational wavefunctions $\Psi_{\nu,J}$ have been obtained {\sc level} can also calculate
the Einstein coefficients by
\begin{equation}\begin{split}\label{formula.Einstein}
& A_{\nu'J'\rightarrow \nu"J"} = \\
& \left( \frac{16 \pi^3}{3 \varepsilon_0 \hbar} \right) \frac{S(J',J'')}{2J'+1} \, \tilde{\nu}^3 \mid\langle\psi_{\nu',J'}|M(r)|\psi_{\nu",J"}\rangle\mid^2
\end{split}\end{equation}
where
$S(J',J'')$ are
the H\"{o}nl-London rotational intensity factors, resulting from integration over the rotational degrees of freedom,
$\tilde{\nu}$ is the wavenumber of the transition, $M(r)$ is the diatomic dipole moment curve and is given as input.
In the common case where $\tilde{\nu}$ is measured in cm$^{-1}$ and $M$ in Debye the constant in brackets in Eq.~(\ref{formula.Einstein}) assumes the value $3.1361891\times10^{-7}$~cm$^3$ D$^{-2}$ s$^{-1}$.
Calculations were carried out for the electronic ground state of the molecular systems under investigation.
The following subsections examine each molecule considered and give detailed information on the potential energy and
dipole moment curves used to perform the calculations. Comparisons with existing data are also provided where possible.
\subsection{HD}
The role of HD in cooling phenomena of primordial plasma has been widely examinated in several studies (Flower et al. 2000, Lipovka et al. 2005, Ripamonti 2007). The effect of HD cooling on the formation of massive primordial stars was analysed by Yoshida et al.~(2007) and McGreer \& Bryan~(2008).
Despite its lower fractional abundance,
HD is a more efficient coolant than molecular hydrogen H$_2$, especially for $T<2000~$K (GP98).
This is because HD molecules possess a small permanent dipole moment ($\approx 8.3\times 10^{-4}$ D, Abgrall et al.~1982)
due to the asymmetry of the nuclei; by contrast, in non-deuterated molecular hydrogen
only very weak quadrupole transitions are allowed.
Lipovka et al. (2005) provide a two-parameter fit of the collisional cooling function, depending on temperature and
H fractional abundance. Here we evaluate
a radiative cooling function which takes into account both dipole and quadrupole electric transitions, for which we used the
calculations by Abgrall et al.~(1982).
Results are shown in Fig.~\ref{HD}, and compared with the high-density limit HD cooling function fit provided by Lipovka et al.~(2005),
where the contribution of the lowest ro-vibrational levels up to $v=3$ were considered.
A deviation from the high-density limit collisional contribution to cooling function is found for
temperatures above $700~$K, where highly excited ro-vibrational states are expected to be populated.
\begin{figure}
\includegraphics[width=0.5\textwidth]{f1.eps}
\caption{HD radiative cooling function. {\it Circles}: dipole contribution, present calculation;
{\it blue dotted-dashed curve}: quadrupole contribution;
{\it red dashed curve}: high-density limit fit by Lipovka et al. 2005;
{\it green solid curve}: fit given in Table~\ref{fits}.}
\label{HD}
\end{figure}
\subsection{HD$^+$}
The HD cation and its collisional contribution to the cooling during the evolution of primeval plasma
has been examined by Glover \& Abel (2008), where calculations were
carried out assuming that there were no significant differences between the molecular ions H$_2^+$ and HD$^+$
in the high-density cooling limit.
These authors provide an analytical expression for $\Lambda_{\mathrm{LTE}}$, calculated using
molecular data from Karr \& Hilico~(2006) on HD$^+$ (energy levels) and Peek et al.~(1979) on H$_2^+$ (quadrupole moments).
Our calculations of the HD$^+$ radiative cooling function were carried out using the potential
energy and dipole moment curves provided by Esry \& Sadeghpour~(1999), in which an adiabatic reformulation of the HD$^+$
Hamiltonian that recovers the isotopic splitting of electronic
states is presented. The resulting cooling curve is shown in Fig.~\ref{HD+},
and compared with the fit by Glover \& Abel~(2008). It should be noted that
our calculations predict significantly increased cooling at temperatures
below 100~K, where the cooling by HD$^+$ rotational transitions is very
much more efficient than the one due to its H$_2^+$ parent ion.
\begin{figure}
\includegraphics[width=0.5\textwidth]{f2.eps}
\caption{HD$^+$ radiative cooling function: {\it Circles}: present calculation; {\it red dashed curve}: fit by Glover \& Abel 2008;
{\it green solid curve}: fit given in Table~\ref{fits}.}
\label{HD+}
\end{figure}
\subsection{HeH$^+$}
HeH$^+$ is the first molecule formed in the initial stages of chemical synthesis in the primordial Universe chemistry
(GP98, Lepp et al.~2002, Dalgarno~2005, Hirata \& Padmanhaban 2006).
It represents one of the reagents for the process:\\
HeH$^+$+H$\rightarrow$ H$_2^+$+He\\
\noindent which, together with H+H$^+$ radiative association, is the main formation pathway for the molecular hydrogen cation.
The equilibrium, permanent dipole moment of HeH$^+$ is approximately equal to 1.66 D (Pavanello et al.~2005).
Engel et al.~(2005) computed line lists for different HeH$^+$ isotopologues ($^3$HeH$^+$,$^4$HeH$^+$,
$^3$HeD$^+$,$^4$HeD$^+$), and data are available electronically\footnote{exomol - Molecular Line lists for Exoplanet Atmospheres, www.exomol.com} both for ro-vibrational levels and Einstein coefficients.
The cooling functions for all the isotopologues are shown in Figs. \ref{HeH+_a} and \ref{HeH+_b}.
\begin{figure}
\includegraphics[width=0.5\textwidth]{f3.eps}
\caption{HeH$^+$ radiative cooling function contributions. {\it Circles}: $^3$HeH$^+$; {\it squares}: $^3$HeD$^+$;
{\it green solid curve}: $^3$HeH$^+$ fit (Table~\ref{fits}); {\it blue dashed curve}: $^3$HeD$^+$ fit (Table~\ref{fits}).}
\label{HeH+_a}
\end{figure}
\begin{figure}
\includegraphics[width=0.5\textwidth]{f4.eps}
\caption{HeH$^+$ radiative cooling function contributions. {\it Circles}: $^4$HeH$^+$; {\it squares}: $^4$HeD$^+$;
{\it green solid curve}: $^4$HeH$^+$ fit (Table~\ref{fits}); {\it blue dashed curve}: $^4$HeD$^+$ fit (Table~\ref{fits}).}
\label{HeH+_b}
\end{figure}
\subsection{LiH}
LiH represents the primordial molecule with the largest permanent equilibrium
dipole moment ($\approx 5.88$~D, Hertzberg~1978).
Several studies have modelled in detail
lithium kinetics in the primordial Universe chemistry (e.g. Bougleux \& Galli~1997, Lepp et al.~2002),
also explicitly evaluating the cooling properties of lithium hydride due to
collisional processes.
The cooling properties have been described by GP98, where a fit of the collisional cooling function
in the low density limit was given.
$^7$LiH has a singlet $X \, ^1 \Sigma^+$ ground state with an experimental equilibrium distance
$r_e = 3.015$~a$_0$ and a dissociation energy
$D_e = 20,287.7$~cm$^{-1}$ (Stwalley \& Zemke~1993).
In virtue of its small number of electrons LiH has been the object of
many quantum mechanical studies, among which we may mention
the pioneering one by \cite{Mulliken1936} and the
recent ones by \cite{bande2010}, \cite{Cooper2009} and \cite{Gadea2009}.
These studies, however,
do not provide in a readily usable format
the potential energy curve and dipole moment curve we need
to compute the cooling function. We therefore independently computed the required quantities.
\subsubsection{Computational detail}\label{section-LiH}
The potential energy and dipole moment curves have been calculated using the program
\cite{ref.molpro}, supplemented with the \cite{mrcc} package.
Descriptions of the quantum chemical methods used, along with references to the
original papers, can be found, e.g., in the review by \cite{Lodi2010}.
The basis sets used for hydrogen belong to aug-cc-pV$n$Z, ($n = $ T, Q, 5)
family by \cite{Dunning1989}.
For lithium we used the recent aug-cc-pwC$n$Z(-DK) ($n=$ T, Q, 5) basis sets by \cite{prasher}.
Basis sets with the DK suffix were especially devised to be used with the
relativistic {\sc dkh} hamiltonian (see below).
Computationally it would be entirely possible to use larger basis sets but, regrettably, no larger ones
are available for lithium.
The restricted Hartree-Fock ({\sc rhf}) method applied to LiH wrongly dissociates to the ionic limit
Li$^+$(1$s^2$)+H$^{-}$(1$s^2$)
instead of the correct neutral pathway Li(1$s^2$2$s^1$)+H(1$s^1$).
This behaviour is a direct consequence of the {\sc rhf} model, which upon dissociation
restricts molecular fragments to be spin-singlets (see e.g. Lodi \& Tennyson~2010).
Because the dissociation limit is wrong, methods such as {\sc ccsd(t)}
which rely on the {\sc rhf} solution being a good approximation
to the exact wave function may be expected to
experience problems at long bond lengths.
In fact, the {\sc ccsd(t)} module in {\sc molpro} does not converge
for $r$ greater than about 6~a$_0$ (using the default settings).
This may be slightly suprising because, as LiH has only 2 valence electron, method such as {\sc ccsd(t)}
are formally equivalent to full configuration interaction ({\sc fci}) in the valence space and should therefore be quite accurate.
It is the non-linearity of the equations to be solved, together with the poor starting guess,
that result in these numerical problems.
We calculated for six geometries all-electron full configuration interaction ({\sc fci})
reference energies in the aug-cc-pwCQZ basis sets
using the {\sc ccsdtq} code of {\sc mrcc}.
Computation of each {\sc fci} point involves about $36 \times 10^6$
configuration state functions and took between 30h and 50h on a 4-core Intel Xeon 5160
workstation with 16~GiB of memory.
We compared these {\sc fci} values with several single-reference and multi-reference methods;
the results of the comparisons, not discussed here, indicate that a very good balance of
speed and accuracy is provided by the internally-contracted, multi-reference averaged coupled pair
functional ({\sc ic-acpf}) method based on a full-valence (2-electron, 5-orbital) complete active space
(8 reference configurations).
This method was therefore used for further calculations.
With respect to the {\sc fci} energies the {\sc ic-acpf} method predicts a dissociation energy
too low by about 6.0~cm$^{-1}$.
Computation of a {\sc ic-acpf} energy in the aug-cc-pCw5Z-DK basis set took about 4~min on the same Xeon 5160 machine.
To account for scalar-relativistic effects we considered the use of the traditional mass-velocity
one-electron Darwin operator ({\sc mvd1}) or, alternatively, of the fourth-order
Dirac-Kroll-Hess hamiltonian ({\sc dkh4}).
Relative energies computed by the two methods agree to better
than $\pm 0.05$~cm$^{-1}$. We chose the {\sc dkh4} approach.
Relativistic corrections are small, affecting relative energies only by $\pm 2$~cm$^{-1}$.
The contribution of relativistic corrections to the dissociation energy is $-0.5$~cm$^{-1}$.
Finally, we remark that in our tests differences in relative energies between DK-contracted and
aug-cc-pwC$n$Z basis sets are negligible, less than $\pm 0.05$~cm$^{-1}$ for
energies up to dissociation.
The energy and dipole curves were computed for 300 uniformely-spaced grid points for
$r =1.00$~a$_0$ to $r = 16.00$~a$_0$ with spacing $0.05$ a$_0$.
The final energy values were obtained for each grid point
by basis-set extrapolation of the $n=$T, Q and 5 values;
following common practice (see, e.g., Lodi \& Tennyson~2010) the {\sc cas-scf} energies
were extrapolated using the functional form $E_n = E_\infty + A e^{-\alpha n}$ while
the differences between {\sc ic-acpf} and {\sc cas-scf} energies were extrapolated
using $E_n = E_\infty + A / n^3 $.
A minor complication arose because, for all basis sets employed,
the {\sc cas-scf} energies experience a discontinuity jump of
about 6~cm$^{-1}$ around $r=11$~a$_0$ if {\sc rhf}
orbitals are used as starting guess; as a result the {\sc ic-acpf} energies
also experience a discontinuity jump of about 1.2~cm$^{-1}$.
This problem is undoubtedly another consequence of the poor {\sc rhf} starting guess; it
was circumvented by starting calculations at small bond lengths and then
concatenating subsequent calculations for increasing $r$ so that
the {\sc cas-scf} orbitals from the preceeding geometry were used
as starting guess.
The final potential energy curve has an equilibrium bond length $r_e = 3.014$~a$_0$ and a dissociation energy $D_e=20,294$~cm$^{-1}$; the dipole moment curve has an equilibrium dipole moment of 2.293~a.u., reaching
a maximum of 3.003~a.u. for $r = 5.180$~a$_0$.
We may assign to $D_e$ an uncertainty of the order of $\pm 15$~cm$^{-1}$, mainly due to basis set incompleteness.
For comparison, the non-relativistic `exact BO' value quoted by \cite{Cooper2009} is $D_e = 20,298.8$~cm$^{-1}$.
The {\sc pec} and the {\sc dmc} are plotted in Fig.~\ref{LiH-plot}.
\begin{figure}
\includegraphics[width=0.5\textwidth]{f5.eps}
\caption{LiH potential energy curve ({\sc pec}) and dipole moment curve ({\sc dmc}). The two curves use different scales for the $y$ axis; the left-hand side legend refers to the {\sc pec}, the right-hand side one to the {\sc dmc}.}\label{LiH-plot}
\end{figure}
\begin{figure}
\includegraphics[width=0.5\textwidth]{f6.eps}
\caption{LiH radiative cooling function.{\it Circles}: present calculation; {\it green solid curve}: fit given in Table~\ref{fits}.}
\label{LiH}
\end{figure}
\subsection{LiH$^+$}
LiH$^+$ represents the most abundant molecular species containing lithium at low redshift. Indeed, while in
the very beginning phases LiH is much more abundant than its cation, this behaviour reverses at $z\sim20$,
as predicted by Dalgarno \& Lepp~(1987) and subsequently found by kinetics
models (e.g. GP98, Bougleaux \& Galli~1997). The reason
of this behaviour is the presence of a residual
ionization fraction which enhances one of the
main formation channel, namely ion-atom radiative association.
So far LiH$^+$ has not been characterised experimentally.
The present calculations, discussed below, predict a very modest
dissociation energy $D_e$ of 1130.5~cm$^{-1}$, an equilibrium bond length
$r_e = 4.130$~a$_0$ and a permanent equilibrium dipole (measured in the
molecular centre-of-mass reference frame) of 0.2773~a.u. (0.7048~D).
Very specialised, high-accuracy methods are applicable to LiH$^+$ due to its small size.
Very recently \cite{Bubin2011} presented non-adiabatic, relativistic-corrected
values for the first five rotationless ($J=0$) vibrational states with a stated accuracy of better than
$0.1$~cm$^{-1}$.
Among other recent studies treating LiH$^+$ we may mention \cite{Gadea2009} and \cite{Magnier2004}.
However, again as in the case of LiH, because of the different focus these recent
studies do not provide in a usable-format both the
potential energy curve and the dipole moment curve we need to compute the cooling function.
We therefore independently computed the necessary quantities.
\subsubsection{Computational details}
Like the LiH case discussed in Section~\ref{section-LiH}
we used the aug-cc-pwC$n$Z-DK ($n = $ T, Q, 5) basis sets
and we accounted for scalar relativistic effects by specifying
the {\sc dkh4} hamiltonian. Calculations were done using \cite{ref.molpro}.
The restricted open-shell Hartree-Fock method ({\sc rohf}) applied to LiH$^+$
correctly dissociates to Li$^+$ + H.
Single-reference methods are therefore expected to fare very well, in contrast
to the LiH case discussed in Section~\ref{section-LiH}.
As a preliminary test we compared using the aug-cc-pwCTZ-DK basis set
38 full configuration interaction
({\sc fci}) relative energies in the range 3.00--11.00~a$_0$ with
energies computed by various other methods.
Here is a brief summary of the comparison,
in the form {\sc method}/error: {\sc rohf}/98.9~cm$^{-1}$;
{\sc cisd}/1.8~cm$^{-1}$;
{\sc uccsd}/1.7~cm$^{-1}$; {\sc cisd+p}/1.4~cm$^{-1}$;
{\sc cisd+q}/0.6~cm$^{-1}$; {\sc uccsd(t)}/0.3~cm$^{-1}$.
The reported errors are one half of the
non-parallelity error, which is defined as the difference between the
maximum and the minimum of the modulus of the energy differences
between two methods (Li \& Paldus~1995).
As expected, {\sc uccsd(t)} produces extremely accurate results; also
note that this method is asymptotic to {\sc fci}
upon dissociation as it is size-extensive and exact for
two-electron systems.
We therefore used {\sc rhf-uccsd(t)} as implemented in \cite{ref.molpro} for further calculations.
Because LiH$^+$ is a charged species the value of the dipole moment depends on the choice of
the origin of the axes. As discussed by \cite{Bunker-Jensen}, the correct choice
of the origin is the molecular centre-of-mass.
{\sc molpro} by default does set the origin at the centre-of-mass but uses
isotopically-averaged nuclear masses, which is not appropriate in our case.
We therefore specified the following masses: $m({\mathrm{Li}^+)}=7.015455$~u and
$m(\mathrm{H})=1.0078250321$~u. These are \emph{atomic} masses, to partially account
for the electron contribution to the centre-of-mass coordinates;
$m({\mathrm{Li}^+)}$ was obtained by subtracting one electron mass
to the atomic mass of Li.
With this choice of masses and taking the molecule to lie along the $x$ axis
the coordinates of the nuclei for an inter-nuclear distance $r$ are:
$x(\mathrm{Li}^+) = 0.125613 r$ and $x(\mathrm{H}) = -0.874387 r$.
The dipole moment $M(r)$ is therefore asymtotic
to $M(r) \to 0.125613 r$ in this reference system.
We present as figure~\ref{plot-LiH+} a plot of the {\sc pec} and of the
quantity $0.125613 r - M(r)$, namely the difference between the
asymptotic dipole and the actual one. We chose not to plot directly $M(r)$
as, because of the asymptotic form, it would very nearly appear as a straight line.
We report here below the rotationless ($J=0$) vibrational terms $E(\nu+1)-E(\nu)$ together
with the difference with the very accurate values by \cite{Bubin2011} for $\nu=0,1,2,3,4$:
354.59(0.44); 261.79(0.78); 170.00(0.59); 89.82(-0.07); 35.26(0.18).
Our computed spectroscopic values are therefore very accurate,
with errors of less than 1~cm$^{-1}$.
\begin{figure}
\includegraphics[width=0.5\textwidth]{f7.eps}
\caption{LiH$^+$ potential energy curve ({\sc pec}) and the quantity $0.125613 r - M(r)$ (see text for details). The two curves use different scales for the $y$ axis; the left-hand side legend refers to the {\sc pec}, the right-hand side one to the {\sc dmc}.}\label{plot-LiH+}
\end{figure}
\begin{figure}
\includegraphics[width=0.5\textwidth]{f8.eps}
\caption{LiH$^+$ radiative cooling function: {\it circles}: present calculation; {\it green solid line}: fit given in Table~\ref{fits}.}
\label{LiH+}
\end{figure}
\subsection{H$_3^+$}
H$_3^+$ is one of the most interesting molecular ion in astrophysics, especially for its role in the interstellar medium chemistry. Its cooling properties
have been recently considered, both in planetary
atmosphere conditions by \cite{b18bis} and in the primordial Universe by \cite{b8}. The former work gives
a fit for the radiative cooling function as a function of temperature based
on the {\it ab initio} line list of Neale et al.~(1996).
\subsection{H$_2$D$^+$}
In the deuterium chemistry of primordial Universe, H$_2$D$^+$ represents the most complex triatomic molecule usually introduced
\citep{stancil,vol}. Recently, Sochi \& Tennyson (2010) have calculated a comprehensive line list of H$_2$D$^+$ frequencies
and transition probabilities. Table~\ref{fits} provides an improved fit for the radiative cooling function at low temperatures.
\begin{figure}
\includegraphics[width=0.5\textwidth]{f9.eps}
\caption{H$_2$D$^+$ radiative cooling function: {\it circles}: calculation by Sochi \& Tennyson~2010; {\it green solid line}: fit given in Table~\ref{fits}.}
\label{LiH+_sochi}
\end{figure}
\section{Conclusions}\label{conclusioni}
In the present work, the cooling functions of the most abundant molecular species in the primordial Universe have been
studied; calculations of the radiative contributions have been taken into account, under the hypothesis of
local thermal equilibrium distributions of ro-vibrational levels.
Radiative cooling functions are shown for each molecule. Analytic fits to
each of these functions were obtained in the form:
\begin{equation}
\log_{10}~W =\sum_{n=0}^N a_{n}(\log_{10} T)^{n}.
\end{equation}
Coefficients for these fits, $a_{n}$,
are provided in Table~\ref{fits}. The fits are valid only in the
temperature range specified in the figures.
Potential energy and dipole moment curves calculated in the present work,
together with the computed transition wavenumbers and Einstein
coefficients,
can be downloaded from www.exomol.com.
\begin{table*}
\centering
\begin{minipage}{140mm}
\caption{Radiative cooling function fits.}
\begin{tabular}{@{}lllllllll@{}}
\hline
Molecule & & Coefficients&Molecule & & Coefficients&Molecule & & Coefficients\\
\hline
HD &N=5 &$a_0=-55.5725$ &HD$^+$ &N=7&$a_0=-6.04917$& $^3$HeH$^+$ &N=7 &$a_0=-10.2807$ \\
&&$a_1=56.649$& &&$a_1=-60.0312$ &&&$a_1=-62.9415$ \\
&&$a_2=-37.9102$& &&$a_2=98.8361$ &&&$a_2=118.348$\\
&&$a_3=12.698$& &&$a_3=-77.5575$ &&&$a_3=-97.3721$\\
&&$a_4=-2.02424$& &&$a_4=33.4951$ &&&$a_4=42.5517$\\
&&$a_5=0.122393$& &&$a_5=-8.07092$ &&&$a_5=-10.1946$\\
&&& &&$a_6=1.01514$ &&&$a_6=1.26335$\\
&&& &&$a_7=-0.0519287$ &&&$a_7=-0.0633433$\\
&&&&&&&&\\
$^4$HeH$^+$ &N=7 &$a_0=-7.58736$&$^3$HeD$^+$ &N=7 &$a_0=26.2717$& $^4$HeD$^+$ &N=7 &$a_0=28.5384$ \\
&&$a_1=-68.2966$& &&$a_1=-168.493$ &&&$a_1=-172.458$ \\
&&$a_2=122.847$& &&$a_2=247.288$ &&&$a_2=249.811$\\
&&$a_3=-99.444$& &&$a_3=-184.299$ &&&$a_3=-184.786$\\
&&$a_4=43.1409$& &&$a_4=77.0755$ &&&$a_4=76.9479$\\
&&$a_5=-10.3034$& &&$a_5=-18.2012$ &&&$a_5=-18.1306$\\
&&$a_6=1.27565$& &&$a_6=2.26293$ &&&$a_6=2.25209$\\
&&$a_7=-0.0639846$& &&$a_7=-0.11512$ &&&$a_7=-0.114559$\\
&&&&&&&&\\
LiH &N=6 &$a_0=-31.894$ &LiH$^+$ &N=7 &$a_0=-23.5448$ &H$_2$D$^+$ &N=6 &$a_0=33.8462$ \\
&&$a_1=34.3512$& &&$a_1=9.93105$ &&&$a_1=-188.249$ \\
&&$a_2=-31.0805$& &&$a_2=-8.6467$ &&&$a_2=253.037$\\
&&$a_3=14.9459$& &&$a_3=2.13166$ &&&$a_3=-168.02$\\
&&$a_4=-3.72318$& &&$a_4=2.43072$ &&&$a_4=59.3597$\\
&&$a_5=0.455555$& &&$a_5=-1.69457$ &&&$a_5=-10.6334$\\
&&$a_6=-0.0216129$& &&$a_6=0.384871$ &&&$a_6=0.759029$\\
&&& &&$a_7=-0.0300114$ &&&\\
&&&&&&&&\\
H$_3^+$ &&fit by \cite{b18bis}&&&&&&\\
\hline
\label{fits}
\end{tabular}
\end{minipage}
\end{table*}
\section*{Acknowledgments}
We thank Eveline Roueff and Herve Abgrall for making available their calculations on dipolar and quadrupolar emission
probabilities of HD. C.M.C. would also acknowledge Savino Longo for helpful discussion on collisional cooling functions, and MIUR
and Universit\`{a} degli Studi di Bari, that partially supported this work (\textquotedblleft fondi di Ateneo 2010\textquotedblright).
|
\section{Introduction}
In this note we consider area-minimizing integral currents $T$ of dimension $m$ in $\R{m+n}$.
The following theorem is the cornerstone of the regularity theory.
It was proved for the first time by De Giorgi \cite{DG} for $n=1$ and then extended later by several authors (the constant $\omega_m$ denotes, as usual, the Lebesgue measure
of the $m$-dimensional unit ball).
\begin{theorem}\label{t:degiorgi}
There exist constants ${\varepsilon},\beta>0$ such that, if $T$ is an area-minimizing integral current
and $p$ is a point in its support such that $\theta (T, p) =1$, $\supp (\partial
T)\cap B_r (p)=\emptyset$ and $\|T\| (B_r (p))\leq (\omega_m +{\varepsilon})\,r^m$,
then $\supp(T)\cap B_{r/2} (p)$ is the graph of a $C^{1,\beta}$ function $f$.
\end{theorem}
Once established this ${\varepsilon}$-regularity result, the regularity theory proceeds further by deriving
the usual Euler--Lagrange equations for the function $f$.
Indeed, it turns out that $f$ solves a system of elliptic partial differential equations
and the Schauder theory then implies that $f$ is smooth
(in fact analytic, using the classical result by Hopf \cite{Hopf}).
In his Big regularity paper \cite{Alm}, Almgren observes that
an intermediate regularity result can be derived as
a consequence of a more complicated construction
without using the nonparametric PDE theory of minimal surfaces
(i.e.~without deriving the
Euler-Lagrange equation for the graph of $f$).
Indeed, given a minimizing current $T$ and a point $p$ with $\theta (T, p)= Q\in
\N$,
under the hypothesis that the excess is sufficiently small,
Almgren succeeds in constructing a $C^{3,\alpha}$ regular surface
(called {\em center manifold}) which, roughly speaking, approximates
the ``average of the sheets of the current'' (we refer to \cite{Alm} for further details).
In the introduction of \cite{Alm} it is observed that,
in the case $Q=1$, the center manifold coincides with
the current itself, thus implying directly the $C^{3,\alpha}$ regularity.
The aim of the present note is to give a simple direct proof of this remark,
essentially following Almgren's
strategy for the construction of the center manifold in the simplified setting $Q=1$.
At this point the following comment is in order: the excess decay leading to Theorem~\ref{t:degiorgi}
remains anyway a fundamental step in the proof of this paper (see Proposition~\ref{p:degiorgi_improved} below) and, as far as we understand,
of Almgren's approach as well.
One can take advantage of the information contained in
Theorem~\ref{t:degiorgi} at several levels but
we have decided to keep its use to the minimum.
\section{Preliminaries}
\subsection{Some notation}\label{s:notation}
From now on we assume, without loss of generality, that $T$ is an area-minimizing
integer rectifiable current in $\R{m+n}$ satisfying the following assumptions:
\begin{equation*}
\hspace{2.5cm} \partial T = 0\; \text{in}\; B_1(0),
\quad \theta (T, 0)=1
\quad\text{and}\quad \|T\| (B_1)\leq \omega_m +{\varepsilon},
\hspace{2.5cm}
\text{(H)}
\end{equation*}
with the small constant ${\varepsilon}$ to be specified later.
In what follows, $B^m_r (q)$, $B^n_r (u)$ and $B^{m+n}_r (p)$ denote the
open balls contained,
respectively, in the Euclidean spaces $\R{m}$, $\R{n}$ and $\R{m+n}$.
Given a $m$-dimensional plane $\pi$, ${\mathcal{C}}^{\pi}_r (q)$ denotes the cylinder
$B^m_r (q)\times \pi^\perp \subset \pi\times \pi^\perp = \R{m+n}$ and
$\proj^\pi:\pi\times\pi^\perp\to\pi$ the corresponding orthogonal projection.
Central points, supscripts and subscripts will be often omitted when
they are clear from the context.
We will consider different systems of cartesian coordinates in $\R{m+n}$.
A corollary of De Giorgi's excess decay theorem (a variant of which is precisely
stated in Proposition \ref{p:degiorgi_improved} below) is that, when ${\varepsilon}$ is sufficiently small,
the current has a unique tangent plane at the origin (see Corollary \ref{c:decay_everywhere}). Thus, immediately after
the statement of Corollary \ref{c:decay_everywhere}, the most important
system of coordinates, denoted by $x$, will be fixed once and for all in such a way that
$\pi_0=\{x_{m+1} = \ldots = x_{m+n} = 0\}$ is the tangent plane to $T$ at $0$.
Other systems of coordinates will be denoted by $x'$, $y$ or $y'$.
We will always consider positively oriented systems $x'$, i.e.~such that
there is a unique element $A\in SO (m+n)$ with $x'(p) = A \cdot x(p)$ for every point $p$.
An important role in each system of coordinates will be played by the oriented $m$-dimensional
plane $\pi$ where the last $n$ coordinates vanish (and by its orthogonal complement $\pi^\perp$).
Obviously, given $\pi$ there are several systems of coordinates $y$ for which $\pi = \{y_{m+1}= \ldots
= y_{m+n}=0\}$. However, when we want to stress the relation between $y$ and $\pi$ we will
use the notation $y_\pi$.
\subsection{Lipschitz approximation of minimal currents}
The following approximation theorem can be found in several accounts of the regularity theory
for area-minimizing currents.
It can also be seen as a special case of a much more general result due to Almgren
(see the third chapter of \cite{Alm}) and reproved in a simpler way in
\cite{DLSp2}. As it is customary the (rescaled)
cylindrical excess is given by the formula
\begin{equation}\label{e:cyl_ex}
\textup{Ex} (T, \mathcal{C}^\pi_r) := \frac{\|T\| (\mathcal{C}_r^\pi) - \omega_m
r^m}{\omega_m\,r^m}
= \frac{1}{2\,\omega_m\,r^m} \int_{\mathcal{C}_r^\pi} |\vec{T}- \vec{\pi}|^2\,
d\|T\|,
\end{equation}
(where $\vec{\pi}$ is the unit simple vector orienting $\pi$
and the last equality in \eqref{e:cyl_ex} holds when we assume
$\proj_\sharp (T\res \mathcal{C}_r^\pi) = \a{B_r (p}$).
\begin{propos}\label{p:approx}
There are constants $C>0$ and $0<\eta,{\varepsilon}_1<1$ with the following property.
Let $r>0$ and $T$ be an area-minimizing integer rectifiable $m$-current in ${\mathcal{C}}^\pi_{r}$ such that
\[
\partial \, T=0,\quad
\proj^\pi_\#(T)=\a{B^m_r}\quad\text{and}\quad
E := \textup{Ex} (T, {\mathcal{C}}^\pi_{r})\leq {\varepsilon}_1.
\]
Then, for $s=r(1-C E^\eta)$, there exists a Lipschitz function $f: B_{s} \to \R{n}$
and a closed set $K\subset B_{s}$ such that:
\begin{subequations}
\begin{gather}
{\rm {Lip}} (f) \leq C E^\eta;\label{e:approx1}\\
|B_{s}\setminus K|\leq C\,r^m\, E^{1+\eta} \quad\text{and}\quad
\textup{graph} (f|_K) = T \res (K\times \R{n});\label{e:approx2}\\
\left|\|T\|({\mathcal{C}}_{s}) -\omega_m\,s^m - \int_{B_{s}} \frac{|Df|^2}{2}\right|
\leq C \,r^m\,E^{1+\eta}.\label{e:approx3}
\end{gather}
\end{subequations}
\end{propos}
This proposition is a key step in the derivation of Theorem~\ref{t:degiorgi}.
In the appendix we include a short proof in the spirit of \cite{DLSp2}.
Clearly, Theorem~\ref{t:degiorgi} can be thought as a much finer version of this approximation.
However, an aspect which is crucial for further developments is that
several important estimates can be derived directly from Proposition~\ref{p:approx}.
\subsection{De Giorgi's excess decay}
The fundamental step in De Giorgi's proof of Theorem~\ref{t:degiorgi} is the
decay of the quantity usually called ``spherical excess''
(where the minimum is taken over all oriented $m$--planes $\pi$):
\begin{equation*
\textup{Ex} (T, B_r (p)) := \min_\pi \textup{Ex} (T, B_r (p), \pi),
\quad\text{with}\quad
\textup{Ex} (T, B_r (p), \pi) := \frac{1}{2} \hspace{0.1cm}-\hspace{-0.45cm}\int_{B_r (p)} |\vec{T}-\vec{\pi}|^2 d\|T\|.
\end{equation*}
\begin{propos}\label{p:degiorgi_improved}
There is a dimensional constant $C$ with the following
property. For every $\delta,{\varepsilon}_0>0$, there is ${\varepsilon}>0$ such that, if (H) holds, then
$\textup{Ex} (T, B_1)\leq {\varepsilon}^2_0$ and
\begin{equation}\label{e:decay_everywhere}
\textup{Ex} (T, B_r (p)) \leq C\, {\varepsilon}^2_0\, r^{2-2\delta}
\quad \mbox{for every $r\leq 1/2$ and every $p\in B_{1/2}\cap \supp (T)$.}
\end{equation}
\end{propos}
From now on we will consider the constant $\delta$ fixed. Its
choice will be specified much later.
\begin{definition}\label{d:admissible}
For later reference, we say that a plane $\pi$ is \textit{admissible}
in $p$ at scale $\rho$ (or simply that $(p,\rho,\pi)$ is \textit{admissible})
if
\begin{equation}\label{e:good}
\textup{Ex} (T, B_{\rho} (p), \pi) \leq C_{m,n}{\varepsilon}^2_0\, \rho^{2-2\delta},
\end{equation}
for some fixed (possibly large) dimensional constant $C_{m,n}$.
\end{definition}
Proposition~\ref{p:degiorgi_improved} guarantees that, for every
$p$ and $r$ as in the statement, there exists always an admissible plane $\pi_{p,r}$.
The following is a straightforward consequence of Proposition~\ref{p:degiorgi_improved}
which will be extensively used.
\begin{corol}\label{c:decay_everywhere}
There are dimensional constants $C$, $C'$ and $C''$ with the following
property. For every $\delta, {\varepsilon}_0>0$, there is ${\varepsilon}>0$ such that,
under the assumption (H):
\begin{itemize}
\item[(a)] if $(p,\rho,\pi)$ and $(p',\rho',\pi')$ are admissible
(according to Definition \ref{d:admissible}), then
\begin{equation*}
|\vec{\pi} - \vec{\pi}'| \leq C\,{\varepsilon}_0 \big(\max \{\rho, \rho', |q-q'|\}\big)^{1-\delta};
\end{equation*}
\item[(b)] there exists a unique tangent plane $\pi_p$ to $T$ at every
$p \in \supp (T)\cap B_{1/2}$; moreover, if $(p,\rho,\pi)$ is admissible then
$|\pi-\pi_p|\leq C'\,{\varepsilon}_0\,\rho^{1-\delta}$ and, vice versa,
if $|\pi-\pi_p|\leq C''\,{\varepsilon}_0\,\rho^{1-\delta}$, then
$(p, \rho, \pi)$ is admissible;
\item[(c)] for every $q\in B_{1/4}^m$, there exists a unique $u\in \R{n}$ such that
$(q,u)\in \supp(T)\cap B_{1/2}$.
\end{itemize}
\end{corol}
\begin{remark}
An important point in the previous corollary is that the constant $C''$ can
be chosen arbitrarily large, provided the constant $C_{m,n}$ in Definition \ref{d:admissible}
is chosen accordingly. This fact is an easy consequence of the proof given in the appendix.
\end{remark}
Theorem~\ref{t:degiorgi} is clearly contained in the previous corollary
(with the additional feature that the H\"older exponent $\beta$ is equal to $1-\delta$,
i.e.~is arbitrarily close to $1$).
In order to make the paper self-contained, we will include also a proof of
Proposition~\ref{p:degiorgi_improved} and Corollary~\ref{c:decay_everywhere} in the appendix.
\subsection{Two technical lemmas}
We conclude this section with the following two lemmas which will be needed in the sequel.
Consider two functions $f:D\subset\pi_0\to \pi_0^\perp$ and $f': D'\subset\pi\to \pi^\perp$,
with associated systems of coordinates $x$ and $x'$, respectively, and
$x'(p)=A\cdot x(p)$ for every $p\in\R{m+n}$.
If for every $q'\in D'$ there exists a unique $q\in D$ such that $(q', f'(q')) = A \cdot (q, f(q))$ and vice versa, then
it follows that $\textup{graph}_{\pi_0}(f)=\textup{graph}_\pi(f')$, where
\[
\textup{graph}_{\pi_0} (f) := \big\{(q, f(q))\in D\times \pi_0^\perp\big\}
\quad\text{and}\quad
\textup{graph}_\pi (f') := \big\{(q', f'(q'))\in D'\times \pi^\perp\big\}.
\]
The following lemma compares norms of functions (and of differences of functions)
having the same graphs in two nearby system of coordinates.
\begin{lemma}\label{l:rotation}
There are constants $c_0,C>0$ with the following properties.
Assume that
\begin{itemize}
\item[(i)] $\|A-{\rm Id}\|\leq c_0$, $r\leq 1$;
\item[(ii)] $(q,u)\in\pi_0\times\pi_0^\perp$ is given and $f,g: B^m_{2r} (q)\to \R{n}$ are Lipschitz functions such that
\begin{equation*}
{\rm {Lip}} (f), {\rm {Lip}} (g) \leq c_0\quad\text{and}\quad |f(q)-u|+|g(q)-u|\leq c_0\, r.
\end{equation*}
\end{itemize}
Then, in the system of coordinates $x'= A\cdot x$, for $(q',u') = A \cdot(q,u)$, the following holds:
\begin{itemize}
\item[(a)] $\textup{graph}_{\pi_0} (f)$ and $\textup{graph}_{\pi_0} (g)$ are the graphs of two Lipschitz functions $f'$ and $g'$, whose domains of definition contain both $B_{r} (q')$;
\item[(b)] $\|f'-g'\|_{L^1 (B_{r} (q'))}\leq C\,\|f-g\|_{L^1 (B_{2r} (q))}$;
\item[(c)] if $f\in C^4 (B_{2r} (q))$, then $f'\in C^4 (B_{r} (q'))$, with the
estimates
\begin{eqnarray}
\|f'- u'\|_{C^3}&\leq& \Phi \left(\|A-\Id\|, \|f-u\|_{C^3}\label{e:est_C3}\right)\, ,\\
\|D^4 f'\|_{C^0} &\leq& \Psi \left(\|A-\Id\|, \|f-u\|_{C^3}\right)
\left(1+ \|D^4 f\|_{C^0}\right)\, ,
\label{e:est_C4}
\end{eqnarray}
where $\Phi$ and $\Psi$ are smooth functions.
\end{itemize}
\end{lemma}
\begin{proof} Let $P: \R{m\times n}\to \R{m}$ and
$Q: \R{m\times n}\to \R{n}$ be the usual orthogonal projections.
Set $\pi=A(\pi_0)$ and
consider the maps $F, G: B_{2r} (q)\to \pi^\perp$ and $I, J: B_{2r} (q)\to \pi$
given by
\[
F (x) = Q (A((x,f(x)))\quad\text{and}\quad G(x) = Q (A((x, g(x))),
\]
\[
I(x)= P (A((x, f(x)))\quad\text{and}\quad J (x) = P (A((x, g(x))).
\]
Obviously, if $c_0$ is sufficiently small, $I$ and $J$ are injective Lipschitz maps.
Hence, $\textup{graph}_{\pi_0} (f)$ and $\textup{graph}_{\pi_0} (g)$ coincide, in the new coordinates, with the graphs of the functions
$f'$ and $g'$ defined respectively in $D:= I (B_{2r} (q))$ and $\tilde{D}:= J (B_{2r} (q))$
by $f' = F \circ I^{-1}$ and $g'= G \circ J^{-1}$.
If $c_0$ is chosen sufficiently small, then we can find a constant $C$
such that
\begin{equation}\label{e:Lip_bound}
{\rm {Lip}} (I), \; {\rm {Lip}} (J),\; {\rm {Lip}} (I^{-1}),\; {\rm {Lip}} (J^{-1}) \leq 1+C\,c_0,
\end{equation}
and
\begin{equation}\label{e:Linfty_bound}
|I (q)-q'|, |J (q)-q'|\leq C\,c_0\, r.
\end{equation}
Clearly, \eqref{e:Lip_bound} and \eqref{e:Linfty_bound} easily imply (a).
Conclusion (c) is a simple consequence of the inverse function theorem.
Finally we claim that, for small $c_0$,
\begin{equation}\label{e:claim}
|f'(x')-g'(x')|\leq 2 \,|f (I^{-1} (x')) - g (I^{-1} (x'))|
\quad\forall \;x'\in B_r(q'),
\end{equation}
from which,
using the change of variables formula for biLipschitz homeomorphisms
and \eqref{e:Lip_bound}, (b) follows.
In order to prove \eqref{e:claim}, consider
any $x'\in B_r (q')$, set $x:= I^{-1} (x')$ and
\[
p_1 := (x, f(x))\in \pi_0\times \pi_0^\perp,\quad
p_2 := (x, g(x))\in \pi_0\times \pi_0^\perp\quad
\text{and}\quad
p_3 := (x', g'(x'))\in \pi\times \pi^\perp.
\]
Obviously $|f'(x')- g'(x')|= |p_1-p_3|$ and $|f(x)- g(x)|=|p_1-p_2|$.
Note that, $g(x)= f (x)$ if and only if $g'(x')= f' (x')$, and in this case \eqref{e:claim} follows trivially.
If this is not the case, the triangle with vertices $p_1$, $p_2$ and $p_3$ is non-degenerate.
Let $\theta_i$ be the angle at $p_i$.
Note that, ${\rm {Lip}} (g)\leq c_0$ implies $|90^\circ-\theta_2|\leq C c_0$ and $\|A-{\rm Id}\|\leq c_0$ implies
$|\theta_1|\leq C c_0$, for some dimensional constant $C$.
Since $\theta_3 = 180^\circ - \theta_1 - \theta_2$, we conclude
as well $|90^\circ - \theta_3|\leq C c_0$.
Therefore, if $c_0$ is small enough, we have $1 \leq 2\sin \theta_3$, so that,
by the Sinus Theorem,
\begin{equation*}
|f'(x')-g'(x')|= |p_1-p_3| = \frac{\sin \theta_2}{\sin \theta_3}\, |p_1-p_2|
\leq 2 \, |p_1-p_2| = 2 \,|f(x)-g(x)|,
\end{equation*}
thus concluding the claim.
\end{proof}
The following is an elementary lemma on polynomials.
\begin{lemma}\label{l:poly}
For every $n,m\in \N$, there exists a constant $C(m,n)$ such that,
for every polynomial $R$ of degree at most $n$ in $\R{m}$ and every positive $r>0$,
\begin{equation}\label{e:poly}
|D^k R (q)| \leq \frac{C}{r^{m+k}} \int_{B_r (q)} |R|\, \quad \text{for all }\; k\leq n\; \text{ and all }\;q\in\R{m}.
\end{equation}
\end{lemma}
\begin{proof} We rescale and translate the variables by setting
$S (x) = R (rx+q)$. The lemma is then reduced to show that
\begin{equation}\label{e:norms}
\sum_{k=0}^n |D^k S (0)|\leq C \int_{B_1 (0)} |S|,
\end{equation}
for every polynomial $S$ of degree at most $n$ in $\R{m}$, with $C=C(n,m)$.
Consider now the vector space $V^{n,m}$ of polynomials of degree
at most $n$ in $m$ variables. $V^{n,m}$ is obviously finite dimensional.
Moreover, on this space, the two quantities
\[
\|S\|_1:= \sum_{k=0}^n |D^k S (0)| \quad \text{and}\quad
\|S\|_2:= \int_{B_1 (0)} |S|
\]
are two norms.
The inequality \eqref{e:norms} is then a corollary
of the equivalence of norms on finite-dimensional vector spaces.
\end{proof}
\section{The approximation scheme and the main theorem}
The $C^{3,\alpha}$ regularity of the current $T$ will be deduced from
the limit of a suitable approximation scheme. In this section we describe
the scheme and state the main theorem of the paper.
We start by fixing a nonnegative kernel $\varphi\in C^\infty_c (B^m_1)$
which is radial and satisfies $\int\varphi =1$.
As usual, for $\tau>0$, we set $\varphi_\tau (w) := \tau^{-m} \varphi (w/\tau)$.
Consider the area-minimizing current
$S=T\res(B_{1/2}^{m+n}\cap{\mathcal{C}}^{\pi_0}_{1/4})$
(recall that $\pi_0 = \{x_{m+1} = \ldots = x_{m+n} = 0\}$ is the tangent
plane to $T$ at $0$).
From Corollary~\ref{c:decay_everywhere} (b) and (c),
it is simple to deduce the following: if $p=(q,u)\in\pi\times\pi^\perp$,
$\rho\leq 2^{-6}$ and $\pi$ form an admissible triple $(p,8\,\rho,\pi)$
with $p\in \supp(T)\cap B_{1/16}$, then
\[
\proj^\pi_\#(S\res {\mathcal{C}}^\pi_{8\rho}(q))=\a{B_{8\rho}(q)}
\quad\text{and}\quad
\partial S=0\quad\text{in}\quad {\mathcal{C}}^\pi_{8\rho}(q).
\]
From now on, we will assume that $C_{m,n}\,{\varepsilon}^2_0\,2^{-3(2-2\delta)}\leq
{\varepsilon}_1$,
where ${\varepsilon}_1$ is the constant of Proposition~\ref{p:approx} and $C_{m,n}$ the
constant of Proposition~\ref{p:degiorgi_improved}.
This assumption guarantees the existence of the Lipschitz approximation of
Proposition~\ref{p:approx},
which we restrict to $B^m_{6\rho} (q)$,
$f: B^m_{6\rho} (q)\subset\pi \to \pi^\perp$.
Then, consider the following functions:
\begin{itemize}
\item[($I_1$)] $\hat{f} = f * \varphi_{\rho}$;
\item[($I_2$)] $\bar{f}$ such that
\begin{equation*}
\left\{
\begin{array}{l}
\Delta \bar{f} = 0 \quad \text{on }\,B^m_{4\rho} (q),\\
\bar{f}|_{\partial B^m_{4\rho} (q)} = \hat{f};
\end{array}\right.
\end{equation*}
\item[($I_3$)] $g: B^m_\rho (q')\subset\pi_0\to \pi_0^\perp$,
with $x(p)=(q',u')\in\pi_0\times\pi_0^\perp$, such that
$\textup{graph}_{\pi_0}(g)=\textup{graph}_{\pi}(\bar f)$ in the cylinder ${\mathcal{C}}_\rho (q')\subset\pi_0\times \pi_0^\perp$.
\end{itemize}
\begin{remark}
In order to proceed further, we need to show the existence of $g$ as in ($I_3$).
We wish, therefore, to apply Lemma \ref{l:rotation} to the function $\bar{f}$.
First recall that $|\pi-\pi_0|\leq C{\varepsilon}_0|p|^{1-\delta}\leq C{\varepsilon}_0$ by
Corollary \ref{c:decay_everywhere}. Thus, assumption (i) in Lemma \ref{l:rotation}
is satisfied provided ${\varepsilon}_0$ is chosen sufficiently small.
Next note that, by the interior estimates for the harmonic functions and \eqref{e:approx1}, one has
\[
{\rm {Lip}}(\bar f|_{B_{3\rho}})\leq C{\rm {Lip}}(\hat f|_{B_{4\rho}})\leq C\,E^\eta\, .
\]
Moreover, if we consider the ball $B_s (p)$ with $s= \rho E^{\eta/(2m)}$, by the monotonocity
formula, $\|T\| (B_s (p))\geq \omega_m \rho^m E^{\eta/2}$. Thus, by \eqref{e:approx2},
the graph of $f$ contains a point in $B_s (p)$. Using the Lipschitz bound \eqref{e:approx1},
we then achieve $\|f-u\|_{C^0 (B_{6\rho} (q))}\leq C \rho E^{\eta/2}$, which in turn implies
$\|\bar{f} - u\|_{C^0 (B_{4\rho} (q))}$. Recalling that $E\leq C {\varepsilon}^2_0 \rho^{2-2\delta}$,
we conclude that condition (ii) in Lemma \ref{l:rotation} is satisfied when
${\varepsilon}_0$ is sufficiently small.
Therefore Lemma \ref{l:rotation}(a) guarantees that the function $g$ exists.
\end{remark}
\begin{remark}
It is obvious that in order to define the function $g$ we could have used,
in place of the $f$ given by Proposition \ref{p:approx},
the function whose graph gives the current $T$ in $B_{6\rho} (p)$.
This would have simplified many of the computations below. However, as mentioned
in the introduction, we hope that our choice helps in the understanding of
the more general construction of Almgren.
\end{remark}
The function $g$ is the main building block of the
construction of this paper. It is called the {\em $(p,\rho,\pi)$-interpolation of $T$} or,
if $\textup{Ex} (T, B_{8\rho} (p)) = \textup{Ex} (T, B_{8\rho} (p),\pi)$,
simply the {\em $(p, \rho)$-interpolation of $T$}.
The main estimates of the paper are contained in the following proposition.
\begin{propos}\label{p:main}
There are constants $\alpha,C>0$ such that, if $g, g'$ are respectively $(p,\rho,\pi)$- and
$(p',\rho,\pi')$-interpolations, then
\begin{subequations}\label{e:ABC}
\begin{gather}
\rho^{1-\alpha} \|D^4 g\|_{C^0}+ \|g\|_{C^3}\leq C,\label{e:A}\\
\sum_{\ell=0}^4
\rho^{\ell-3-\alpha}\|D^\ell g (x) - D^\ell g' (x)\|_{C^0}\leq C \quad\text{in }\;B_\rho(p)\cap B_{\rho}(p'),\label{e:B}\\
|D^3 g (q) - D^3 g' (q')|\leq C |q-q'|^\alpha,
\quad\text{with }\;p=(q,u),\,p'=(q',u').\label{e:C}
\end{gather}
\end{subequations}
\end{propos}
\subsection{Approximation scheme}
Let $5<n_0<k_0$ be natural numbers and consider the cube
$Q=[-2^{-n_0}, 2^{-n_0}]^m$.
For $k\geq k_0$, we consider the usual subdivision
of $\R{m}$ into dyadic cubes of size $2\cdot 2^{-k}$, centered
at points $c_i = 2^{-k} i \in 2^{-k} \Z^m$.
The corresponding closed cubes of the subdivision are then denoted by $Q_i$ and
we consider below only those $Q_i$'s which have nonempty intersection
with $Q$.
According to Corollary~\ref{c:decay_everywhere} and to the previous observations,
for every $c_i$ there exists a unique $u_i$ such that
$p_i = (c_i, u_i)\in \supp (T)\cap B_{1/16}$. Moreover,
for every constant $C$, if $k_0$ is large enough,
we can consider the $(p_i, C\, 2^{-k})$-interpolation $g_i$
for all $k\geq k_0$.
Let $\psi\in C^\infty_c ([-\textstyle{\frac{5}{4}},
\textstyle{\frac{5}{4}}]^m)$ be a nonnegative function such that,
if we define $\psi_i (q):= \psi (2^{k} (q-c_i))$, then
\[
\sum_{i\in \Z^{m}} \psi_i \equiv1\;\text{in}\;Q\, .
\]
Denote by $\mathcal{A}_i$ the set of indices $j$
such that $Q_j$ and $Q_i$ are adjacent.
Note that the choice of $\psi$ guarantees $\psi_i\,\psi_j=0$ if $j\not\in \mathcal{A}_i$.
Moreover the cardinality of $\mathcal{A}_i$ is (bounded by) a
dimensional constant independent of $k$ and, if $q\in Q_i$, then
in a neighborhood of $q$ we have
\begin{equation}\label{e:sum_der=0}
\sum_{j\in \mathcal{A}_i} \psi_j=1
\quad\text{and}\quad
\sum_{j\in \mathcal{A}_i} D^\ell \psi_j (q) = 0 \quad
\text{for all }\;\ell>0.
\end{equation}
We are now ready to state and prove the central theorem of this note.
\begin{theorem}\label{t:main}
There are dimensional constants $n_0<k_0$ with the following
properties.
Given an area-minimizing current $T$ as in (H) and $k\geq k_0$,
consider the functions $h_k : {Q} \to \R{n}$ given by
$h_k := \sum_i \psi_i\, g_i$. Then,
\begin{equation}\label{e:main_estimate}
\|h_k\|_{C^{3,\alpha}}\leq C,
\end{equation}
for some dimensional constants $\alpha>0$ and $C$ (which, in particular,
do not depend on $k$).
Moreover, the graphs of $h_k$ converge, in the sense of currents, to
$T\res (Q\times \R{n})\cap B_{1/2}$, thus implying that $T$ is a $C^{3,\alpha}$ graph in
a neighborhood of the origin.
\end{theorem}
\begin{proof}[Proof of Theorem~\ref{t:main}]
Given $k$, consider a cube $Q_i$ of the corresponding dyadic decomposition and a point $q\in Q_i$.
We already observed that, in a neighborhood of $q$,
$h_k = \sum_{j\in \mathcal{A}_i} \psi_j g_j$.
Moreover, from the definition, we have that
\begin{equation}\label{e:psi}
\|D^\ell \psi_j\|_{C^0}= 2^{k\ell}\, \|D^\ell\psi\|_{C^0}= C_\ell\, 2^{k\ell} \quad\text{for every }\;\ell\in\N.
\end{equation}
The $C^0$ estimate of $h_k$ follows trivially from \eqref{e:A}, since
\[
|h_k (q)|\leq \sum_{j\in\mathcal{A}_i} \|\psi_j\|_{C^0}\|g_j\|_{C^0}
\leq C.
\]
As for the $C^1$ estimate, we write the first derivative of $h_k$ as follows,
\[
D h_k (q) = \sum_{j\in \mathcal{A}_i} \big( D \psi_j (q) g_j (q) +
\psi_j (q) D g_j (q)\big) \stackrel{\eqref{e:sum_der=0}}{=} \sum_{j\in \mathcal{A}}
\big( D \psi_j (q) (g_j (q)- g_i (q)) + \psi_j (q) D g_j (q)\big),
\]
from which, using \eqref{e:A}, \eqref{e:B} and \eqref{e:psi}, we deduce
\[
|D h_k (q)|\leq \sum_{j\in \mathcal{A}_i} \big(\|D\psi_j\|_{C^0} \|g_i -g_j\|_{C^0}
+ \|\psi_j\|_{C^0} \|Dg_j\|_{C^0}\big)\leq C.
\]
With analogous computations, we obtain
\begin{align*}
|D^2 h_k (q)|\leq{}& \sum_{j\in\mathcal{A}_i} \big(\|D^2 \psi_j\|_{C^0}
\|g_i-g_j\|_{C^0} + \|D \psi_j\|_{C^0} \|Dg_j - Dg_i\|_{C^0}+\|\psi_j\|_{C^0} \|D^2 g_j\|_{C^0}\big)\leq C,\\
|D^3 h_k (q)|\leq{}& \sum_{j\in\mathcal{A}_i} \big(\|D^3 \psi_j\|_{C^0}
\|g_j-g_i\|_{C^0} + \|D^2 \psi_j\|_{C^0} \|Dg_j - Dg_i\|_{C^0}+\\
&+ \|D \psi_j\| \|D^2 g_j - D^2 g_i\|_{C^0} + \|\psi_j\|_{C^0} \|D^3 g_j\|_{C^0}\big)\leq C,\\
|D^4 h_k (q)|\leq{}& \sum_{j\in\mathcal{A}_i} \big(\|D^4 \psi_j\|_{C^0}
\|g_j-g_i\|_{C^0} + \|D^3 \psi_j\|_{C^0} \|Dg_j - Dg_i\|_{C^0}+\\
&+ \|D^2 \psi_j\| \|D^2 g_j - D^2 g_i\|_{C^0} + \|D\psi_j\|_{C^0} \|D^3 g_j-D^3g_i\|_{C^0}+\|\psi_j\|_{C^0} \|D^4 g_j\|_{C^0}\big)\\
\leq {}& C\,2^{k(1-\alpha)},
\end{align*}
where $C$ is a constant independent of $k$.
Now, let $q,q'\in B_{1/2}$ and consider the cubes $Q_i$ and $Q_j$ such that
$q\in Q_i$ and $q'\in Q_j$.
If the two cubes are adjacent, then we have $|q-q'|\leq C 2^{-k}$ and, therefore,
\[
|D^3 h_k (q) - D^3 h_k (q')|\leq \|D^4 h_k\|_{C^0} |q-q'|\leq C \,2^{k(1-\alpha)}\, |q-q'| \leq C\, |q-q'|^\alpha.
\]
If $Q_i$ and $Q_j$ are not adjacent, then $2\,|q-q'|\geq \max\{|c_i-c_j|, 2^{-k}\}$.
Since $\supp(\psi)\subset [-\textstyle{\frac{5}{4}},
\textstyle{\frac{5}{4}}]^m$, $D^3 h_k (c_i)= D^3 g_i (c_i)$ for every $i$ and
from \eqref{e:C} it follows that
\begin{align*}
|D^3 h_k (q)-D^3 h_k (q')| \leq{}&
|D^3 h_k (q)-D^3 h_k (c_i)| + |D^3 g_i (c_i)-D^3 g_j (c_j)|+\\
&+ |D^3 h_k (c_j) - D^3 h_k (q')|\\
\leq{}& C \,2^{-k} \|D^4 h_k\|_{C^0}
+ C |c_j-c_i|^\alpha
\leq C 2^{-k\alpha} + C |c_i-c_j|^\alpha\leq C |q-q'|^\alpha.
\end{align*}
This concludes the proof of \eqref{e:main_estimate}.
We finally come to the convergence of the graphs of $h_k$ in the sense of currents.
Obviously, by compactness we can assume that a subsequence of $h_k$
(not relabelled) converges in
the $C^3 (Q)$ norm to some limiting $C^{3,\alpha}$ function $h$.
On the other hand, by Corollary \ref{c:decay_everywhere}
and Proposition \ref{p:approx}, it follows easily that the support of
$T\res (Q\times\R{n})\cap B_{1/2}$ is contained in the graph of $h$.
But then, by the Constancy Theorem, $T\res (Q\times\R{n})\cap B_{1/2}$
must coincide with an integer multiple of the graph of $h$. Our assumptions
imply easily that the multiplicity is necessarily $1$.
\end{proof}
\section{$L^1$-estimate}
The rest of the paper is devoted to the proof of Proposition~\ref{p:main}.
A fundamental point is
an estimate for the $L^1$ distance between the harmonic function $\bar f$
introduced in step ($I_2$) of the approximation scheme
and the function $f$ itself.
A preliminary step is the following estimate on the Laplacian of $\hat f$, which
is a simple consequence of the first variation formula for area-minimizing currents.
\begin{lemma}\label{l:laplace}
There exists $\delta, \gamma, C, \lambda>0$ such that, if
$(p,8\rho, \pi)$ is admissible and $\hat{f}$ is as in ($I_1$), then
\begin{equation}\label{e:laplace}
\|\Delta \hat{f}\|_{C^0 (B^m_{5\rho})}\;\leq\; C \rho^{1+\lambda},
\end{equation}
\begin{equation}\label{e:laplace0}
\int_{B_{5\rho}}\bigg\vert \int_{B_\rho(w)} Df(z)\cdot D\gamma
(w-z) \,dz\bigg\vert\,dw\leq
C\, E^{1+\eta}\,\rho^m\,\|D\gamma\|_{L^1},
\quad\forall\gamma\in C_c^1(B_{\rho},\R{n}),
\end{equation}
where $\eta$ is the constant in Proposition~\ref{p:approx}
and $E =\textup{Ex} (T, B_{8\rho} (p), \pi)$.
\end{lemma}
\begin{proof}
Let $\mu$ be the measure defined by $\mu(A):=\|T\|(A\times\pi^{\perp})$.
We start showing that the approximation $f$ given by
Proposition~\ref{p:approx} satisfies
\begin{equation}\label{e:variation}
\left\vert\int Df\cdot D\kappa\right\vert
\leq C\int |D\kappa|\,|Df|^3\,dx+C\int |D\kappa|\,{\bf 1}_{B_{6\rho}\setminus K}\, (dx+ d\mu(x)),
\end{equation}
for every $\kappa\in C_c^1(B_{6\rho},\R{n})$.
Consider the vector field $\chi(x,y)=(0,\kappa (x))$.
From the minimality of the current $T$, we infer that the first variation of the mass in direction $\chi$
vanishes, $\delta T(\chi)=0$.
We set $T_f=\textup{graph}(f)$. Since $\delta T (\chi)=0$, we get
\begin{equation}\label{e:variation2}
\left\vert\int Df\cdot D\varphi\right\vert\leq
\left\vert\int Df\cdot D\varphi-\delta T_f(\chi)\right\vert+
\left\vert \delta T(\chi)-\delta T_f(\chi)\right\vert.
\end{equation}
The first variation $\delta T_f (\chi)$ is given by the formula
\begin{align*}
\int_{{\mathcal{C}}_{6\rho}}{\text {div}}_{\vec T_f}\chi \, d\|T_f\|&=\frac{d}{ds}\Big\vert_{s=0}
\int_{B_{6\rho}}\sqrt{1+|D f+sD \kappa|^2+\textstyle{\sum_{|\alpha|\geq2}}M_\alpha(D f+sD \kappa)^2}\,dx\\
&=\int_{B_{6\rho}}\frac{D f\cdot D\kappa+\textstyle{\sum_{|\alpha|\geq2}}M_\alpha(D f)\left.\frac{d}{ds}\right|_{s=0} M_{\alpha}(D f+sD \kappa)}{\sqrt{1+|D f|^2+\sum_{|\alpha|\geq2}M_\alpha(D f)^2}}.
\end{align*}
It follows then that
\begin{align*}
\left\vert\int_{{\mathcal{C}}_{6\rho}}\hspace{-0.2cm}{\text {div}}_{\vec T_f}\chi \, d\|T_f\|-
\int_{B_{6\rho}}\hspace{-0.2cm}Df\cdot D\kappa\right\vert\leq{}&
\int_{B_{6\rho}}\hspace{-0.2cm}|D f| |D\kappa|\left(\sqrt{1+|D f|^2+\textstyle{\sum_{|\alpha|\geq 2}} M_\alpha(D f)^2}-1\right)+\\
& +\left\vert\int_{B_{6\rho}}\textstyle{\sum_{|\alpha|\geq2}}M_\alpha(D f)\left.\frac{d}{ds}\right|_{s=0} M_{\alpha}(D f+sD \kappa)\right|\\
\leq {}& C\int_{B_{6\rho}}|D\kappa|\,|Df|^3.
\end{align*}
We next estimate the second term in the right hand side of
\eqref{e:variation2}:
\begin{align*}
\big\vert \delta T(\chi)-\delta T_f(\chi)\big\vert&\leq
\int_{B_{6\rho}\setminus K\times\R{n}}{\text {div}}_{\vec T}\chi \, d\|T\|
+\int_{B_{6\rho}\setminus K\times\R{n}}{\text {div}}_{\vec T_f}\chi \, d\|T_f\|\\
&\leq \int_{}{\bf 1}_{B_{6\rho}\setminus K}(x)|D\kappa|(x)\, d\mu(x)
+ C\int_{}{\bf 1}_{B_{6\rho}\setminus K}(x)|D\kappa|(x) \, dx,
\end{align*}
where we have used
the Lipschitz bound on $f$ to estimate the second integral in the right hand
side of the first line.
This concludes the proof of \eqref{e:variation}.
We now come to the proof of \eqref{e:laplace}.
From \eqref{e:variation} and Proposition~\ref{p:approx}, it follows straightforwardly that
\begin{equation}\label{e:step1}
\left\vert\int_{B_{6\rho}}D f\cdot D\kappa\right\vert\leq C\, E^{1+\eta}\,\rho^m\,\|D\kappa\|_{L^\infty},
\quad\text{for every }\;\kappa\in C_c^1(B_{6\rho},\R{n}).
\end{equation}
Then, putting together the previous estimates, we conclude that
\begin{align*}
\|\Delta \hat f\|_{L^{\infty}(B_{5\rho})}&=\sup_{\gamma\in C^1_c(B_{5\rho}),\|\gamma\|_{L^1}\leq1}\int D\hat f\cdot D\gamma
=\sup_{\gamma\in C^1_c(B_{5\rho}),\|\gamma\|_{L^1}\leq1}
\int D f\cdot D(\gamma*\varphi_{\rho})\\
&\stackrel{\mathclap{\eqref{e:step1}}}{\leq}\sup_{\gamma\in C^1_c(B_{5\rho}),\|\gamma\|_{L^1}\leq1} C\,E^{1+\eta}\rho^m\,\|D(\gamma*\varphi_\rho)\|_{L^\infty}
\leq C\, E^{1+\eta}\,\rho^m\,\|D\varphi_{\rho}\|_{L^\infty}\\
&\leq C\, E^{1+\eta}\, \rho^{-1}\leq C\,\rho^{(2-2\delta)(1+\eta)-1}.
\end{align*}
Therefore, \eqref{e:laplace} follows choosing $\delta$ sufficiently small
with respect to $\eta$.
For the proof of \eqref{e:laplace0}, it is enough to notice that, from
\eqref{e:variation} and Proposition~\ref{p:approx}, we get
\begin{align*}
&\int_{B_{5\rho}}\left\vert \int_{B_{\rho} (w)} Df(z)\cdot D\gamma (w-z)
\,dz\right\vert\,dw\\
\leq& C\int_{B_{5\rho}}
|D\gamma|*|Df|^3+C\int_{B_{5\rho}} |D\gamma|*{\bf 1}_{\R{m}\setminus
K}+C\int_{B_{5\rho}} |D\gamma|*\big( \mu\res(\R{m}\setminus K)\big)\\
\leq& \|D\gamma\|_{L^1} \left( C E^\eta\,\int_{B_{6\rho}}|Df|^2+ |B_{6\rho}\setminus K|
+ \mu(B_{6\rho}\setminus K)\right)
\leq C\, E^{1+\eta}\,\rho^m\,\|D\gamma\|_{L^1}.
\end{align*}
\end{proof}
Now we come to the $L^1$-estimate for the harmonic approximation $\bar f$.
\begin{propos}\label{p:L1}
Let $(p,8\rho,\pi)$ be admissible and $\bar{f}$ be as in ($I_2$).
Then, there exists $\alpha>0$ such that
\begin{equation}\label{e:L1prov}
\|\bar f-f\|_{L^1(B_{4\rho})}\leq C\, \rho^{m+3+\alpha}.
\end{equation}
\end{propos}
\begin{proof}
First we estimate the $L^1$ distance between $\bar f$ and $\hat f$.
Using the Poincar\'e inequality and a simple integration by parts, we infer that
\begin{equation*}
\|\bar f-\hat{f}\|_{L^1 (B_{4\rho})}^2\leq C \, \rho^{m+2}\,
\|\nabla (\bar f-\hat{f})\|_{L^2(B_{4\rho})}^2
= C \,\rho^{m+2} \int_{B_{4\rho}} \Delta \hat{f} \,(\bar{f} - \hat{f}),
\end{equation*}
from which
\begin{equation*
\|\hat{f}-\bar{f}\|_{L^1 (B_{4\rho})}\leq
C \,\rho^{2+m} \,\|\Delta \hat{f}\|_\infty
\stackrel{\eqref{e:laplace}}{\leq} C \,\rho^{m+3+\lambda}.
\end{equation*}
In order to prove \eqref{e:L1prov}, then it is enough to prove the following inequality,
\begin{equation}\label{e:Linfty}
\|\hat{f} - f\|_{L^1 (B_{4\rho})} \leq C \rho^{m+3+\alpha}.
\end{equation}
For every $z\in B_{4\rho}$, from the definition of $\hat{f}$ we have
\begin{equation}\label{e:explicit}
\hat{f} (z)-f (z) = \int \varphi_\rho (z-y) (f(y)-f (z))\, dy.
\end{equation}
To simplify the notation assume $z=0$ and rewrite \eqref{e:explicit} as
\begin{align*}
\hat{f} (0) -f (0) &= \int \varphi_\rho (y) \int_0^{|y|}
\frac{\partial f}{\partial r} \left( \tau \frac{y}{|y|}\right)\, d\tau\, dy
= \int \varphi_\rho (y) \int_0^{|y|} \nabla f \left(\tau\frac{y}{|y|}\right)
\cdot \frac{y}{|y|}\, d\tau\, dy\notag\\
&= \int \varphi_\rho (y) \int_0^1 \nabla f (\sigma y) \cdot y\, d\sigma\, dy
= \int \int_0^1 \varphi_\rho \left(\frac{w}{\sigma}\right)\, \nabla f (w) \cdot
\frac{w}{\sigma^{m+1}}\, d\sigma\, dw\notag\\
&= \int \nabla f (w) \cdot \underbrace{w \left(\int_0^1 \varphi_\rho
\left(\frac{w}{\sigma}\right)\,\sigma^{-m-1}\, d\sigma\right)}_{=: \Phi (w)}\, dw
\end{align*}
More generally, for every $z\in B_{4\rho}$, we have
$\hat{f} (z) -f (z)= \int \nabla f (w) \cdot \Phi (w-z)\, dw$
and
\begin{equation*}
\|\hat f-f\|_{L^1(B_4\rho)}=\int_{B_{4\rho}}\left\vert \int \nabla f (w) \cdot \Phi (w-z)\, dw\right\vert\,dz.
\end{equation*}
Since $\varphi$ is radial, the function $\Phi$ is a gradient.
Indeed, it can be easily checked that, for any $\psi$, the vector field $\psi (|w|)\, w$ is curl-free.
Moreover, $\supp (\Phi)$ is compactly contained in $B_\rho$.
Hence, we can apply \eqref{e:laplace0} and get
\begin{equation}\label{e:norm1}
\|\hat f-f\|_{L^1(B_4\rho)}\leq C\, E^{1+\eta}\,\rho^m\, \|\Phi\|_{L^1}.
\end{equation}
By a simple computation,
\begin{align*}
\|\Phi\|_{L^1}=\int_{\R{m}}\int_0^1 |w|\,\varphi
\left(\frac{w}{\rho\sigma}\right)\,\rho^{-m}\sigma^{-m-1}\, d\sigma\,dw
=\rho \int_{\R{m}}\int_0^1|y|\,\varphi(y)\,d\sigma\,dy\, .
\end{align*}
The last integral is a constant which depends only on $\varphi$.
Thus, \eqref{e:Linfty} follows from \eqref{e:norm1}.
\end{proof}
A simple consequence of the $L^1$-estimate is a comparison between
harmonic approximations at different scales.
\begin{corol}\label{c:alpha}
Assume $(p, 16\,r, \pi)$ is an admissible triple and
let $\bar f_1$ and $\bar f_2$ be as in $(I_2)$, with $\rho=r$ and $\rho=2\,r$ respectively.
Then, if $p=(q,u)\in \pi\times \pi^\perp$,
\begin{equation}\label{e:stima_alpha}
\sum_{\ell=0}^4 r^{\ell-3-\alpha}\|D^\ell \bar f_1-D^\ell \bar{f}_2\|_{C^0 (B^m_{3r/2} (q))}
\leq C .
\end{equation}
\end{corol}
\begin{proof}
It is enough to show that
\begin{equation}\label{e:L1bis}
\|\bar f_1-\bar f_2\|_{L^1(B_{2r})}\leq C\, r^{m+3+\alpha},
\end{equation}
because then the conclusion of the lemma follows easily from the classical
mean-value property of harmonic functions.
Clearly, from the admissibility of $(p,16\,r,\pi)$ and
Corollary~\ref{c:decay_everywhere},
it follows that $|\pi-\pi_p|\leq C\,r^{2-2\delta}$.
Hence, always by the same corollary
$E_2:=\textup{Ex}(T, B_{16 r}(p), \pi)\leq C r^{2-2\delta}$.
Then, in view of Proposition \ref{p:L1}, in order to show \eqref{e:L1bis},
it suffices to prove
\begin{equation}\label{e:L1prov2}
\|f_1-f_2\|_{L^1(B_{2r})}\leq C\, r^{m+3+\alpha}.
\end{equation}
Note first that $f_1$ and $f_2$ coincide
on a set $K$ with $|B_{2r} \setminus K|\leq C E_2^{1+\eta} r^m$.
Moreover, since the Lipschitz constants of $f_1$ and $f_2$ are bounded by
a universal constant $C$,
we have $|f_1 (z)- f_2 (z)|\leq C r$ for every $z\in B_{2r}$.
Therefore, we conclude \eqref{e:L1prov2} from
\begin{equation*
\|f_1-f_2\|_{L^1 (B_{2r})} \leq C r |B_{2r}\setminus K|\leq C \,r\, E_2^{1+\eta}\,r^m
\leq C \, r^{m+1+(1+\eta)(2-2\delta)}.
\end{equation*}
\end{proof}
\section{Proof of Proposition~\ref{p:main}}
The proof of \eqref{e:A} in Proposition~\ref{p:main} is given by a simple iteration
of Corollary \ref{c:alpha} on dyadic balls.
\begin{lemma}\label{l:beta}
Let $g_1, g_2$ be respectively the $(p, \rho, \pi)$- and the
$(p, 2^N\rho, \pi)$-interpolation (under the assumption of admissibility \eqref{e:good}).
Then, for $p=(q',u')\in \pi_0\times\pi_0^\perp$, it holds
\begin{equation}\label{e:A'}
\|g_1\|_{C^3} + \rho^{1-\alpha} \|D^4 g_1\|_{C^0}\leq C ,
\end{equation}
\begin{equation}\label{e:stima_beta}
|D^3 g_1 (q') -D^3 g_2 (q')| \leq C (2^N \rho)^\alpha.
\end{equation}
\end{lemma}
\begin{proof}
Recalling Lemma \ref{l:rotation}, it suffices to show \eqref{e:A'} for the function
$\bar f_1$.
Let $n_0$ be the biggest integer such that $2^{n_0+3}\rho\leq \frac{1}{2}$ and
for every $k\leq n_0-1$ set $r_k=2^k\,\rho$.
If $\pi_k$ is such that $\textup{Ex}(T, B_{8 r_k}, \pi_k)=\textup{Ex}(T, B_{8 r_k})$,
then, by Corollary~\ref{c:decay_everywhere} (b), $|\pi-\pi_k|\leq C\,r_k^{1-\delta}$.
Hence, we conclude that the admissibility condition \eqref{e:good} holds with $r=r_k$,
so that we can consider the approximation $\bar f_k$ as in ($I_2$) for $r_k$.
From Corollary \ref{c:alpha}, we get
\begin{equation}\label{e:stima_alpha2}
\|D^\ell \bar f_k-D^\ell \bar{f}_{k+1}\|_{C^0 (B^m_{3r_k/2} (q))}
\leq C r_k^{3+\alpha-\ell}\leq
C\, 2^{-(n_0-k)(3+\alpha-\ell)} \quad\text{for }\;\ell\in \{0,1,2,3,4\}.
\end{equation}
Note that the series $\sum_i 2^{-i(3+\alpha-\ell)}$ is summable for $\ell\leq 3$.
Therefore, $\|\bar f_1\|_{C^3}\leq C+\|\bar f_{n_0}\|_{C^3}$.
On the other hand, since $r_{n_0}>1/32$, it is easy to see that $\|\bar f_{n_0}\|_{C^3}\leq C$
for some universal constant $C$, so that $\|\bar f_1\|_{C^3}\leq C$.
In the same way we have $\|D^4 \bar f_1\|_{C^0}\leq C\, \rho^{\alpha-1}$.
Then, \eqref{e:A'} follows from Lemma \ref{l:rotation} (c) .
Finally, Corollary \ref{c:alpha} obviously implies that
\begin{equation}\label{e:ancoraL^1}
\int_{B^m_{3r_k/2} (q)} |\bar f_k - \bar f_{k+1}| \leq C r_k^{m+3+\alpha}.
\end{equation}
Hence, using again Lemma \ref{l:rotation}, we conclude
\begin{equation}\label{e:ancoraL^1_2}
\int_{B^m_{r_k} (q')} |g_k - g_{k+1}| \leq C r_k^{m+3+\alpha}.
\end{equation}
Let $P_k$ and $P_{k+1}$ be the third order Taylor polynomials at $q'$ of $g_k$ and $g_{k+1}$.
From the estimate $\|D^4 g_k\|, \|D^4 g_{k+1}\|\leq C r_k^{\alpha-1}$
and \eqref{e:ancoraL^1_2}, we easily infer
\begin{equation*
\int_{B^m_{r_k} (q')} |P_k - P_{k+1}| \leq C\, r_k^{m+3+\alpha}.
\end{equation*}
Hence, applying Lemma \ref{l:poly}, we then get
\begin{equation}\label{e:iterateD^3}
|D^3 g_k (q')-D^3 g_{k+1} (q')| =
|D^3 P_k (q')-D^3 P_{k+1} (q')| \leq C\, r_k^\alpha.
\end{equation}
Arguing as above, the estimate \eqref{e:stima_beta}
follows from \eqref{e:iterateD^3} and a simple iteration.
\end{proof}
The final step in the proof of Proposition~\ref{p:main} consists in comparing two different
interpolating functions defined at the same scale but for nearby balls and varying planes
$\pi$.
We do this in the following two separate lemmas.
\begin{lemma}\label{l:gamma}
Let $g_1$ and $g_2$ be the $(p,\rho, \pi)$- and $(p, \rho, \pi')$-interpolating functions
where as usual $(p,8\rho,\pi)$ and $(p,8\rho, \pi')$ are admissible.
Then,
\begin{equation}\label{e:gamma}
\sum_{\ell=0}^3\rho^{\ell-3-\alpha}\|D^\ell g_1-D^\ell g_2\|_{C^0 (B^m_\rho (q))} \leq C.
\end{equation}
\end{lemma}
\begin{proof}
As before, we first show that
\begin{equation}\label{e:g1g2}
\|g_1-g_2\|_{L^1 (B_{3/2 \rho} (q))}\leq C \rho^{m+3+\alpha}.
\end{equation}
Denote by $f_1, f_2$ the Lipschitz approximations given by Proposition~\ref{p:approx}
in the coordinates associated to $\pi, \pi'$ and
let $h_1, h_2: B_\rho (q)\to \pi_0^\perp$ be the Lipschitz functions whose graphs coincide
with the graphs of ${f}_1$ and $f_2$ respectively.
From Lemma~\ref{l:rotation} and Proposition~\ref{p:L1}, we have
\begin{equation*
\|g_i-h_i\|_{L^1 (B_{3/2 \rho} (q))}\leq
\|f_i-\bar{f}_i\|_{L^1 (B_{2\rho} (q_i))}\leq C \,\rho^{m+3+\alpha},
\end{equation*}
where $(q_1, u_1)$, $(q_2, u_2)$ and $(q,u)$ are the coordinates of $p$ in
$\pi\times \pi^\perp$, $\pi'\times \pi'^\perp$ and $\pi_0\times \pi_0^\perp$ respectively.
Therefore, for \eqref{e:g1g2} it is enough to show
\[
\|h_1-h_2\|_{L^1 (B_{3/2 \rho} (q))}\leq C \rho^{m+3+\alpha}.
\]
To see this, consider the set $A=\{h_1\neq h_2\}$.
From Proposition~\ref{p:approx} if follows that
\begin{equation*
|A|\leq \mathcal{H}^m \big(\textup{graph} (h_1)\,\triangle\, \textup{graph} (h_2)\big)\leq C \rho^{m+2+\alpha}.
\end{equation*}
Then, if $x\in A$ and $y\in B_{3\rho/2}\setminus A$,
since $h_1 (y)=h_2 (y)$ and ${\rm {Lip}} (h_i)\leq C$, we have
\[
|h_1 (x) - h_2 (x)|\leq |h_1 (x)-h_1 (y)|+|h_2 (y) - h_2 (x)| \leq C|y-x|\leq C \rho,
\]
from which $\|h_1-h_2\|_{L^1 (B_{3/2 \rho} (q))}\leq C\,r\,|A|\leq C\, \rho^{m+3+\alpha}$.
From \eqref{e:g1g2} we are ready to conclude.
Let $x\in B_\rho (q)$ and $P_i$ be the third order Taylor expansions of $g_i$ at $x$.
Arguing as in Lemma \ref{l:beta}, we conclude
\[
\|P_1-P_2\|_{L^1 (B_{\rho/2} (x))}\leq C \rho^{m+3+\alpha}.
\]
Using Lemma \ref{l:poly} we then conclude
\begin{equation}\label{e:scaled_est}
|D^k P_1 (x) - D^k P_2 (x)|\leq C \rho^{3-k+\alpha}
\quad \text{for }\;k\in \{0,1,2,3\}.
\end{equation}
On the other hand, since $D^k P_i (x)= D^k g_i (x)$,
\eqref{e:scaled_est} implies the desired estimates.
\end{proof}
\begin{lemma}\label{l:delta}
Let $g_1$ and $g_2$ be, respectively, the $(p,\rho, \pi)$- and $(p', \rho, \pi)$-interpolating
functions, where $(p,\rho,\pi)$ and $(p',\rho,\pi)$ are admissible.
Assume that $p=(q,u)$, $p'=(q',u')$ with $|q-q'|\leq \rho/16$.
Then,
\begin{equation}\label{e:delta}
\sum_{\ell=1}^4 \rho^{\ell-3-\alpha}\|D^\ell g_1-D^\ell g_2\|_{C^0 (B^n_\rho (q)\cap B^n_\rho (q'))} \leq C.
\end{equation}
\end{lemma}
The proof of this lemma exploits only a portion of the same computations used for Lemma~\ref{l:gamma}
and is left to the reader.
The proof of \eqref{e:B} follows straightforwardly from Lemma~\ref{l:gamma} and Lemma~\ref{l:delta};
while the proof of \eqref{e:C} is given below.
\begin{proof}[Proof of \eqref{e:C}]
Consider $R:= 16\,|q-q'|$ and let $h$, $k$ and $h'$ be the $(q, R, \pi)$-,
$(q, R, \pi')$- and $(q', R, \pi')$-interpolations, respectively.
By Corollary \ref{c:decay_everywhere}, if $|q-q'|$ is small enough,
we can apply Lemma \ref{l:gamma} and Lemma \ref{l:delta} to conclude that
$$
|D^3 h (q) - D^3 k (q)| + |D^3 k (q') - D^3 h' (q')|\leq C R^\alpha.
$$
On the other hand, by \eqref{e:A}, $\|D^4 k\|\leq C R^{\alpha-1}$,
and so $|D^3 h (q) - D^3 h' (q')|\leq R^\alpha$.
Since by Lemma \ref{l:beta} we know that $|D^3 g (q) - D^3 h (q)|\leq C
R^\alpha$
and $|D^3 g' (q') - D^3 h' (q')|\leq C R^\alpha$, the desired conclusion
follows.
\end{proof}
|
\section{Introduction}
\label{sec:intro}
The relation between the stellar mass of galaxies and their host dark-matter haloes
has become a key point of reference for many different theoretical and observational studies.
It summarizes in a simple way the complexity of galaxy formation physics when evolved
within growing dark-matter structure. A special attention was given
to the relation between galaxies and their host \emph{subhaloes}, which are the sub-structure bound density
peaks inside a halo.
The `abundance matching' (hereafter ABM) methodology is an important theoretical tool for
constraining the mass relation between galaxies and their host subhaloes
\citep{Kravtsov04,Vale04,Conroy06,Shankar06,Guo10a,Moster10,Behroozi10}.
In this simple empirical approach one assigns a model galaxy to each subhalo within a
cosmological $N$-body simulation. Assuming there is a one-to-one, monotonic
relation between the stellar mass ($m_{\star}$) and the subhalo mass
($M_{\rm infall}$, see section \ref{sec:subhalos}), the abundance of galaxies and subhaloes can be matched,
yielding a unique relation between $m_{\star}$ and $M_{\rm infall}$.
Surprisingly, this simple model provides a good fit to the observed clustering
properties of galaxies.
The ABM approach thus offers a practical solution to the relation between
subhaloes and galaxies, without going into the complex details of
galaxy formation physics. It can be used to constrain the mass
relation at various redshifts, to predict the star-formation rate in
galaxies \citep{Conroy09}, to study the
merger-rates of galaxies \citep{Hopkins10}, and to interpret
high-resolution hydrodynamical simulations \citep{Sawala11}.
In comparison to models based on the halo occupation distribution
\citep[][and references therein]{Berlind02,Cooray02,Tinker05,Zehavi05},
ABM uses explicit information on the location and mass of subhaloes,
decreasing the number of free parameters needed in the model.
The success of ABM is intriguing and raises several interesting questions:
Is it based on the only possible set of assumptions that can reproduce
the abundance and clustering of galaxies?
Do we miss models that result in a different $m_{\star}-M_{\rm infall}$ relation? What are the
important assumptions made by ABM? How can we explore these
assumptions and test to what level they are constrained by observations?
In this work we try to address these questions. We specifically focus our results
on the freedom in the $m_{\star}-M_{\rm infall}$ relation.
In ABM, the stellar mass is assigned to each subhalo according to $M_{\rm infall}$.
For a satellite subhalo\footnote{We define `satellite subhaloes' as all the substructure clumps
within a {\scshape fof~} group, except the most massive one.}, $M_{\rm infall}$ is defined as the mass at the last time
it was the most massive substructure within its {\scshape fof~} group. This is
a reasonable assumption because the subhalo mass of satellite galaxies can be
significantly stripped after falling into a larger dark-matter halo \citep[e.g.][]{Zentner05,vdBosch05},
more so than the stellar mass of its galaxy \citep[for stellar stripping see][]{Monaco06,Purcell07,Conroy07a,Yang09a}.
On the other hand, for a central subhalo, $M_{\rm infall}$ is defined as its current mass.
\citet{Wang06} found that the relation between $M_{\rm infall}$ and the stellar mass
of galaxies is tight in semi-analytic models, justifying the above
definition of $M_{\rm infall}$.
There are various other assumptions made by ABM, which are mainly related to the treatment of
satellite galaxies. In a recent paper \citep[][hereafter paper I]{Neistein11}, we have
examined these assumptions using the semi-analytic model (SAM) of \citet[][]{Neistein10}.
Although ABM models assume that for a given $M_{\rm infall}$, the stellar mass of central and satellite galaxies
is the same, there are various effects within SAMs that violate this assumption.
First, the relation between stellar mass and subhalo mass evolves with
redshift for central galaxies, affecting satellite galaxies at the time of infall \citep[the
typical infall time is a few Gyrs ago;][]{Wang07}.
Second, the stellar mass of satellite galaxies might be
different already at the time of infall from that of central galaxies at the same time. This is
because galaxies that join larger groups as satellites are located
in large-scale environment of higher density than the average. Consequently, the
dark-matter merger-histories of these galaxies are already
different at early epochs \citep[see also][]{Gao05,Harker06}.
Third, once a galaxy becomes a satellite, its stellar mass might
still grow. This is especially true for models in which
gas stripping in satellite galaxies is modeled
on time-scales of a few Gyrs \citep{Weinmann10}.
All the effects above are consistent with various studies about the properties
of central and satellite galaxies
\citep{Weinmann06,vdLinden07,Khochfar08,Skibba09,Pasquali09,Skibba11}.
In paper I, we showed that the effects above can change not only
the $m_{\star}-M_{\rm infall}$ relation for satellite galaxies, but also the auto-correlation function (CF) of
galaxies.
If satellite and central galaxies with a given $M_{\rm infall}$ are randomly redistributed in a SAM, the CFs can vary by
up to a factor of four. Even when just redistributing satellite and central galaxies among themselves, the
modifications in the CFs can reach a factor of 2 for massive galaxies.
We have shown in paper I that the CFs of SAM
galaxies can be reproduced accurately by the ABM approach only when the
stellar mass of satellite galaxies is assumed to depend on both
$M_{\rm infall}$ and the host halo mass at $\emph{z}=0$. This finding is very
useful as it saves us the complex modeling of $m_{\star}$ as a function of
the various effects mentioned above.
In this paper we
make use of the conclusions made in paper I, and add more ingredients to an
ABM model. We would like to study the influence of the new ingredients on the relation
between $M_{\rm infall}$ and $m_{\star}$, while the models are constrained to fit the observational data.
We allow the stellar mass of satellite galaxies to depend on both the $M_{\rm infall}$ and
the host halo mass, and to deviate from the behaviour of central galaxies.
In addition, we use two ingredients for modeling the location
and abundance of satellite galaxies: Subhaloes
that were lost by the cosmological simulation at high
redshift, but can still host galaxies at $\emph{z}=0$, are identified
using an estimate for dynamical friction. The location of unresolved
subhaloes is fixed using either the location of the most bound particle of the
last identified subhalo, or using an analytical model with a free parametrization.
Once the population of galaxies is broken into two sub-populations,
the model cannot be constrained by the abundance of galaxies alone. The
number of models that can fit the observed stellar mass function of
galaxies is infinite. We therefore check a very large number of models ($\sim10^{12}$)
and look for those that fit both the observed stellar mass function,
and the observed auto-correlation function of galaxies. We will show
that even when using these two constraints, there is a
significant amount of freedom in the relation between $m_{\star}$ and
$M_{\rm infall}$. Satellite galaxies might occupy a significantly lower value
of $M_{\rm infall}$ and thus might be more abundant than central galaxies at a fixed $m_{\star}$. The
dependence of satellite galaxies on the halo mass is able to compensate for
this effect in terms of clustering, and is crucial for fitting the observed auto-correlation
functions.
The observational constraints used in this work include the stellar mass function of galaxies, as
derived by \citet{Li09}. For the CFs of galaxies we use the
methodology presented by \citet{Li06},
using the same stellar masses as in \citet{Li09}.
These stellar masses are based on redshift and
five-band photometric data, as described in detail by \citet{Blanton07}. The galaxy sample is based
on the final data release \citep[DR7;][]{Abazajian09} of the Sloan Digital Sky Survey
\citep[SDSS;][]{York00}. Although there might be both random and systematic
deviations in the stellar masses used here, the two
observational constraints are self-consistent.
This paper is organized as follows. Section \ref{sec:formalism} describes in detail
the approach we use in this work. We elaborate on the new ingredients used here,
and the way we implement the models. In section \ref{sec:examples} we demonstrate the
results of our formalism by showing a few different models
that fit the data well. Our method for scanning the parameter space is discussed in section
\ref{sec:results}, where we explain how the parameter space is sampled, and how we
select good models. In this section we also show the properties of the successful
models. Lastly, we summarize our results and discuss them in section \ref{sec:discuss}.
\section{The formalism}
\label{sec:formalism}
In this section we describe the formalism developed for modeling the abundance and
clustering of galaxies. In general, we assume that the stellar mass of galaxies depends
solely on the properties of the host dark-matter haloes. This assumption is similar to
the abundance matching approach \citep[e.g.][]{Vale04}. However, we significantly extend
the ingredients of the model with respect to previous studies. This is
motivated by the analysis done in paper I, based on the semi-analytic models of
\citet{Neistein10}. The additional ingredients
are related to the assumptions regarding satellite galaxies: their abundance,
location, and the effect of the host halo mass on their stellar mass.
We start by summarizing the various ingredients of the formalism, a detailed description of
each component is given in the following sub-sections.
\begin{enumerate}
\item All the subhaloes from a cosmological $N$-body simulation at $\emph{z}=0$ are selected.
\item In addition, we use subhaloes that have been merged into a
bigger structure at high redshift, but might host galaxies
that survive until $\emph{z}=0$. This is due to the effects of dynamical
friction and stripping on galaxies, which are not included in the $N$-body simulation.
\item We use a freely tunable prefactor in the dynamical friction formula.
\item We assign one galaxy with a specific stellar mass ($m_{\star}$) to each subhalo from
the full sample defined in items (i) \& (ii).
\item For central subhaloes within their {\scshape fof~} group, we assume
that the stellar mass of galaxies depends on the host subhalo mass at $\emph{z}=0$
only ($M_{\rm infall}$).
\item For satellite subhaloes, we assume
that the stellar mass of galaxies depends on two parameters: the host subhalo mass at
the time of infall, and the host halo mass at $\emph{z}=0$.
\item We allow the dependence of stellar mass on the subhalo mass to be different
for central and satellite subhaloes.
\item The location of subhaloes that are not found in the
simulation at $\emph{z}=0$ is set by either the location of their most
bound particle, or by an analytical model.
\end{enumerate}
Items (iii), (vi), (vii), (viii) in the list are new with respect to
previous ABM models, and their influence on clustering and abundance
of galaxies has not been tested before.
\subsection{Subhaloes and merger-trees}
\label{sec:subhalos}
We use merger trees extracted from the Millennium $N$-body simulation \citep{Springel05}.
This simulation was run using the cosmological parameters $(\Omega_{\rm m},\,\Omega_{\Lambda},\,h,\,\sigma_8)=
(0.25,\,0.75,\,0.73,\,0.9)$, with a particle mass of $8.6\times10^8\,\,\ifm{h^{-1}}{M_{\odot}}$ and a box size
of 500 $h^{-1}$Mpc. The merger trees are based on snapshots spaced by $\sim$250 Myr, linking
\emph{subhaloes} identified by the \textsc{subfind} algorithm \citep{Springel01}.
Subhaloes are defined as the bound
density peaks inside {\scshape fof~} groups \citep{Davis85}. More details on the simulation and the
subhalo merger-trees can be found in \citet{Springel05} and \citet{Croton06}.
The subhalo mass, $M_h$, corresponds to the number of particles inside a subhalo,
as identified by \textsc{subfind}. The \emph{infall} mass of the subhalo,
$M_{\rm infall}$, is defined as
\begin{equation}
\label{eq:minfall}
M_{\rm infall} = \left\{ \begin{array}{ll}
M_h & \textrm{if central within its {\scshape fof~} group}\\ \;\;\;\
& \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \\
M_{h,p}(\z_{\rm infall}) & \textrm{otherwise}
\end{array} \right.
\end{equation}
Here $\z_{\rm infall}$ is the lowest redshift at which the main progenitor\footnote{
Main-progenitor histories are derived by following back in time the most massive
progenitor in each merger event.}
of the subhalo $M_h$ was the most massive within its {\scshape fof~} group, and $M_{h,p}$ is
the main progenitor mass at this redshift.
In addition to $M_{\rm infall}$ we will use the \emph{halo} mass
$M_{200}$, which is defined as the mass within the radius where the halo has an
over-density 200 times the critical density of the simulation.
In general, the mass $M_{200}$ should include both the central subhalo within a group,
and all of its satellite subhaloes. However, due to the spherical symmetry forced on
$M_{200}$ and the specific over density being used here, it might deviate from the
exact {\scshape fof~} group mass. In what follows, we will refer to $M_{200}$ as the `halo mass' to
indicate that this mass is computed over a larger spatial region than the subhalo mass.
For satellite subhaloes, we assign the same value of $M_{200}$ as
is computed for their central subhalo
within the same group. Due to the effects above, the value of
$M_{\rm infall}$ is often similar or higher than $M_{200}$ for central subhaloes\footnote{
For central subhaloes, $M_{200}$ is slightly smaller than $M_{\rm infall}$. The average
difference is $\sim0.08$ dex, with an RMS scatter of $\sim0.06$ dex}.
\subsection{Satellite subhaloes: dynamical friction and location}
Satellite subhaloes are defined as all subhaloes inside
a {\scshape fof~} group except the most massive (central) subhalo. Within the $N$-body
simulation, satellite subhaloes lose their mass
while falling into a bigger subhalo, and thus might fall below the resolution limit
used in the \textsc{subfind} algorithm (20 particles for the Millennium simulation used here).
However, the galaxies residing inside these satellite subhaloes might live
longer, as they are more dense, and thus less vulnerable to
stripping. This effect is significant even for relatively high resolution
cosmological simulations, as was shown in paper I. It can
modify the abundance of subhaloes even at two orders of magnitude above
the minimum subhalo mass resolved by the simulation.
In order to take this effect into account, we model the time it takes the galaxy to fall into
the central galaxy by dynamical friction. At the last time the satellite subhalo
is resolved we compute its distance from the central subhalo ($r_{\rm sat}$),
and estimate the dynamical friction time using the formula of \citet{Binney87},
\begin{equation}
t_{\rm df} = \alpha_{\rm df} \cdot \, \frac{1.17 V_v r_{\rm sat}^2}{G M_{h,2}\ln\left(
1+ M_{h,1}/M_{h,2} \right) } \, .
\label{eq:t_df}
\end{equation}
Here $M_{h,1}$ is the mass of the central subhalo, $V_v$ is its virial
velocity, and
$M_{h,2}$ is the mass of the satellite subhalo.
Once a satellite subhalo falls together with its central
subhalo into a larger group, we update $t_{\rm df}$ for both
objects according to the new central subhalo.
The dynamical friction estimate is computed using the mass of subhaloes only, before galaxies
are being modeled. Therefore, the formula deviates from its proper definition,
as it does not include the effect of the galaxy mass on dynamical friction. However,
since this formula uses various simplified assumptions with a larger
uncertainty \citep[for example,
the exact trajectories of the satellite subhaloes are ignored, see][]{Boylan08},
one general constant, $\alpha_{\rm df}$, is being used here to absorb all the related
uncertainties. Semi-analytic models often use
$\alpha_{\rm df}\sim 2-3$ \citep{DeLucia07,Khochfar09,Neistein10},
in agreement with more detailed studies of dynamical friction processes
\citep{Colpi99,Boylan08,Jiang08,Mo10}. Since we ignore contribution due to the galaxy
mass, we allow more freedom in $\alpha_{\rm df}$ as is discussed below.
To summarize, three types of subhaloes exist in our models:
\begin{itemize}
\item {\bf central subhaloes:} most massive subhaloes within their {\scshape fof~} group.
\item {\bf satellite subhaloes:} all subhaloes except central subhaloes.
\item {\bf unresolved subhaloes:} subhaloes that were last identified at high
$\emph{z}$, and are added according to the dynamical friction formula
above. All the unresolved subhaloes are also satellite subhaloes.
\end{itemize}
The location of unresolved subhaloes is not given by the simulation. We therefore
need to estimate the location of the galaxies that are assigned to these subhaloes.
This issue was treated by e.g. \citet{Croton06} who used the location
of the most bound particle of the last identified subhalo, as is given by the dark-matter
simulation at $\emph{z}=0$. The same location is adopted here as well. However, we would like
to take into account the uncertainty involved in this method. Obviously,
one dark-matter particle inside a collisionless $N$-body simulation cannot
mimic accurately the location of an extended galaxy, with a possibly very different
mass. We therefore use the following analytical model as an additional option:
\begin{equation}
r= r_{\rm sat} \left( 1-\tau^p \right) ^{1/q} \,.
\label{eq:sat_loc}
\end{equation}
Here $\tau$ is the fraction of time spent out of all the estimated dynamical friction time
until $\emph{z}=0$, $r$ is the distance we adopt at $\emph{z}=0$ from the central subhalo, and $p,q$
are constants. In appendix \ref{sec:sat_loc} we derive this model by following the angular momentum of a particle
inside a spherical gravitational potential of the type $\rho\propto r^{-\gamma}$. We
show that for a specific values of $p$ and $q$, the model is able to reproduce the location of subhaloes
as is modeled by the most bound particle. This parametrization
enables us to check various modifications to the location of unresolved subhaloes.
\subsection{Counting subhaloes}
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig1.eps,width=9cm} }}
\caption{The abundance of satellite subhaloes, $\phi_s$, as a function of $M_{\rm infall}$ and
$M_{200}$, using the full Millennium simulation with $\alpha_{\rm df}=3$. \emph{Contours} show constant values
of $\phi_s$ as labeled in units of Log Mpc$^{-3}$ dex$^{-2}$.
$\phi_s$ is bigger than zero only below the \emph{dotted-dashed} line.
The \emph{thin dashed} line corresponds to $M_{\rm infall}=M_{200}$ and is shown as a reference.}
\label{fig:nsat_2d}
\end{figure}
The usual way to implement ABM is to first populate subhaloes
with galaxies, and only then to compute the auto correlation functions (CFs) of
galaxies. In our approach there are many possible models, mainly
depending on the different mass relations for satellite and central
galaxies. Since computing the statistics of pairs is the most demanding
computational step, it is not possible to scan a significant number of
models in the usual technique. Here we present a new way of
computing the CFs which is extremely efficient when many
models are needed. We first compute the statistics
of \emph{subhalo} pairs, and save them as a function of $M_{\rm infall}$ and
$M_{200}$. Only then are the galaxy CFs computed. In this section we explain how the statistics of
subhaloes is defined and computed, the next subsection discusses the way this
information is used to model galaxies.
We want to constrain our models against the stellar mass function of galaxies. In order to do
so we will use the mass function of central and satellite subhaloes:
\begin{equation}
\phi_c(M_{\rm infall})=\frac{1}{V}\frac{{\rm d}{N^c}}{{\rm d}{\logM_{\rm infall}}} \,,
\end{equation}
\begin{equation}
\phi_s(M_{\rm infall},M_{200})=\frac{1}{V}\frac{{\rm d}^2{N^s}}{{\rm d}{\logM_{\rm infall}} \, {\rm d}{\logM_{200}}} \,.
\end{equation}
Here $V$ is the volume of the simulation box, and $N^c,N^s$ are the numbers of central and satellite
subhaloes respectively. In Fig.~\ref{fig:nsat_2d} we show the two dimensional mass function $\phi_s$,
for a model with $\alpha_{\rm df}=3$. A similar behaviour as presented here is valid for
$0.1\leq\alpha_{\rm df}\leq10$, where the low-mass contour can vary by $\sim0.5$ dex, and the contour
of massive subhaloes is hardly affected.
In order to compute the CFs of galaxies, we start by computing the number
of subhalo pairs, $N_p$, for the total sample of subhaloes. In case both subhaloes within the pair are central
subhaloes, we count this pair into $\psi_{cc}$:
\begin{equation}
\psi_{cc}(M_{\rm infall}^1,M_{\rm infall}^2, r ) = \frac{1}{V^2} \frac{{\rm d}^3{N_p^{cc}}}{{\rm d}{\logM_{\rm infall}^1}\, {\rm d}{\logM_{\rm infall}^2}\,{\rm d}{\log{r}}} \,.
\label{eq:psi_cc}
\end{equation}
Here $M_{\rm infall}^1$, $M_{\rm infall}^2$ are the infall mass of the first and second
subhaloes in the pair, and $r$ is the distance between these subhaloes within
the $x$-$y$ plane. Distances are computed only within the $x$-$y$ plane
in order to compute the \emph{projected} auto-correlation function,
as is described below. In practice we divide the range in Log$M_{\rm infall}$ and
Log($r$) into 100 and 50 bins respectively, and save $\psi_{cc}$ as a multi-dimensional
histogram.
In a similar way we define the pair statistics of central-satellite,
and satellite-satellite subhaloes,
\begin{equation}
\psi_{ss}(M_{\rm infall}^1,M_{\rm infall}^2, M_{200}^1, M_{200}^2, r ) \,,
\end{equation}
\begin{equation}
\psi_{cs}(M_{\rm infall}^1,M_{\rm infall}^2, M_{200}^2, r ) \,.
\label{eq:psi_cs}
\end{equation}
Note that for satellite subhaloes the number of pairs
is saved as a function of both $M_{\rm infall}$ and $M_{200}$.
This is done in order to properly model the dependence of
stellar mass on $M_{200}$. It should be emphasized that the statistics
of satellite subhaloes, i.e. $\phi_s$, $\psi_{ss}$, $\psi_{cs}$,
depend on the dynamical friction constant, $\alpha_{\rm df}$, and on the
location we adopt for unresolved subhaloes ($p$ and
$q$ from Eq.~\ref{eq:sat_loc}).
\subsection{Definition of a model - domains in stellar mass}
\label{sec:def_model}
Previous studies have often used an analytic functional form to describe the
relation between $M_{\rm infall}$ and $m_{\star}$. For example, \citet{Moster10}
suggested:
\begin{equation}
m_{\star} = f \left( M_{\rm infall} \right) = c_1 M_{\rm infall} \left[ c_2 M_{\rm infall}^{c_3} +
M_{\rm infall}^{c_4} \right]^{-1} \,,
\end{equation}
where $c_i$ are all constants. A straight forward way
to extend this approach here would be to parameterize $f$ as a function of two
variables, $m_{\star}=f(M_{\rm infall},M_{200})$, and to use a different set of
parameters for modeling central and satellite galaxies. However, this
approach requires a priori knowledge of $f$, and the resulting solutions
might be restricted by the specific functional form chosen.
This is especially true when $f$ is allowed to be different for
satellite and central galaxies, so the freedom in its functional
shape might be larger than what was found in previous studies (see the examples in
section \ref{sec:examples} below).
Here we suggest a new way to parameterize the relation between
$M_{\rm infall}$ and $m_{\star}$, which is motivated by the observational data.
The observed CFs of galaxies are computed over
mass bins of width 0.5 dex in stellar mass (hereafter
`domains'). For example, the CF for small mass galaxies is based
on galaxies with stellar mass in the range
\begin{equation}
\textrm{Domain 1: } \,\, [10^{9.27},\,\, 10^{9.77}] \,\, M_\odot\,.
\end{equation}
In order to model the CF in this range we only
need to know the values of $M_{\rm infall},M_{200}$ that correspond to
$m_{\star}=10^{9.27},10^{9.77}$. These are defined by:
\begin{equation}
\label{eq:domain_bound}
f\left(\overline{M}_{\rm infall},\overline{M}_{200} \right) = 10^{9.27} \,, \,\,\, f\left(\widetilde{M}_{\rm infall},\widetilde{M}_{200}\right) = 10^{9.77} \,.
\end{equation}
Such constraints imply that in case of a smooth and monotonic $f$, the
'boundaries' of the domain correspond to curves within the $M_{\rm infall},M_{200}$
plane. The relation that defines a boundary can thus be written as:
\begin{equation}
\label{eq:boundary}
\overline{M}_{\rm infall}=U_1\left( \overline{M}_{200} \right) \,,\,\, \widetilde{M}_{\rm infall}=U_2\left( \widetilde{M}_{200} \right) \,,
\end{equation}
where $U_i$ is the curve function.
All subhaloes that are located between $U_1$ \& $U_2$
(their $M_{\rm infall},M_{200}$ masses follow the constraint $U_1(M_{200})\!\leq\!M_{\rm infall}\!\leq\!U_2(M_{200})$)
should contribute galaxies to the CF of the first domain.
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig2.eps,width=9cm} }}
\caption{
The dependence of $m_{\star}$ on ($M_{\rm infall}$,$M_{200}$) for satellite galaxies according to model $A$.
This model uses $\alpha_{\rm df}=3,p=0.5,q=0.8$.
The solid lines show $U_i^s$ that obey the equation $f\left(U_i^s(M_{200}),M_{200}\right)
=m_{\star,i}$, where $m_{\star,i}=$ 9.27, 9.77, 10.27, 10.77, 11.27,
11.77 in units of $\logM_\odot$.
Diamonds are placed at the
median $M_{\rm infall}$ values along each line. The dotted-dashed line and the thin dashed
line are the same as in Fig.~\ref{fig:nsat_2d}.
}
\label{fig:model_a1}
\end{figure}
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig3.eps,width=9cm} }}
\caption{
The relation between $m_{\star}$ and $M_{\rm infall}$ in model $A$.
For satellite subhaloes we plot various values of $M_{\rm infall}$ per a given $m_{\star}$, including
median (thick dashed lines), one standard deviation from the median (thin dashed lines), and
the full range (thin dotted-dashed lines).
The thin solid line corresponds to the universal baryonic fraction, $m_{\star}=0.17M_{\rm infall}$.
Model $A$ violates this fraction limit, and thus will not be used in the rest of the paper.}
\label{fig:model_a2}
\end{figure}
The standard abundance matching approach assumes that $m_{\star}$ depends
only on $M_{\rm infall}$. In terms of our language, this means that the
functions $U_i$'s are all constants, with no dependence on $M_{200}$.
In this work, we extend this assumption in two ways: we assume that
$U_i$ might be a power-law for satellite galaxies\footnote{For central galaxies the values of
$M_{200}$ are always very similar to $M_{\rm infall}$ so there is no added value in allowing the stellar mass
of central galaxies to depend on $M_{200}$.}, and that for each domain there
might be a different $U_i$ for central and satellite galaxies. This can be
summarized as:
\begin{eqnarray}
\label{eq:alpha_i}
\textrm{Central subhaloes:} & & U_i^c = \alpha_i \, ,\\ \nonumber
\textrm{Satellite subhaloes:} & & U_i^s = \beta_i M_{200} ^{\delta_i} \,.
\end{eqnarray}
Since the observed CFs are based on 5 domains in stellar mass, we
need to specify $U_i$'s at 6 domain boundaries. Our model therefore
includes 6 free parameters ($\alpha_i$) for central subhaloes,
and additional 12 parameters ($\beta_i,\delta_i$) for satellite
galaxies. We choose to limit our models only to power-law dependence
on $M_{200}$ following the results of paper I, and as a first order approximation. It will allow us to test
how important the standard assumption of constant $U_i$ is.
It might well be that a more complex behaviour would add a
significant amount of freedom to the models.
In order to model the stellar mass function, we
will use the same domains, and demand that the stellar
mass function will be reproduced once integrated over each domain. This guarantees
that in case it is needed, a detailed solution of the type $m_{\star}=f(M_{\rm infall},M_{200})$ exist.
However, since our models allow the mass relation above to deviate between central and
satellite galaxies, the detailed behaviour of $m_{\star}$ within each
domain is not well constrained. In general, there might be many
different interpolations of the kind $m_{\star}=f(M_{\rm infall},M_{200})$ within each
domain (between adjacent $U_i$'s). These will not change the computed CF, and will fit the
observed stellar mass functions. The range in stellar
mass for each domain is relatively small, so this freedom is negligible in comparison to the
results we will show below.
In general, modeling a scatter in $m_{\star}$ for a given $M_{\rm infall}$ and $M_{200}$ is possible
within our formalism. However, it demands a detailed knowledge of the functional
form, $m_{\star}=f(M_{\rm infall},M_{200})$. This addition does not allow us to scan the different models
in a very efficient way. We therefore do not treat such a scatter in this work.
Nonetheless, the dependence of $m_{\star}$ on $M_{200}$ for satellite galaxies
results in a variation in $m_{\star}$ as a function of $M_{\rm infall}$. This effect will
be discussed below.
\subsection{How to compute CFs?}
\label{sec:model_howto}
To summarize, each model in this work is defined by the following parameters:
\begin{description}
\item[a)] The value of $\alpha_{\rm df}$ used in the dynamical
friction formula, as defined by Eq.~\ref{eq:t_df}.
\item[b)] The values $p,q$ used in Eq.~\ref{eq:sat_loc} for
modeling the location of unresolved subhaloes. Alternatively, we use the
location as given by the most bound particle.
\item[c)] For central subhaloes, 6 values of $\alpha_i$ that
define the domain boundaries $U_i^c$ (Eq.\ref{eq:alpha_i}).
\item[d)] For satellite subhaloes, 6 values of $\beta_i$, and
6 values of $\delta_i$ that correspond to the $U_i^s$ boundaries.
\end{description}
Once the parameters $\alpha_{\rm df}$, $p$, $q$ are chosen, we
construct the subhalo statistical functions $\phi$ and $\psi$. The
parameters in items (c) and (d) above are then used as integration limits
for $\phi$ and $\psi$.
By integrating $\phi$ we compute the total number of subhaloes within the domain:
\begin{eqnarray}
\label{eq:N}
\lefteqn{
N = V\int_{U_i^c}^{U_{i+1}^c} \phi_c\, {\rm d}{\logM_{\rm infall}} \, + } \\ \nonumber & &
V\int_{U_i^s}^{U_{i+1}^s} \phi_s \, {\rm d}{\logM_{\rm infall}} \, {\rm d}{\logM_{200}} \,.
\end{eqnarray}
A similar integration of $\psi$ over each domain results in $N_p(r_p)$ -- the total
number of pairs within each radial bin $r_p$. The projected auto-correlation function,
$w_p(r_p)$, is then defined as the deviation
in the number of pairs from the average value per volume:
\begin{equation}
w_p(r_p) = \left[ \frac{L^2}{N^2} \frac{V^2\,N_p(r_p)}{A_p} - 1 \right] L \,.
\end{equation}
Here $A_p$ is the 2-dimensional area covered by the radial bin $r_p$, and $L$ is the size of
the simulation box in $h^{-1}$Mpc.
We have tested numerically that our
methodology agrees with the standard ABM approach for various different
models. In our approach the models are based on quantifying $\phi$ and $\psi$ only,
so we do not construct a full realization of galaxies per each model. This is in
contrast to standard ABM models, where a list of all subhaloes within a given simulation
is necessary.
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig4a.eps,width=9cm} }}
\centerline{ \hbox{ \epsfig{file=fig4b.eps,width=9cm} }}
\caption{Same as Figs.~\ref{fig:model_a1} \& \ref{fig:model_a2} but for model $B$. Here $\alpha_{\rm df}=3$,
and the location of unresolved subhaloes is set by the most bound particle.}
\label{fig:model_b}
\end{figure}
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig5a.eps,width=9cm} }}
\centerline{ \hbox{ \epsfig{file=fig5b.eps,width=9cm} }}
\caption{Same as Figs.~\ref{fig:model_a1} \& \ref{fig:model_a2} but for model $C$. Here $\alpha_{\rm df}=3,p=0.5,q=0.8$.}
\label{fig:model_c}
\end{figure}
\section{A few examples}
\label{sec:examples}
In this section we present a few specific models that were found using a large parameter search
within our formalism. The details of the search will be presented in the next section.
Here we discuss a few examples to demonstrate the parametrization needed for each model,
and to highlight the variety of models that are able to fit both the stellar
mass function and CFs.
In Figs.~\ref{fig:model_a1} \& \ref{fig:model_a2} we present model $A$, one example out of a family of models
which are extreme with respect to the full population of models.
In this model, $m_{\star}$ is much higher for satellite subhaloes than for central subhaloes,
for a given $M_{\rm infall}$. The difference can reach a factor of 100 in stellar mass.
As can be seen from Fig.~\ref{fig:model_a1}, this model includes a relatively strong dependence
of $m_{\star}$ on $M_{200}$, a feature that was originally seen in the SAMs analyzed in paper I, although less
prominent. We note that previous ABM models were using only
horizontal lines in the $M_{200}-M_{\rm infall}$ plane.
The CFs of model $A$ are plotted in Fig.~\ref{fig:cfs}, showing a reasonable match to the
observed data, with an RMS deviation of 0.2 dex (more details on how we define this deviation
can be found in section \ref{sec:strategy}). Interestingly, the same model fits the stellar mass
function well (see Fig.~\ref{fig:smf_examples}). These results suggest that even though model $A$ is extreme,
it is broadly consistent with the observational constraints adopted here.
However, Fig.~\ref{fig:model_a2} shows that this model violates the limit $m_{\star}<0.17M_{\rm infall}$
for satellite subhaloes. Assuming all baryons within each subhalo are converted into stars,
the maximum stellar mass should equal $m_{\star}=\Omega_B \Omega_m^{-1} M_{\rm infall}=0.17M_{\rm infall}$. Model $A$ exceed
this limit, and is therefore rejected and will not be considered
as a valid model in what follows. It is plotted here to demonstrate
that using our formalism, fitting both the stellar mass function and the CFs is not enough
for constraining the maximum stellar mass per a given $M_{\rm infall}$.
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig6a.eps,width=9cm} }}
\centerline{ \hbox{ \epsfig{file=fig6b.eps,width=9cm} }}
\caption{Same as Figs.~\ref{fig:model_a1} \& \ref{fig:model_a2} but for model $D$. Here $\alpha_{\rm df}=3$,
and the location of unresolved subhaloes is set by the most bound particle.}
\label{fig:model_d}
\end{figure}
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig7.eps,width=9cm} }}
\caption{
The stellar mass functions of galaxies using models $A,B,C,D$ (symbols).
The observed function derived by \citet{Li09} is plotted in dashed line.
Symbols are placed at the centre of each domain.
}
\label{fig:smf_examples}
\end{figure}
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig8.eps,width=9cm} }}
\caption{
The fraction of satellite galaxies out of all galaxies, $f_{\rm sat} = \phi_s/(\phi_s+\phi_c)$,
for the different models presented here.
}
\label{fig:sat_frac}
\end{figure}
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig9.eps,width=9cm} }}
\caption{The ratio between stellar mass and $M_{\rm infall}$ for models $B,C,D$.
Satellite subhaloes are plotted as dashed lines, and are using the median values for a given $m_{\star}$ as shown in
Figs.~\ref{fig:model_b}, \ref{fig:model_c}, and \ref{fig:model_d}.
Thin, medium, and thick lines correspond to models $B,C,D$ respectively. All lines are artificially smoothed
within the $m_{\star}$ domains for a better view. The peak in the stellar fraction can change by more than a factor
of 10 between different models, and between satellite and central subhaloes.}
\label{fig:mass_ratio_bcd}
\end{figure}
\begin{figure*}
\centerline{\psfig{file=fig10.eps,width=150mm,bbllx=30mm,bblly=80mm,bburx=188mm,bbury=200mm,clip=}}
\caption{The projected auto-correlation functions (CFs) derived for models $A,B,C,D$.
Each panel corresponds to galaxies with stellar masses as indicated by the range of Log$M_\odot$.
Lines show the results of the models. The observational
data are using SDSS DR7 with the same technique as in \citet{Li06}, and are shown as error bars.
Models $B,C,D$ fit the observational data to a level of 0.1 dex RMS, while model
$A$ deviates at the level of 0.2 dex (details on the way these errors are computed
can be found in section \ref{sec:strategy}).}
\label{fig:cfs}
\end{figure*}
Models $B$ and $C$ are plotted in Figs.~\ref{fig:model_b} \&
\ref{fig:model_c}, and show
a very different behaviour for low mass subhaloes. The difference for a given $m_{\star}$ between these two models
can reach a factor of 10 in $M_{\rm infall}$ for central and satellite subhaloes. Model $C$ shows a
steep dependence of $m_{\star}$ on $M_{\rm infall}$ for low mass central subhaloes. This means that for
a given subhalo mass, the difference between $m_{\star}$ for satellite and central galaxies might be
very large, more than a factor of 10.
We emphasize that our definition of a `satellite' versus `central' is valid only for subhaloes. It might
be that satellite subhaloes will host more massive galaxies than the central subhalo
within the same group. In these cases the more massive galaxy in the group
might be identified as a `central' galaxy, although its host subhalo is defined as a `satellite' here.
In a recent study, \citet{Skibba11}, have estimated the fraction of haloes that host central galaxies that
are not the most luminous in their group. Using a group catalog based on SDSS \citep{Yang07} they found that this fraction
reaches 25 (40) per cent for low (high) mass haloes. Thus, a different identification of `central' and `satellite' may be used either
in observational studies, or in the analysis of hydrodynamical simulations. In our models, the difference between satellite
and central subhaloes is defined solely according to the dark-matter behaviour, and is motivated by the
different merger history of subhaloes, and their final location.
From Fig.~\ref{fig:model_c} it is evident that model $C$ predicts a
non-negligible number of low mass subhaloes that host galaxies
with very high stellar masses, including galaxies with $m_{\star} \sim0.6\,\cdot\,0.17M_{\rm infall}$.
This shows again that our formalism cannot constrain the mass fraction that is locked in stars,
in contrast to what has been claimed using
more simplified ABM models \citep[e.g.][]{Behroozi10,Guo10a}.
Model $C$ is a good example of a model with a very high abundance of
satellite galaxies.
Although the number of satellite \emph{subhaloes} is smaller than central subhaloes at a given $M_{\rm infall}$,
different $m_{\star}-M_{\rm infall}$ relations for central and satellite subhaloes might result in
a high abundance of satellite \emph{galaxies}.
We plot the satellite fraction (the number of satellite galaxies out of all the galaxies of a given
$m_{\star}$) of all the models in Fig.~\ref{fig:sat_frac}. The satellite
fraction changes significantly between the models.
For model $C$ this fraction reaches unity at low stellar masses, while for models $B$ and $D$ it reaches
a maximum value of $\sim0.3$ and 0.4 respectively.
Model $D$ is shown in Fig.~\ref{fig:model_d}. It is an example for a model in which
$m_{\star}$ for satellites does not strongly depend on $M_{\rm infall}$.
In general, the slope of the $m_{\star}-M_{\rm infall}$ relation, and the location
where it turns over can be related to various processes in galaxy formation
physics. Merger-rates, cooling, feedback and star-formation
should all be combined in order to reproduce the observed $m_{\star}-M_{\rm infall}$
relation \citep[e.g.][]{Shankar06}. Unless our more extreme models like model $D$
are ruled out by other observational constraints,
our results indicate a large amount of freedom in modeling the above processes.
The stellar fractions, $m_{\star}/M_{\rm infall}$, for the few models presented here, are plotted
in Fig.~\ref{fig:mass_ratio_bcd}.
The halo mass at which the global stellar fraction reaches a maximum can range
from $\sim3\times10^{11}$ to $\sim3\times10^{12}\,M_\odot$. In model $D$ this efficiency
is approximately constant for satellite galaxies as a function of $M_{\rm infall}$, showing no global
strong peak.
The CFs of all the models are shown in Fig.~\ref{fig:cfs}.
Models $B,C,D$ fit the observed CFs to a good accuracy, below 0.1 dex RMS,
while model $A$ is less accurate, reaching a level of 0.2 dex RMS. This demonstrates the importance
of using this constraint here, in order to narrow down the range of accepted
models. As can be seen from
Fig.~\ref{fig:smf_examples}, all the models fit the stellar mass
function well.
Another point that should be emphasized regarding Fig.~\ref{fig:cfs}
is the small scale ($<1$Mpc) clustering of massive
galaxies. The CF of different models shows variations at this regime, so
it might be that extending the CF of the most massive domain into
smaller radii will help to constrain the models. This will require
a survey volume much larger than in the current SDSS survey.
\section{Results}
\label{sec:results}
\subsection{Search strategy}
\label{sec:strategy}
Each model within our formalism is defined by 21 free parameters,
where 18 of them define the domains in stellar masses ($U_i^c$ \& $U_i^s$ from Eq.~\ref{eq:alpha_i}) used to compute
the CF, 1 parameter fixes the time-scale for dynamical friction, and 2 parameters
are responsible for the location of unresolved subhaloes. We were able to scan
\emph{all} possible options of the 18 parameters that govern the domains behaviour
(within the resolution adopted here). In addition, we have tested five
$\alpha_{\rm df}$ values in the range 0.1--10,
and a few options for setting the location
of unresolved subhaloes (see Appendix \ref{sec:sat_loc} and Table \ref{tab:models} for more details).
In total, we have tested the stellar mass function for
$\sim10^{12}$ models, approximately $10^7$ out of them were tested
against the observed CFs.
More numerical details regarding the search algorithm can be found in
Appendix \ref{sec:scan_prm}.
The values of $\phi$ and $\psi$ are saved numerically in fine bins. The choice of bin size is important
for several reasons. First, the function $\psi_{ss}$ depends on 5 different variables, so the number of bins
we use for each variable is limited because of computer memory issues. Second, fine bins will require
more evaluations of CFs, as the code checks automatically all possible options. This can slow our code
dramatically. On the other hand, the bin size
limits the search resolution: too big bins will not allow us to find all possible solutions.
As a compromise between these requirements we have chosen the following bin sizes:
$M_{\rm infall}$ is split into bins of 0.02 (0.1) dex for subhaloes with mass smaller (bigger) than
$10^{12}\,M_\odot$, data per $r$ and $M_{200}$ are saved in bins of 0.1 dex.
We use small bins for low mass $M_{\rm infall}$ because the typical difference between adjacent $U_i$ is smaller
for small mass $M_{\rm infall}$ (see e.g., model $C$ above).
As described in section \ref{sec:def_model}, the domain boundaries
are defined using power-law relations. Writing Eq~.\ref{eq:alpha_i} in terms of Log mass gives
$\logM_{\rm infall}=\log\beta_i+\delta_i\logM_{200}$. Our search
algorithm checks all the possible values of $\beta_i$ according to the bins of $M_{\rm infall}$.
$\delta_i$ is modeled in terms of the line slope, $\delta_i=\tan \theta_i$. We sample
$\theta_i$ in steps of $6.75$ degrees, over the range $[-90,\, 45]$.
Each model is accepted if it follows the conditions below:
\begin{itemize}
\item The model fits the stellar mass function of \citet{Li09} with an accuracy that is better than 20 per cent.
This criterion is applied separately to each domain by integrating the stellar mass function over
the domain range (0.5 dex in $m_{\star}$). We have checked that the range of models
shown below changes in a very minor way, when demanding a better fit that resembles the statistical
errors from \citet{Li09} (these are 5,5,5,10,20 per cent for each domain, in order of increasing mass).
Here we choose to use a constant accuracy of 20 per cent to account for systematic uncertainties in the
stellar mass function \citep[comparisons against other measurement can be found in][]{Guo10a,Bernardi10}.
\item The model fits the logarithm of the observed CF (see Fig.~\ref{fig:cfs}) to better than an RMS value of 0.1 dex
(26 per cent). This estimate is based on all points in the range $0.03<r<30$ Mpc $h^{-1}$, sampled in bins separated by
0.2 dex in Log($r$). In order to test the effect of this fit accuracy we also show results using 0.2 dex deviation.
A detailed discussion of these issues is given below.
\item The model does not include individual points that deviate from the observed CF to more than a factor of 2.
\item The stellar mass function is fitted also
for masses larger than the most massive domain (i.e. for~$m_{\star}>10^{11.77}\,M_\odot$).
This domain should include only 23 galaxies, when using the full Millennium simulation.
Since this is a very small number, we allow
our models to deviate an additional Poisson error from the nominal value. This means that
the number of galaxies with masses larger than $10^{11.77}\,M_\odot$ can have any value
in the range [14, 34].
\item The stellar fraction (i.e. $m_{\star}/M_{\rm infall}$) does not exceed the universal fraction of
0.17 for all subhaloes in the sample.
\end{itemize}
We adopt an accuracy of 0.1 and 0.2 dex for fitting the CFs due to the following reasons.
First, the Millennium simulation being used here is based on a cosmological model with
$\sigma_8=0.9$, higher than the most updated measurements of $\sigma_8=0.8$ \citep{Jarosik11}.
This should give rise to some deviation between our models and observations. We find that
there is a minimum deviation of 0.08 dex between all the models discussed here,
and the observed CFs. Since this deviation is dominated by scales larger than 1 Mpc, it is probably
related to the different cosmological model assumed.
Second, due to the finite bins that are used for saving $\psi$, the accuracy of our search algorithm
is approximately 0.03 dex. As a result of the above, our minimum fit criterion is chosen to be 0.1 dex.
However, since the error bars presented for the observed CFs are only statistical, it might be that systematic
uncertainties would introduce errors that are larger than 0.1 dex.
The CFs depend strongly on various galaxy properties like color, star-formation, and morphology \citep{Li06}.
Consequently, if the stellar masses of galaxies
are systematically biased for galaxies of a given property (for example the high star-forming
galaxies), this could change the CFs considerably.
We therefore additionally consider a fitting criterion of 0.2 dex, to demonstrate the effect of the
possible systematic uncertainties.
The minimum subhalo mass ($\sim6.3\times10^{10}\,M_\odot$),
enforced by the Millennium simulation seems to limit our models (see models $A$ and $C$ above).
It might be that using simulations of higher resolution will permit more models.
The Millennium II simulation \citep{Boylan09} is a good candidate for such a study. However,
high resolution simulations are naturally based on much smaller volume than what is being used here,
resulting in a smaller statistical sample and non-negligible cosmic variance effects. Since our study is
aiming at fitting the CFs to a high accuracy of 0.1 dex, we focus our study only on the Millennium simulation.
\subsection{The mass relation}
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig11.eps,width=9cm} }}
\caption{A summary of all the models that fit the observed stellar mass function and the auto correlation functions
of galaxies at $\emph{z}=0$. Lines show the envelope of all relations between $m_{\star}$ and $M_{\rm infall}$ (we use \emph{median}
values of $M_{\rm infall}$ per a given $m_{\star}$ for each model). The results are separated into central and satellite subhaloes.
Thin dotted-dashed line corresponds to the median $M_{\rm infall}$ for all satellite galaxies residing in haloes more
massive than $10^{15}\,M_\odot$. The models
reproduce the observed stellar mass function to better than 20 per cent, and the CFs to better than 0.1 dex RMS. The thin solid
line shows the $m_{\star}-M_{\rm infall}$ relation using the same behaviour for satellite and central galaxies, with no dependence on $M_{200}$.}
\label{fig:models_all}
\end{figure}
The main results of the parameter search are summarized in Figs.~\ref{fig:models_all} and
\ref{fig:models_58}. In case we force the models to reproduce the observed CFs to a high
accuracy of 0.1 dex, the range of accepted models occupies a region of $\sim$1 dex in the
$m_{\star}-M_{\rm infall}$ plane, as shown in Fig.~\ref{fig:models_all}.
Interestingly, the uncertainty is small for central massive subhaloes. On the other
hand, low mass central subhaloes can host galaxies with very low stellar masses. The range of
accepted models increases significantly in Fig.~\ref{fig:models_58}, where the accuracy of fitting
the CFs is set to 0.2 dex. We have checked that the distributions of models within
the plotted envelope in Figs.~\ref{fig:models_all}-\ref{fig:models_58}
are roughly uniform, so the range of models is not affected by outliers.
The number of models presented in Fig.~\ref{fig:models_all} and \ref{fig:models_58} is quite large.
Each domain includes $\sim10^5$ successful models in Fig.~\ref{fig:models_all}, and $\sim10^6$ models in
Fig.~\ref{fig:models_58}. We only take into account domains
for which nearby $U_i$ from adjacent domains coincide, and the set of
all 5 domains covers the full mass range. Due to the above, we compute the median $M_{\rm infall}$
for a given $m_{\star}$, which does not demand combinations of domains of different $m_{\star}$.
Computing the opposite relation, median $m_{\star}$ per a given $M_{\rm infall}$, is much more complicated
within our formalism. However, since the relation for central galaxies
does not include any scatter, it represents a median $m_{\star}$ per a given $M_{\rm infall}$ as well.
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig12.eps,width=9cm} }}
\caption{Similar to Fig.~\ref{fig:models_all} but allowing the models to deviate from the
observed CFs by up to 0.2 dex RMS.
Results from weak lensing analysis by \citet{Mandelbaum06} are plotted in symbols. Squares and circles refer to the
\emph{mean} value of early and late type galaxies respectively, with error bars that reflect 95 per cent confidence level.
The fraction of late-type galaxies out of the full sample is 0.74, 0.60, 0.46, 0.32, 0.20, 0.11, 0.05
(ordered in increasing $m_{\star}$). }
\label{fig:models_58}
\end{figure}
In our models, the stellar mass of satellite galaxies depends on their host
halo mass, which correlates with the
number of neighboring galaxies in observational studies. As a result, the range in the $m_{\star}-M_{\rm infall}$ relation changes
as a function of the host halo mass. This is shown in Fig.~\ref{fig:models_all} by plotting the envelope of all median
relation in the $m_{\star}-M_{\rm infall}$ plane, but taking into account only galaxies that reside in clusters (their host halo mass
is bigger than $10^{15}\,M_\odot$). For these satellite galaxies, the range in $M_{\rm infall}$ for a given $m_{\star}$ is larger
than the one for all galaxies.
Fig.~\ref{fig:models_58} includes the results of weak lensing analysis
from \citet{Mandelbaum06}.
These estimates are for the host halo mass\footnote{\citet{Mandelbaum06} define the halo mass as the mass
enclosed within a radius that corresponds to 180 times the mean density, this mass is higher than $M_{200}$ \citep[see also][]{Weinmann06a},
but agrees quite well ($<0.1$ dex) with $M_{\rm infall}$ for central subhaloes used here.} of central galaxies, and should be compared to the solid lines
of our models. The square symbols represent a subsample of early type galaxies, while the circles correspond
to late-type galaxies. As is mentioned in the figure caption, the fraction of late-type galaxies is
very small at the most massive stellar mass bin (0.05).
Surprisingly, our results deviate from those of \citet{Mandelbaum06} at the high mass end, even when
the range of models is large, using an uncertainty of 0.2 dex in matching the CF.
This is similar to what was found by previous ABM studies, summarized by
\citet{Behroozi10}. Apparently, for a given value of $m_{\star}$,
the weak lensing results constrain the host halo mass by providing mainly an upper
threshold (the low values of the host halo often reach the universal baryonic fraction).
The opposite seems to be true for central galaxies in our analysis.
Deviations between the two studies can be due to various effects:
\begin{itemize}
\item \citet{Mandelbaum06} use a specific halo model for computing the lensing
signal, which differs from the set of models being used here.
\item There is some uncertainty
in their study owing to the width of the $M_{\rm infall}$-distribution, for a given $m_{\star}$.
\item Their stellar mass estimates are based on \citet{Kauffmann03} while the observations
of \citet{Li09} use the method of \citet{Blanton07}.
The difference between these estimates is discussed in both \citet{Li09} and \citet{Guo10a}, and its effect
on the stellar mass function is probably small. However, it might be that the difference between the methods
are more significant for computing the CFs.
\item It might be that our range of models is too narrow, or
that there are some systematic differences between the two approaches.
This could be due to our search resolution, the assumed
underlying cosmological models, or the assumed IMF.
\end{itemize}
A better, more self-consistent way to compare our results against weak lensing would demand
a direct estimate of the lensing signal from our models. This can then be compared to
the observed shear signal. Such an analysis is however outside the scope of this work.
A recent study by \citet{Leauthaud11} have used a model based on halo occupation distribution
and demonstrated the strength of applying this additional constraint.
The results of \citet{Mandelbaum06} indicate that the population of late-type massive central galaxies live inside
haloes with lower masses than early-type galaxies. The difference between early and late
type galaxies indicates that the scatter in the $m_{\star}-M_{\rm infall}$ relation might be significant.
A large scatter was also found by \citet{More11}, based on satellite kinematics. We do not model such a scatter
for central galaxies in this study. Previous works \citep[e.g.][]{Moster10} have shown that including a scatter modifies
the $m_{\star}-M_{\rm infall}$ relation at the massive end, such that for a given $M_{\rm infall}$, $m_{\star}$ decrease with increasing scatter.
Interestingly, the results of \citet{Mandelbaum06} reach the universal
baryonic fraction, $m_{\star}/M_{\rm infall}\sim0.17$, for part of the galaxy population. This is very similar to what
is found here, and can be seen in model $C$ (Fig.~\ref{fig:model_c}).
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig13.eps,width=9cm} }}
\caption{The effect of different parameters on the median relation between $m_{\star}$
and $M_{\rm infall}$. Each panel shows a selection of models out of the full sample, as
plotted in Fig.~\ref{fig:models_all}.
The \emph{upper left} panel takes into account all models for which the behaviour
for satellite and central subhaloes is the same ($\pm0.2$ dex), and no dependence on $M_{200}$ is allowed.
In the \emph{middle left} panel we plot models with a similar median $M_{\rm infall}$ for
central and satellite galaxies, at a given $m_{\star}$. The \emph{lower left} panel plots models
for which there is no dependence on $M_{200}$.
Models using the default value of $\alpha_{\rm df}=3$ and location set by the most bound particle
are plotted in the \emph{upper right} panel. The \emph{middle right} panel summarizes models
that use the location as given by the most bound particle.
The \emph{lower right} panel summarizes all models with the same dynamical
friction constant, $\alpha_{\rm df}=3$.}
\label{fig:models_cuts}
\end{figure}
In Fig.~\ref{fig:models_cuts} we plot various sub-samples of the models from Fig.~\ref{fig:models_all}, showing the influence
of different parameters on the relation between $m_{\star}$ and $M_{\rm infall}$. Our results agree with previous studies in predicting
a very tight $m_{\star}-M_{\rm infall}$ relation, once satellite and central galaxies are considered to have very similar
median relations (at the level of $\sim0.2$ dex), and $M_{200}$ does
not affect $m_{\star}$. This tight relation holds even though the models include the new ingredients related to $\alpha_{\rm df}$ and
the location of satellite subhaloes. This proves that
the dependence of $m_{\star}$ on $M_{200}$, and the difference between central and satellite galaxies,
are responsible for the range of allowed models shown in Figs.~\ref{fig:models_all}-\ref{fig:models_58}.
It is worth noting that our models do not include solutions that have \emph{exactly} the same
mass relation for central and satellite galaxies, and can still fit the CFs to a level of 0.1 dex. The minimum deviation
between central and satellite galaxies in the upper left panel of Fig.~\ref{fig:models_cuts} is roughly 0.12 dex in $M_{\rm infall}$.
This small difference might be an artifact of the
underlying cosmological parameters assumed by the Millennium simulation.
The models shown here use the two populations of galaxies to compensate for each other, keeping the overall $m_{\star}-M_{\rm infall}$ relation
relatively similar. This effect is shown in the middle left panel of Fig.~\ref{fig:models_cuts}. Once we force the models
to have a similar median $M_{\rm infall}$ for satellite and central galaxies, the range of models decreases, especially at the high
mass end. It might be that constraints on the differences between $m_{\star}$ of the two populations, for a given $M_{\rm infall}$, could be
enforced ad-hoc. However, it is not clear what this limit should be. A reasonable demand would be that the full distribution
of the two populations would overlap to some level, allowing a transition of central galaxies into satellites. This condition
seems to be fulfilled by our models (see Figs.~\ref{fig:model_b}-\ref{fig:model_d}). Since we do not model a distribution
for the $m_{\star}$ of central galaxies, we do not explore this issue further here.
The effect of $\alpha_{\rm df}$ and satellite locations is explored in Fig.~\ref{fig:models_cuts}. It seems that both elements add freedom
to our models, although the effect of satellite location is slightly stronger. Higher resolution cosmological simulations
might help to constrain $\alpha_{\rm df},p,q$. However, these simulations are based on dark-matter only, while the results here
are affected strongly by the baryonic components of galaxies \citep[see e.g.][]{Boylan08,Jiang08}.
\begin{figure}
\centerline{ \hbox{ \epsfig{file=fig14.eps,width=9cm} }}
\caption{Similar to Fig.~\ref{fig:models_all}, but here the CFs are constrained to fit
one of our models, instead of the observational
data. We use as a reference the simplest model where satellite and central
galaxies have the same relation between $m_{\star}$ and $M_{\rm infall}$, and $\alpha_{\rm df}=3$.
This shows that the uncertainty in the median relation between $m_{\star}$ and $M_{\rm infall}$
does not depend strongly on the reference model, and its underlying cosmological parameters.}
\label{fig:models_ms}
\end{figure}
Lastly, in Fig.~\ref{fig:models_ms}, we show the results of a
parameter search when the reference CFs are not the observed ones,
but instead are taken from one model out of the models tested here. The range of accepted models
in this case is related to the internal degrees of freedom of the models, and is less related to the choice
of the cosmological parameters, the method adopted for computing $m_{\star}$, or to the specific physical assumptions used in our formalism.
The fact that Fig.~\ref{fig:models_ms} is similar to Fig.~\ref{fig:models_all}
indicates that the underlying cosmological model does not affect the
range of acceptable models significantly.
This is because here the reference CFs are based on the same cosmology, but use a specific
model for the relation between $m_{\star}$ and subhaloes.
As was pointed out by \citet{Cacciato09}, the effect of the underlying cosmological model should
be important when constraints from abundance and clustering of galaxies are combined with
galaxy-galaxy lensing measurements. This issue should be further examined using
an appropriate $N$-body cosmological simulation.
\section{Summary and discussion}
\label{sec:discuss}
In this work we have studied the relation between the stellar mass of
galaxies ($m_{\star}$) and the mass of their host subhaloes ($M_{\rm infall}$). Our models are
constrained by the abundance of galaxies, and their auto-correlation function (CF) at
$\emph{z}=0$. We have shown that
once the population of galaxies is broken into
two sub-populations of central and satellite galaxies,
the allowed range in the $m_{\star}-M_{\rm infall}$ relation
for each population reaches a factor of $\sim10$. The range of accepted models
depends on the accuracy by which the models reproduce
the observed abundance and clustering of galaxies.
As was demonstrated in our previous work
\citep[][paper I]{Neistein11}, a different $m_{\star}-M_{\rm infall}$ relation for central and
satellite galaxies is expected in galaxy formation models, although
the strength of the deviation cannot be constrained easily.
The shape of the $m_{\star}-M_{\rm infall}$ relation has a large degree of freedom, resulting in a
stellar fraction (the ratio between
$m_{\star}$ and $M_{\rm infall}$) that can peak anywhere between $\sim3\times10^{11}$ and $3\times10^{12}\,M_\odot$. It
can even be close to
a constant as a function of $M_{\rm infall}$ for satellite
galaxies. Interestingly, the stellar fraction can reach high
values, comparable to the maximum universal value of 0.17, for low mass satellite galaxies
residing in low mass haloes.
Although it seems unlikely that the stellar fraction would exceed 0.17, our models
are not able to constrain this ratio, and instead we are forced to use it as an ad-hoc constraint.
Our models are similar to the `abundance matching' (ABM) approach, and are directly relating a
galaxy with a specific $m_{\star}$ to each subhalo within a large $N$-body simulation.
However, in comparison to previous studies, we include various additional ingredients for modeling
satellite galaxies. Their stellar mass
depends on both the host subhalo mass ($M_{\rm infall}$) and on the halo
mass at $\emph{z}=0$, and might differ from the stellar
mass of central galaxies of the same $M_{\rm infall}$. We include subhaloes
that have merged into more massive subhaloes at high-$\emph{z}$ but might host
galaxies that will survive until $\emph{z}=0$, due to long dynamical friction time-scales.
The location of these `unresolved' subhaloes is set by either the
location of their most bound particle, or by an analytical model.
We found that the most important ingredient for fixing the range of accepted models
is the way stellar masses for satellite galaxies are modeled.
A model with two populations of galaxies includes a large amount of freedom in
matching the observed stellar mass function. We therefore constrain
our models also against the observed auto-correlation function (CF) of
galaxies. We developed a new formalism to compute the model CF based
on the pair statistics of subhaloes. This enabled us to scan
systematically a significant part of the parameter space. We have
tested $\sim10^{12}$ different ABM models, and estimated the CF for
$\sim10{^7}$ models that showed a good match to the stellar mass
function. The accuracy by which we fit the CF affects
significantly the range of accepted models within the $m_{\star}-M_{\rm infall}$ plane.
Consequently, a detailed study of the systematic uncertainties involved in
this measurement is crucial in order to better constrain our models.
The range allowed in the $m_{\star}-M_{\rm infall}$ relation might be larger than our
prediction due to various effects. There are still a large number of
models that we did not test here. These include complicated
dependence of $m_{\star}$ on the halo mass, different
estimates for dynamical friction and satellite locations,
higher resolution of the underlying $N$-body simulation, different
observed stellar mass functions \citep[e.g.][]{Bernardi10},
and modeling a random scatter in $m_{\star}$. On the other hand, the range of
models might be modified once we use subhaloes from an $N$-body simulation using
the most up-to-date cosmological parameters.
Previous estimates for the uncertainty in the $m_{\star}-M_{\rm infall}$ relation were emphasizing
the contribution from uncertainty in
the stellar mass function, resulting in 0.25 dex
\citep{Behroozi10}. This uncertainty should be added to what we find here.
In this paper we have reproduced the abundance
and clustering of galaxies only at $\emph{z}=0$ when constraining the $m_{\star}-M_{\rm infall}$ relation. It is not likely that data from higher redshifts
would limit the $m_{\star}-M_{\rm infall}$ relation to a narrow range.
This is because the same uncertainties discussed here would be valid at high-$\emph{z}$, in
addition to the larger observational errors inherent at these redshifts.
Moreover, once the $m_{\star}-M_{\rm infall}$ relation is fixed at some redshift range,
it is not straightforward to decide which models violate the physics of galaxy formation
by linking galaxies at different epochs. For example, as we showed
in paper I, the $m_{\star}-M_{\rm infall}$ relation for satellite galaxies at the time
of infall might already be different from the relation for central
galaxies at the same epoch.
Our empirical results have implications for various aspects of galaxy formation.
\citet{Guo10a} and \citet{Sawala11} argued that stellar fractions at a fixed subhalo
mass derived from basic ABM models are systematically lower than the
results of detailed hydro-simulations. Our results that allow for a much larger spread
of stellar mass at a fixed subhalo mass, are less definitive in this respect. This
is especially true for dwarf galaxies, for which there is no available clustering
data. An additional implication is related to
the cumulative energy injected into the interstellar
medium in a supernova feedback-constrained scenario. Shankar et al. \citep[2006; see also][]{Dekel03}
re-derived the expected trend of stellar and halo
mass of the type $m_{\star} \propto f_{\rm surv} M_{\rm infall}^{\alpha}$. Here
$1 < \alpha < 2$ and $f_{\rm surv}$ is the fraction of surviving stellar mass.
Based on the observed rather low stellar mass in haloes with $M_{\rm infall} \lesssim 10^{11}\, M_\odot$
inferred from a basic ABM model, they argued that supernova
feedback appeared to be insufficient to remove the gas associated with the host halo.
The results of this paper, however, show that some ABM models allow for much
larger stellar mass fraction in low mass haloes,
thus providing some hints towards the solution of this puzzling issue.
Another possible non-trivial consequence of our results concerns Halo Occupation
Distribution (HOD) models. As outlined in e.g., \citet{Berlind02}, this class
of models is based on a parameterized conditional probability $P(N|M)$ that a dark
matter halo of virial mass $M$ contains $N$ galaxies. Adding prescriptions on the spatial
distribution of subhaloes derived from accurate $N$-body simulations, HOD models have
been rather successful in reproducing the two-point correlation
function of different classes of galaxies and at different redshifts.
However, once we allow for a different ranking between satellite and central
galaxies, as was found in this paper, the shape of $P(N|M)$ might
become rather different from its usual functional form. This could induce
non-trivial degeneracies in these models,
allowing for different occupation
distributions that still match the data.
The methodology developed in this work can be used to study the clustering properties
of AGNs, cold gas, and subsets of galaxies divided by e.g. luminosity
or color. As was demonstrated here, our method already enabled us to better fit the
observed stellar mass functions and CFs of all galaxies, with
the current data sets. It might become more important when studying clustering properties of objects
which tend to depend more strongly on environment, like the cold gas mass within galaxies.
Our method may be able to shed some light on the limitations in
modeling the observed high-$\emph{z}$ galaxies using standard halo models \citep{Quadri08,Tinker10,Wake11}.
In addition, it might be useful for future surveys with high quality data and a large survey volume,
for which the simple abundance
matching approach might not be flexible enough to provide an accurate fit to the data.
Improved constraints on the $m_{\star}-M_{\rm infall}$ relation might be obtained from observational
quantities, other than the stellar mass function and clustering used here.
These might include: conditional stellar mass functions per
halo mass \citep[e.g.][]{Yang09},
satellite kinematics \citep{More11}, weak lensing \citep{Mandelbaum06,Cacciato09}, star-formation histories,
and dynamical tracers such as the velocity dispersion and circular
velocity \citep[e.g.][]{Dutton10}. In addition, physically
motivated models like hydrodynamical simulations or semi-analytic models
are crucial to obtain insights on the difference between the evolution of
satellite and central galaxies. More effort to understand the fundamental relation
between the mass of subhaloes and galaxies is clearly needed both on the observational and theoretical side.
\section*{Acknowledgments}
We thank the referee for a detailed and constructive report that helped to improve the
presentation of this work.
We also thank Peter Behroozi, Charlie Conroy, Umberto Maio, Rachel Mandelbaum,
Risa Wechsler, and Simon White for useful discussions.
The Millennium Simulation databases used in this paper and the web application providing online access to
them were constructed as part of the activities of the German Astrophysical Virtual Observatory.
FS acknowledges support from the Alexander von Humboldt Foundation.
MBK acknowledges support from the Southern California Center for Galaxy Evolution,
a multi-campus research program funded by the University of California office of
research.
\bibliographystyle{mn2e}
|
\section{Introduction}
It is a great honour for me to give a plenary talk in the Chandra Centenary
Symposium. As a graduate student of Gene Parker in the early 1980s, I had
the privilege of working for four years in an office about 4 or 5 doors down
the corridor from Chandra's office. Those of you who had visited University
of Chicago in those days may know that most of the astronomy
faculty and students
were in a building called Astronomy and Astrophysics Center. The building
next to it---Laboratory for Astrophysics and Space Research---mainly housed
the large cosmic ray research group. However,
two of the most beautiful offices in that
building were given to two theoretical astrophysicists---Chandra
and Parker. Chandra stopped taking students
after a heart attack in the 1970s and there were no students working with him
when I was in Chicago. For a while I was the only theory student having a very
nice office in that building. When I was attending the first AAS meeting of
my life, somebody asked me at the dinner
table, ``How big is your theory group?'' I replied:
``It is a very small theory group with only three members.'' The next question was,
``Who are the members?'' I casually said: ``Oh, besides myself, the other two
members of our small theory group are Subrahmanyan Chandrasekhar and Eugene Parker.''
There was such an aura around Chandra that, like most other graduate students,
I was in an awe and always tried my best to keep away from him. Since I was
the only other Indian in the building, Chandra seemed somewhat curious about
me and often asked Gene what I was doing. Gene used to tell me that I should overcome
my fear of Chandra and should talk to him some time. A few days after my first
paper dealing with the solar dynamo problem appeared in ApJ [1],
while walking along the corridor, I saw Chandra coming from the opposite direction.
Normally we would walk past each other as if we were strangers. That day, to my
utter consternation, Chandra suddenly stopped when he came close to me and looked
straight into my eyes. Then he said: ``I have seen your paper. It is a nice
piece of work.'' Without giving me any time to recover from my dazed state or to
respond, Chandra immediately started walking away. I should mention that I had
always been a great admirer of Chandra's style of writing. Although that first
paper of mine presented a relatively unimportant calculation, it was deliberately
written in imitation of the Chandra style. You may want to compare that paper
[1] with the famous S.\ Candlestickmaker paper! I have a hope that,
if Chandra were present here today, he would have taken some interest in the
subject of my presentation.
All of you know about the 11-year periodicity of the sunspot cycle. There
also seems to be a 50-year periodicity in this field which you may not be
aware of! So let me begin by telling you about this 50-year periodicity.
It appears that major breakthroughs in this field take place approximately
at the intervals of 50 years. (1) A little more than 150 years ago, the
German amateur astronomer Schwabe [2] reported the first discovery of
the sunspot cycle. (2) About 100 years ago, Hale [3] found the evidence
of Zeeman splitting in the spectra of sunspots, thereby concluding that
sunspots are regions of concentrated magnetic field. It may be mentioned
that this was a momentous discovery in the history of physics because this
was the first time somebody found a conclusive evidence of large-scale
magnetic fields outside the Earth's environment. Now we know that magnetic
fields are ubiquitous in the astronomical universe. With Hale's discovery, it
also became clear that the
sunspot cycle is essentially a magnetic cycle of the Sun.
(3) About 50 years ago, Parker [4] finally formulated the turbulent dynamo
theory, which still provides the starting point of our understanding how
magnetic fields arise in astronomical systems.
Even without a theoretical model of this 50-year periodicity, you should
be able to make a simple extrapolation and predict that another major
breakthrough in this field should be taking place right now. We are going
to argue that such a breakthrough is indeed happening at the present time.
Some of the earlier breakthroughs were achieved single-handedly by extraordinary
individuals like Hale and Parker. Now we probably live in a less heroic
age. The present breakthrough is a result of efforts due to many groups
around the world, in which our group in Bangalore also has made some
contributions.
\section{Some observational considerations}
\begin{figure}[t]
\centerline{\includegraphics[height=13cm,angle=+90]{fig1.ps}}
\caption{The yearly averaged number of sunspots plotted against time
for the period 1610--2000.}
\end{figure}
Let us begin by looking at Fig.~1, which plots the sunspot
number as a function of time from 1610. Galileo and some of
his contemporaries were the first scientists to study sunspots
systematically. The initial entries in Fig.~1 are based on
their records. Then, for nearly a century, sunspots were
rarely seen---a period known as the {\it Maunder minimum}. Afterwards
the sunspot number has varied periodically with a rough period
of about 11 years, although we see a considerable amount of irregularity.
Some cycles are stronger than the average and some are weaker.
An intriguing question is whether we can predict the strength
of a cycle in advance. Simple methods like expanding the last
few cycles in a Fourier series and continuing the series to predict
the next cycles have failed completely in the past. It is clearly
not a problem of merely extending a mathematical series and
we presumably need a proper understanding of what causes the
irregularities of the cycles if we hope to predict a future
cycle successfully. When we discuss the causes of irregularities
in sunspot cycles in \S5, we shall address the question whether
our understanding of sunspot cycles
at the present time is good enough to make
such predictions.
A few years after Schwabe's discovery of the sunspot cycle [2],
Carrington [5] noted that sunspots seemed to appear at lower
and lower latitudes with the progress of the solar cycle.
It may be mentioned that
individual sunspots live from a few days to a few
weeks. Most of the sunspots in the early phase of a solar
cycle are seen between $30^{\circ}$ and
$40^{\circ}$. As the cycle advances, new
sunspots are found at increasingly lower latitudes. Then a
fresh cycle begins with sunspots appearing again at high
latitudes. Maunder [6] made the first graphical representation
of this. In a time-latitude plot, the
latitudes where sunspots were seen at
a particular time can be marked by black bars. Fig.~2 shows one
such plot. The explanation of the grey-scale background will
be provided later. The sunspot
distribution in a time-latitude plot is often referred to as a
{\it butterfly diagram}, since the pattern (the regions marked in black
bars in Fig.~2) reminds one of butterflies.
\begin{figure}[t]
\centering
\centerline{\includegraphics[height=6cm]{fig2.eps}}
\caption{A `butterfly diagram' of sunspots,
with shades of grey showing the latitude-time distribution
of longitudinally averaged weak, diffuse magnetic field
($B$ is in Gauss).}
\end{figure}
We have mentioned Hale's discovery of magnetic fields in
sunspots [3]. A large sunspot has a typical magnetic
field of about 3000 G.
A few years later, Hale and his coworkers
made another significant discovery [7].
Often two large sunspots are seen side by side. Hale {\em et al}
[7] found that they invariably
have opposite polarities. Fig.~3 shows a magnetogram map
of the Sun in which white and black indicate respectively
regions of strong positive and negative polarities, grey
being put in regions where the magnetic field is below a
threshold. A bipolar sunspot pair appears as a white patch and
a black patch side by side. You may note in Fig.~3 that the
right sunspots in the sunspot pairs in the northern hemisphere
are positive, whereas the right sunspots in the sunspot pairs
in the southern hemisphere are negative. This is the case for
a particular cycle. In the next cycle, the polarity reverses.
The right sunspots in the northern hemisphere would become negative
in the next cycle and the right sunspots in the southern
hemisphere would become positive. If we only look at the
sunspot number, we may think that the sunspot cycle has a
period of 11 years. However, on taking account of the configuration
of the magnetic field, we realize that the Sun's magnetic cycle
has actually a period of 22 years.
\begin{figure}[t]
\centerline{\includegraphics[height=8cm,width=8cm]{fig3.ps}}
\caption{A magnetogram image of the full solar disk. The
regions with positive and negative magnetic polarities are
respectively shown in white and black, with grey indicating
regions where the magnetic field is weak.}
\end{figure}
You may note another thing in Fig.~3. The line joining the centres of
a bipolar sunspot pair is, on an average, nearly parallel to the solar
equator. Hale's coworker Joy, however, noted that there is a
systematic tilt of this line with respect to the equator
(the right sunspot in a pair appearing closer to the equator) and
that this tilt increases with latitude
[7]. This result is
usually known as {\it Joy's law}. The tilts, however, show
a considerable amount of scatter around the mean given by Joy's
law. As we shall see later, this law of tilts of sunspot
pairs plays a very important role in solar dynamo theory.
We shall present a detailed discussion in \S3 how the bipolar
sunspot pairs arise. For the time being, let us just mention
that there has to be a strand of sub-surface magnetic field
which occasionally breaks out of the solar surface as shown
in Fig.~7b. Then magnetic field lines would come out of one
sunspot (making its polarity positive) and would go down into
the other sunspot (making its polarity negative). A look at
Fig.~3 suggests that there must be a sub-surface magnetic field
with field lines going from the right to the left in the northern
hemisphere and there must be an oppositely directed magnetic field
in the southern hemisphere. Such a magnetic field in the azimuthal
direction is called a toroidal field. This seems to be the
dominant component of the magnetic field in the Sun. In contrast,
the magnetic field of the Earth seems to be of poloidal nature.
In his seminal paper on the turbulent dynamo, Parker [4]
proposed that the sunspot cycle is produced by an oscillation between toroidal and
poloidal components of the Sun's magnetic
field, just as we see an oscillation between kinetic and potential energies
in a simple harmonic oscillator. This was a truly extraordinary suggestion because almost
nothing was known about the Sun's poloidal field at that time. Babcock
and Babcock
[8] were the first to detect the weak poloidal field having a strength of about 10 G
near the Sun's poles. Over the last few years, there is increasing evidence that the
field outside sunspots is actually not weak and diffuse,
but concentrated in intermittent flux
concentrations [9]. Only in low-resolution magnetograms in which these flux
concentrations are not resolved, the field appears weak and diffuse. This seems
to be the case even in the polar regions [10]. However, we
shall not get into a more detailed discussion of this point here.
Only from mid-1970s, we have systematic data of the Sun's polar
fields. Fig.~4 shows the polar fields of the Sun plotted as a function of time
(N and S indicating north and south poles), with the
sunspot number plotted below. It is clear that the sunspot number, which is a proxy
of the toroidal field, is maximum at a time when the polar field is nearly zero. On the
other hand, the polar field is strongest when the sunspot number is nearly zero. This
clearly shows an oscillation between the toroidal and poloidal components, as envisaged
by Parker [4]. The theoretical reason behind this oscillation
will be discussed in \S4.
\begin{figure}
\center
\includegraphics[width=8cm]{fig4.eps}
\caption{The polar field of the Sun as a function of time
(on the basis of the Wilcox Solar Observatory data)
with the sunspot number shown below. The top panel, with
N and S representing north and south poles, is adapted
from [51]. The yearly modulations in the measurements
of the polar fields is due to the fact that the Sun's polar
axis is slightly inclined to the orbital plane of the
Earth's revolution around the Sun.}
\end{figure}
Let us make some more remarks on the poloidal field of the
Sun. It was found that there were large unipolar patches of
diffuse magnetic field on
the solar surface which migrated poleward
[11]. Even when
averaged over longitude, one finds predominantly one polarity
in a belt of latitude which drifts poleward
[12]. The reversal
of polar field, which occurs at the time of the sunspot
maximum [13], presumably takes place when sufficient field of
opposite polarity has been brought to the poles.
Along with the butterfly diagram of sunspots,
Fig.~2 also shows the distribution
of the longitude-averaged poloidal field in a time-latitude plot.
The various shades of grey
indicate values of the longitude-averaged
poloidal field. While the sunspots appear at lower and lower
latitudes with the progress of the solar cycle, the poloidal
field migrates poleward. The reason behind the poleward
migration of the poloidal field is a meridional circulation
in the Sun which involves a flow of gas at the surface from
the equatorial region to the polar region, having an amplitude
of about 20 m s$^{-1}$ [14].
The poloidal field is carried poleward by this meridional circulation.
It may be noted that all the observations discussed above pertain
to magnetic fields at the Sun's surface.
We have no direct information about
magnetic fields underneath the Sun's surface. In dynamo theory,
we need to study the interactions between the magnetic fields
and velocity fields. So let us now look at the nature of
the velocity fields of the Sun.
Stellar structure models
suggest that the energy produced by nuclear reactions at the
centre of the Sun is transported outward by radiative transfer
to a radius of about $0.7 R_{\odot}$ (where $R_{\odot}$
is the solar radius). However,
from about $0.7 R_{\odot}$ to $R_{\odot}$, energy is transported
by convection. This region is called the {\em convection zone},
within which the plasma is in a turbulent state with hot
gas going up and cold gas coming down. The turbulent diffusivity
of the convection zone is the main source of diffusion in the
dynamo problem. We have already pointed out that there is a
meridional circulation which is poleward near the solar surface
at the top of the convection zone. This meridional circulation
is supposed to be driven by the turbulent stresses in the convection
zone, though our theoretical understanding of this subject is
rather limited at the present time. It is generally believed
that the meridional circulation at the bottom of the convection
zone has to go from the polar region to the equatorial region
in order to conserve mass, although we do not have a direct
evidence for it yet. This meridional circulation plays a tremendously
important role in current dynamo models, as we shall see later.
\begin{figure}
\centering
\includegraphics[width=7cm]{fig5.eps}
\caption{The contours of constant angular velocity inside
the Sun, as obtained by helioseismology. The contours are
marked with rotation frequency in nHz. It may be noted that
frequencies of 340 nHz and 450 nHz correspond respectively
to rotation periods of 34.0 days and 25.7 days. Courtesy:
J. Christensen-Dalsgaard and M. J. Thomson.}
\end{figure}
We finally come to what is probably the most important part of
the Sun's velocity field for us---the differential rotation. Unlike
the Earth which rotates like a solid body, the Sun has the angular
velocity varying over it. It has been known for a long time that
the angular velocity near the Sun's equator is faster than that
at the Sun's polar regions. In the early years of dynamo research,
theorists used to make various assumptions about the distribution
of angular velocity in the Sun's interior. An amazing development
of the last few decades has been helioseismology---the study of
the oscillations of the Sun. These oscillations have allowed us
to probe various properties of the solar interior. One of the
most extraordinary outcomes of helioseismology is that solar physicists
have been able to construct a map of angular velocity distribution
in the interior of the Sun (see, for example,
[15]). A version of this map is shown in
Fig.~5. It is seen that there is strong differential rotation
(i.e.\ a strong gradient of angular velocity) at the bottom of
the solar convection zone. This relatively thin layer of concentrated
differential rotation is called the {\em tachocline}.
We have now come to an end of our discussion of what we know
about the magnetic and the velocity fields of the Sun. The aim of
solar dynamo theory is the following. Given our knowledge of
the velocity fields of the Sun, we need to study the interactions
between the velocity and magnetic fields in the Sun's interior
such that all the surface
observations of magnetic fields are properly explained---a fairly
daunting problem, of which the full solution is still a distant
dream.
\def{\bf v}{{\bf v}}
\def{\bf B}{{\bf B}}
\def\overline{\overline}
\def\cal{E}{\cal{E}}
\def\partial{\partial}
\def{\bf v}{{\bf v}}
\def{\bf B}{{\bf B}}
\section{Formation of sunspots}
All our theoretical considerations are based on magnetohydrodynamics
or MHD. An introduction to its basic concepts can be found in
Chs.~14--16 of [16] or Ch.~8 of [17]. Let
us begin by mentioning some concepts of MHD which we shall be
using repeatedly. We know that a magnetic field has a pressure
$B^2 /2 \mu$ associated with it, along with a tension along the
field lines. The other result which is going to be of central
importance to us is the theorem due to Alfv\'en [18] that, when
the magnetic Reynolds number is sufficiently high, magnetic fields
are frozen in the plasma and get carried by the velocity fields
of the plasma. Because of the high magnetic Reynolds number in
the Sun, we expect this theorem to hold---at least approximately.
Since energy is transported by convection in the layers below
the Sun's surface, sunspots are basically regions of concentrated
magnetic field sitting in a convecting fluid. To understand why
the magnetic field remains concentrated in structures like
sunspots instead of spreading out more evenly, we need to study
the interaction of the magnetic field with the convection in the
plasma. This subject is known as {\it magnetoconvection}.
The linear theory of convection in the presence of a
vertical magnetic field was studied by Chandrasekhar [19].
The nonlinear evolution of the system, however, can only be
found from numerical simulations pioneered by Weiss [20].
Since the tension of magnetic field lines opposes convection, it
was found that space gets separated into two kinds of regions.
In certain regions, magnetic field is excluded and vigorous
convection takes place. In other regions, magnetic field gets
concentrated, and the tension of magnetic field lines suppresses
convection in those regions. Sunspots are presumably such regions
where magnetic field is piled up by surrounding convection. Since
heat transport is inhibited there due to the suppression of convection,
sunspots look darker than the surrounding regions.
Although we have no direct information about the state of the
magnetic field under the Sun's surface, it is expected that the
interactions with convection would keep the magnetic field
concentrated in bundles of field lines throughout the solar
convection zone. Such a concentrated bundle of magnetic field
lines is called a {\it flux tube}.
\begin{figure}[t]
\centerline{\includegraphics[height=5cm,width=10cm]{fig6.eps}}
\caption{The production of a strong toroidal magnetic field
underneath the Sun's surface. {\bf a.} An initial poloidal field
line. {\bf b.} A sketch of the field line after it has been stretched
by the faster rotation near the equatorial region.}
\end{figure}
One important consequence of Alfv\'en's theorem of flux freezing
for the Sun is the following. If there is any poloidal field line
going through the Sun, differential rotation will drag it out
to produce a toroidal field, as shown in Fig.~6. The production
of the toroidal field is expected to be strongest in the tachocline
at the bottom of the convection zone where the gradient of angular
velocity is concentrated,
as seen in Fig.~5. Due to interactions with convection
there, the toroidal field should exist in the form of horizontal
flux tubes. If a part of such a flux tube rises up and
pierces the solar surface as shown in Fig.~7b, then we expect
to have two sunspots with opposite polarities at the same
latitude. But how can a configuration like Fig.~7b arise?
The answer to this question was provided by Parker [21] through
his idea of magnetic buoyancy. We need to have a pressure
balance across the surface of a flux tube. Since the magnetic
field inside the flux tube has a pressure
$B^2/2 \mu$, the interior pressure is a sum of this pressure
and the gas pressure $p_{\rm in}$. On the other hand, the
only pressure outside is the gas pressure $p_{\rm out}$. Hence
we must have
$$p_{\rm out} = p_{\rm in} + \frac{B^2}{2 \mu} \eqno(1)$$
to maintain pressure balance across the surface of a flux tube. It
follows that
$$p_{\rm in} \leq p_{\rm out}, \eqno(2) $$
which often, though not always, implies that the density inside
the flux tube is less than the surrounding density. If this happens
in a part of the flux tube, then that part becomes buoyant and
rises against the gravitational field to produce the configuration
of Fig.~7b starting from Fig.~7a. It is seen in Fig.~6b
that the toroidal fields in the two hemispheres are in the opposite
directions. If parts of these toroidal fields rise in the two
hemispheres to produce the bipolar sunspot pairs, we have a natural
explanation why the sunspot pairs should have the opposite polarity
in the two hemispheres as seen in the magnetogram map of
Fig.~3.
\begin{figure}[t]
\centerline{\includegraphics[height=5cm,width=10cm]{fig7.eps}}
\caption{Magnetic buoyancy of a flux tube. {\bf a.} A nearly
horizontal flux tube under the solar surface. {\bf b.} The flux
tube after its upper part has risen through the solar surface.}
\end{figure}
It can be shown that
magnetic buoyancy is particularly destabilizing in the interior of
the convection zone, where convective instability and magnetic
buoyancy reinforce each other. On the other hand, if a region is stable
against convection, then magnetic buoyancy can be partially
suppressed there (see, for example, \S8.8 in [22]). Since the
toroidal flux tube is produced at the bottom of the convection zone,
we may expect some parts of it to come into the convection zone and become
buoyant, whereas other parts may remain underneath the bottom of the
convection zone and stay anchored there due to the suppression of
magnetic buoyancy. A part of the flux tube coming within the convection
zone is expected to rise and eventually reach the solar surface to
form sunspots, as sketched in Fig.~7. In order to model the formation
of bipolar sunspots, we have to study the dynamics of flux tubes
rising through the convection zone due to magnetic buoyancy.
The best way to study this problem is to treat it as an initial-value problem.
First, an initial configuration with a magnetic flux ring at the
bottom of the convection zone, having a part coming inside
the convection zone, is specified, and then its subsequent
evolution is studied numerically. The evolution depends on the
strength of magnetic buoyancy, which is in turn determined by the
value of the magnetic field. We shall give arguments in \S4
why most of the dynamo theorists till the early 1990s believed
that the magnetic energy density should be in equipartition with the
kinetic energy density of convection, i.e.
$$\frac{B^2}{2 \mu} \approx \frac{1}{2} \rho v^2. \eqno(3)$$
This suggests $B \approx 10^4$ G on the basis of standard models
of the convection zone. If we use full MHD equations to study the
evolution of the flux tube, then the calculations become extremely
complicated. However, if the radius of cross-section of the flux tube is
smaller than the various scale heights, then it is possible to derive
an equation for flux tube dynamics from the MHD equations [23,
24]. Even this flux tube equation is a sufficiently
complicated nonlinear equation and has to be solved numerically.
The evolution of such magnetic flux tubes due to magnetic buoyancy
(starting from the bottom of the convection zone) was studied by Choudhuri
and Gilman [25] and Choudhuri [26]. It was found that the Coriolis
force due to the Sun's rotation plays a much more important role
in this problem than what anybody suspected before. If the initial
magnetic field is taken to have a strength of around $10^4$ G as
suggested by (3), then
the flux tubes move parallel to the rotation axis and emerge at
very high latitudes rather than at latitudes where sunspots are
seen. Only if the initial magnetic field is taken as strong as
$10^5$ G, then magnetic buoyancy is strong enough to overpower
the Coriolis force and the magnetic flux tubes can rise radially
to emerge at low latitudes.
\begin{figure}[t]
\centerline{\includegraphics[height=7cm,width=9cm]{fig8.eps}}
\caption{Plots of sin(tilt) against sin(latitude) theoretically
obtained for different initial values of magnetic field indicated
in kG. The observational data indicated by the straight line
fits the theoretical curve for initial magnetic field 100 kG
(i.e. $10^5$ G). Reproduced from D'Silva and Choudhuri [27].}
\end{figure}
D'Silva and Choudhuri [27] extended these calculations to look
at the tilts of emerging bipolar regions at the surface. These tilts
are also produced by the action of the Coriolis force on the
rising flux tube. Fig.~8
taken from [27] shows the
observational tilt vs.\ latitude
plot of bipolar sunspots (i.e.\ Joy's law) along with the theoretical
plots obtained by assuming different values of the initial magnetic
field. It is clearly seen that theory fits observations only if
the initial magnetic field is about $10^5$ G. If the magnetic
field is much stronger, then the Coriolis force is unable to
produce much tilt. On the other hand, flux tubes with weaker
magnetic fields are diverted to high latitudes. Apart from providing
the first quantitative explanation of Joy's law nearly three-quarters
of a century after its discovery, D'Silva and Choudhuri [27] put the first
stringent limit on the value of the toroidal magnetic field at the
bottom of the convection zone. Several other authors [28, 29]
soon performed similar calculations and confirmed
the result. Initially some efforts were made to explore whether
flux tubes with magnetic field given by (3) could satisfy various
observational constraints by invoking extra effects [30, 31].
However, the evidence kept mounting that the magnetic field
at the bottom of the convection zone is indeed much stronger than
the equipartition value given by (3).
We already mentioned that the tilts of
active regions have a large amount of scatter around the mean given
by Joy's law. In fact, it is found that active regions
often emerge with initial tilts inconsistent with Joy's law and then
the tilts change in the next few days to come closer
to values given by Joy's law [32]. Longcope and Choudhuri [33]
have argued that the vigorous convective turbulence in the upper layers
of the convection zone exerts a random force on the tops of the rising
flux loops, causing the scatter around the Joy's law, and then the tilt
of the flux tube relaxes to the appropriate value after the emergence
of the top of the tube through the solar surface when
the top is no longer kicked by convective turbulence.
\section{Modelling the cycles from flux transport dynamo}
If we begin by assuming the Sun to have a poloidal field as shown
in Fig.~6a, we saw that various properties of sunspot pairs
can be explained. The differential rotation would stretch this poloidal field
to produce the toroidal field, the interaction with convection
would lead to toroidal flux tubes and then magnetic buoyancy
would make these flux tubes rise to produce the bipolar sunspots.
However, if there is no mechanism to replenish the poloidal
field, then it would decay away and ultimately the whole
process outlined here would stop. We now turn to the question
how the poloidal field is produced. We invoke a mechanism
first proposed by Babcock [34] and Leighton [35].
The name of Leighton should be known to most physicists
as the second author of the celebrated {\em Feynman Lectures}
[36].
Let us now explain what this Babcock--Leighton mechanism is.
We pointed out in \S2 that bipolar sunspots have tilts increasing
with latitude, in accordance with Joy's law. Then we discussed in \S3
how this law was explained by D'Silva and Choudhuri [27] by
considering the action of the Coriolis force on rising flux tubes.
Now, a typical sunspot lives for a few days and the magnetic field
of the sunspot diffuses in the surrounding region by turbulent
diffusion after its decay. When a tilted bipolar sunspot pair
with the right spot nearer the equator and the left spot
at a higher latitude decays, the polarity of the right sunspot
gets more diffused in the lower latitudes and the polarity of the
left sunspot gets more diffused in the higher latitudes.
Take a look at Fig.~3 to visualize this process. This process
essentially gives rise to a poloidal field at the solar surface.
Since sunspots form from the toroidal field due to magnetic buoyancy,
a tilted bipolar sunspot pair can be viewed as a conduit through
which a part of the toroidal field ultimately gets transformed
into the poloidal field. The tilted sunspot pair forms from
the toroidal and we get the poloidal field after its decay.
This is the basic idea of poloidal
field generation proposed by Babcock [34] and Leighton [35].
It may be noted that this Babcock--Leighton mechanism is somewhat
different from the original proposal of Parker [4], which
was elaborated further by Steenbeck, Krause \& R\"adler [37].
According to this original proposal, the turbulence in the
convection zone would involve helical motions due to the Coriolis
force and the toroidal field would be
twisted by this helical turbulence to produce the poloidal
field. However, this process,
often known as the $\alpha$-effect, can occur only if the maximum
value of the toroidal magnetic field is such that the magnetic
energy density does not exceed the kinetic energy of turbulence,
as indicated by (3). As we already pointed out, flux tube
simulations for modelling sunspot formation suggest that the
toroidal field is about one order of magnitude stronger (about
$10^5$ G) compared to what we get from (3) (about $10^4$ G).
If the toroidal field is so strong, then the $\alpha$-effect
as originally envisaged by Parker [4] cannot work and the
Babcock--Leighton mechanism seems to be the likely mechanism
by which the poloidal field is produced.
\begin{figure}
\center
\includegraphics[width=6cm]{fig9.eps}
\caption{A cartoon explaining how the solar dynamo works within
the convection zone.}
\end{figure}
Fig.~9 is a cartoon encapsulating how the solar dynamo operates.
If you understand this cartoon, then you would have got the
central point of this presentation!
The toroidal field is produced in the tachocline by the differential
rotation stretching out the poloidal field. Then this toroidal
field rises due to magnetic buoyancy to produce bipolar sunspots
at the solar surface, where the poloidal field is generated by
the Babcock--Leighton mechanism from these bipolar sunspots.
The poloidal field so generated is carried by the meridional
circulation first to the polar region and then underneath the
surface to the tachocline to be stretched by the differential
rotation---thus completing the cycle. The likely streamlines
of meridional circulation are indicated in Fig.~9. This type
of dynamo model in which the meridional circulation plays
a crucial role is
called a {\it flux transport dynamo}.
Most of the dynamo theorists
at the present time believe that the solar dynamo operates in this
way. Wang, Sheeley and Nash [38] proposed the idea of the flux
transport dynamo. Choudhuri, Sch\"ussler and Dikpati [39] and
Durney [40] were the first to construct two-dimensional models
of the flux transport dynamo to demonstrate that such a dynamo
really does work. Initially it was thought that this type of
dynamo model would not work due to a technical reason. There is
a rule, known as the {\it Parker-Yoshimura sign rule}
[4, 41], which suggests that the type of dynamo outlined
in Fig.~9 would produce a poleward dynamo wave. In other words,
it was feared that such a theoretical model would suggest that sunspots should
appear at higher and higher latitudes with the progress of the
sunspot cycle rather than at lower and lower latitudes. Choudhuri,
Sch\"ussler and Dikpati [39] solved this puzzle by demonstrating
that a sufficiently strong meridional circulation can override
the Parker-Yoshimura sign rule and make the dynamo wave propagate
equatorward. This paved the way for the subsequent growth of
the flux transport dynamo model.
So far in this presentation I have avoided getting into equations.
For those who wish to see the equations, I now show the central
equations of the flux transport dynamo theory.
In spherical coordinates, we write the
magnetic field as
$$
{\bf B} = B (r, \theta) {\bf e}_{\phi} + \nabla \times [ A
(r, \theta) {\bf e}_{\phi}],
\eqno(4)$$
where $B (r, \theta)$ is the toroidal component and $A (r, \theta)$
gives the poloidal component. We can write the velocity field
as ${\bf v} + r \sin \theta \, \Omega (r, \theta) {\bf e}_{
\phi}$, where
$\Omega (r, \theta)$ is the angular velocity in the interior of the
Sun and ${\bf v}$ is the velocity of meridional circulation having components
in $r$ and $\theta$ directions. Then the main equations telling us
how the poloidal and the toroidal fields evolve with time are
$$
\frac{\partial A}{\partial t} + \frac{1}{s}({\bf v}.\nabla)(s A)
= \lambda_T \left( \nabla^2 - \frac{1}{s^2} \right) A + \alpha B,
\eqno(5)$$
$$ \frac{\partial B}{\partial t}
+ \frac{1}{r} \left[ \frac{\partial}{\partial r}
(r v_r B) + \frac{\partial}{\partial \theta}(v_{\theta} B) \right]
= \lambda_T \left( \nabla^2 - \frac{1}{s^2} \right) B
+ s({\bf B}_p.\nabla)\Omega + \frac{1}{r}\frac{d\lambda_T}{dr}
\frac{\partial}{\partial{r}}(r B), \eqno(6)$$
where $s = r \sin \theta$ and $\lambda_T$ is the turbulent
diffusivity inside the convection zone. We should point out that
(5) and (6) are
mean field equations obtained by averaging over the turbulence
in the convection zone and describe the mean behaviour of
the average magnetic field. Since (5) and (6)
are coupled partial differential equations,
nothing much can be done
analytically. Our research group in IISc
Bangalore has developed a numerical
code {\it Surya} for studying the flux transport dynamo problem
by solving (5) and (6). I can
send the code {\it Surya} and a detailed guide for using it
to anybody who sends a request to my e-mail address <EMAIL>.
\begin{figure}
\center
\includegraphics[width=7cm]{fig10.eps}
\caption{A theoretical butterfly diagram
of sunspots superposed on contours of constant $B_r$ at the solar surface
in a time-latitude plot. This figure is taken from Chatterjee, Nandy and
Choudhuri [43].}
\end{figure}
Some of the first results obtained with {\it Surya} were presented by Nandy and
Choudhuri [42] and Chatterjee, Nandy and Choudhuri [43]. I may mention
that a modified version of {\it Surya} has even been used to study the
evolution of magnetic fields in neutron stars [44, 45].
Fig.~10 shows a theoretical butterfly diagram
of sunspots, superposed on contours in the time-latitude plot of the
poloidal field on the solar surface. This theoretical figure obtained
by the code {\it Surya} has to be
compared with the corresponding observational figure given in Fig.~2.
Given the fact that this was one of the first efforts of reproducing this
observational figure from a theoretical model, hopefully most readers will
agree that the match between theory and observations is not too bad.
The original flux transport dynamo model of Choudhuri, Sch\"ussler
and Dikpati [39] was developed at a time when Mausumi Dikpati was
my PhD student. Afterwards she went to work in HAO Boulder and
the parent model led to two offsprings: a high diffusivity model
and a low diffusivity model. The diffusion times across the
convection zone in these two models
are of the order of 5 years and 200 years respectively. The high
diffusivity model has been developed in IISc Bangalore by me
and my successive PhD students
(Choudhuri, Nandy, Chatterjee, Jiang, Karak), whereas the low diffusivity
model has been developed by Dikpati and her co-workers in HAO (Dikpati,
Charbonneau, Gilman, de Toma). The differences between these models
have been systematically studied by Jiang, Chatterjee and Choudhuri
[46] and Yeates, Nandy and Mckay [47]. Both these models are
capable of producing oscillatory solutions resembling solar
cycles. However, when we try to study the variabilities of the
cycles, the two models give completely different results. We need
to introduce fluctuations to cause variabilities in the cycles.
In the high diffusivity model, fluctuations spread all over the
convection zone in about 5 years. On the other hand, in the low
diffusivity model, fluctuations essentially remain frozen during
the cycle period. Thus the behaviours of the two models are totally
different on introducing fluctuations. It may be mentioned that simple
mixing length arguments suggest a reasonably high turbulent diffusivity
(see p.\ 629 of [22]) consistent with what is used in the high
diffusivity model of the IISc Bangalore group.
\section{Irregularities of solar cycles and prospects for predicting
future cycles}
Before coming to the question of what causes the irregularities
of solar cycles, we take another look at the plot of polar
fields in Fig.~4. The polar field at
the end of cycle~22 was weaker than the polar field in the previous
sunspot minimum. We see that this weaker polar field was followed by
the cycle~23 which was weaker than the previous cycle. Does this
mean that there is a correlation between the polar field during
a sunspot minimum and the next sunspot cycle? In the left
panel of Fig.~11, we plot
the polar field in the sunspot minimum along the horizontal axis and
the strength of the next cycle along the vertical axis. Although
there are only 3 data points so far, they lie so close to a
straight line that one is tempted to conclude that there is
a real correlation. There is a joke that astrophysicists
often do statistics with one data point, whereas here
we have three! On the other hand, the right panel
of Fig.~11, which has
the cycle strength along the horizontal axis and the polar field
at the end of that cycle along the vertical axis, has points
which are scattered around. Choudhuri, Chatterjee and Jiang
[48] proposed the following to explain these observations.
While an oscillation between toroidal and poloidal components
takes place, the system gets random kicks at the epochs indicated
in Fig.~12. Then the poloidal field and the next toroidal field
should be correlated, as suggested by the left panel of
Fig.~11. On the other
hand, the random kick ensures that the toroidal field is not
strongly correlated with the poloidal field coming after it, as seen
in the right panel of Fig.~11.
\begin{figure}
\begin{minipage}[c]{0.55\textwidth}
\vspace{0.1cm}
\includegraphics[height=5cm,width=5cm]{fig11a.eps}
\end{minipage}%
\hspace{0.3cm}
\begin{minipage}[c]{0.6\textwidth}
\includegraphics[height=4.5cm,width=5cm]{fig11b.eps}
\end{minipage}
\caption{The left panel shows a plot of the strength
of cycle $n+1$ against the polar field at the end of
cycle $n$. The right panel shows a plot of the polar field
at the end of cycle $n$ against the strength of the cycle
$n$. From Choudhuri [49].}
\end{figure}
If there is really a correlation between the polar field at the
sunspot minimum and the next cycle, then one can use the polar field
to predict the strength of the next cycle [50].
Since the polar field in the just concluded minimum
has been rather weak (as seen in
Fig.~4), several authors [51, 52]
suggested that the coming cycle~24 will be rather
weak. Very surprisingly, the first theoretical prediction based
on a dynamo model made by Dikpati and Gilman [53] is that the
cycle~24 will be very strong. Dikpati and Gilman [53]
assumed the generation of the poloidal field from the toroidal
field to be deterministic, which is not supported by observational
data shown in the right panel
of Fig.~11. Tobias, Hughes and Weiss [54] make the
following comment on this work:
``Any predictions made with such models should be treated with extreme
caution (or perhaps disregarded), as they lack solid physical underpinnings.''
While we also consider many aspects of the Dikpati--Gilman work wrong which
will become apparent to the reader soon, we
cannot also accept the opposite extreme viewpoint
of Tobias, Hughes and Weiss [54], who suggest that the solar dynamo is a
nonlinear chaotic system and predictions are impossible or useless. If that
were the case, then we are left with no explanation for the correlation
seen in the left panel of Fig.~11.
\begin{figure}
\center
\includegraphics[width=12cm]{fig12.eps}
\caption{A schematic cartoon of the oscillation between toroidal
and poloidal components, indicating the epochs when the system
is subjected to random kicks.}
\end{figure}
Let us now finally come to the
theoretical question as to what produces the variabilities
of cycles and whether we can predict the strength of a cycle before its advent.
Some processes in nature can be predicted and some not. We can easily
calculate the trajectory of a projectile by using elementary mechanics.
On the other hand, when a dice is thrown, we cannot predict which side
of the dice will face upward when it falls. Is
the solar dynamo more like the trajectory
of a projectile or more like the throw of a dice? Our point of view is that the
solar dynamo is not a simple unified process, but a complex combination of
several processes, some of which are predictable and others not. Let
us look at the processes which make up the solar dynamo.
The flux transport dynamo model combines three basic processes. (i) The
strong toroidal field is produced by the stretching of the poloidal
field by differential rotation in the tachocline.
(ii) The toroidal field generated in the tachocline
gives rise to sunspots due to magnetic buoyancy and then the decay
of tilted bipolar sunspots produces the poloidal field by the
Babcock--Leighton mechanism. (iii) The poloidal field is
advected by the meridional circulation first to high latitudes and then down
to the tachocline, while diffusing as well. We believe that
the processes (i) and (iii) are reasonably ordered and deterministic.
In contrast, the process (ii) involves an element of randomness due
to the following reason. The poloidal field produced from the decay of
a tilted bipolar region by the Babcock--Leighton process depends on
the tilt. While the average tilt of bipolar regions at a certain
latitude is given by Joy's law, we observationally find quite a large
scatter around this average.
As we already pointed out, the action
of the Coriolis force on the rising flux tubes gives rise to Joy's
law [27], whereas convective buffeting of
the rising flux tubes in the upper layers of the convection zone causes the
scatter of the tilt angles [33].
This scatter in the tilt angles certainly introduces a
randomness in the generation process of the poloidal field
from the toroidal field. Choudhuri,
Chatterjee and Jiang [48] identified it as the main source of
irregularity in the dynamo process, which is in agreement
with Fig.~12. It may be noted that Choudhuri [55] was the
first to suggest several years ago
that the randomness in the poloidal field
generation process is the source of fluctuations in the dynamo.
\begin{figure}
\center
\includegraphics[width=5cm]{fig13.eps}
\caption{A sketch
indicating how the poloidal field produced at C during a
maximum gives rise to the polar field at P during the
following sunspot minimum and the toroidal field at T during the next
sunspot maximum. From [46].}
\end{figure}
The poloidal field gets built up during the declining phase of
the cycle and becomes concentrated near the poles during the
sunspot minimum.
The polar field at the sunspot minimum produced in a theoretical
mean field dynamo
model is some kind of `average' polar field during a typical
sunspot minimum. The observed polar field during
a particular sunspot minimum may be stronger or weaker than this average
field. The theoretical dynamo model has to be updated by feeding
the information of the observed polar field in an appropriate
way, in order to model actual cycles.
Choudhuri, Chatterjee and Jiang [48] proposed to model this
in the following way. They ran the dynamo code from a
minimum to the next minimum in the usual way. After stopping
the code at the minimum,
the poloidal field of the theoretical model was multiplied
by a constant factor everywhere above $0.8 R_{\odot}$ to bring
it in agreement with the observed poloidal field. Since
some of the poloidal field at the bottom of the convection zone may
have been produced in the still earlier cycles, it is left unchanged
by not doing any updating below $0.8 R_{\odot}$. Only the poloidal
field produced in the last cycle which is concentrated in the upper
layers gets updated. After
this updating which takes care of the random kick shown
in Fig.~12, we run the code till the next minimum, when the code
is again stopped and the same procedure is repeated.
Our solutions are now no longer
self-generated solutions from a theoretical model alone, but are
solutions in which the random aspect of the dynamo process has been
corrected by feeding the observational data of polar fields into
the theoretical model.
Before presenting the results obtained with this procedure, we come
to the question how the correlation between the polar field at the sunspot
minimum and the strength of the next cycle as seen in the left panel of Fig.~11 may
arise. This was first explained by Jiang, Chatterjee and Choudhuri
[46]. The Babcock--Leighton process would first produce the poloidal
field around the region C in Fig.~13. Then this poloidal field will
be advected to the polar region P by meridional circulation and will
also diffuse to the tachocline T. In the high diffusivity model, this
diffusion will take only about 5 years and the toroidal field of the
next cycle will be produced from the poloidal field that has diffused
to T. If the poloidal field produced at C is strong, then both the
polar field at P at the end of the cycle and the toroidal field at T
for the next cycle will be strong (and vice versa). We thus see that
the polar field at the end of a cycle and the strength of the next
cycle will be correlated in the high diffusivity model. But this
will not happen in the low diffusivity model where it will take more
than 100 years for the poloidal field to diffuse from C to T
and the poloidal field reaches the tachocline only due to the
advection by meridional circulation taking a time of about 20 years. If we
believe that the 3 data points in the left panel of
Fig.~11 indicate a real correlation,
then we have to accept the high diffusivity model!
\begin{figure}
\center
\includegraphics[width=11cm,height=4.5cm]{fig14.eps}
\caption{The theoretical monthly sunspot number
(solid line) for the last few years as well as
the upcoming next cycle, plotted along with the observational data (dashed line)
for the last few years. From Choudhuri, Chatterjee and
Jiang [48].}
\end{figure}
Finally the solid line in
Fig.~14 shows the sunspot number calculated from our high
diffusivity model [48]. Since
systematic polar field measurements are available only from the
mid-1970s, the procedure outlined above could be applied only from
that time. It is seen from Fig.~14 that our model matches the last
three cycles (dashed line) reasonably well and predicts a weak cycle 24.
It should be stressed that this is an inevitable consequence of
the high diffusivity model in which the strength of the cycle is
correlated with the polar field in the previous sunspot minimum and we
have fed the information in our calculation that the polar field
in the just-concluded minimum was weak.
We now wait for the Sun-god to give
a verdict on this prediction within a couple of years.
It may be mentioned that over the last few years several authors
[43, 46, 56--59] have given several independent arguments in support
of the high diffusivity model. If the next cycle~24 turns out to
be weak (for which there are already enough indications), then
that will provide a further support for the high diffusivity model.
One important related question is whether our dynamo model
can explain occurrences of extreme events like the Maunder
minimum in the seventeenth century. Choudhuri and Karak [60]
showed that the flux transport dynamo model can reproduce
the Maunder minimum if we introduce a set of assumptions in
the theoretical model. Whether this set of assumptions necessary
for producing the Maunder minimum is justified on statistical
grounds is an important question which needs to be investigated.
While the fluctuations in the Babcock--Leighton process seem to
be the main source of irregularities in the sunspot cycle, the
meridional circulation also has fluctuations and it has become
apparent in the last few years that the fluctuations in meridional
circulation also introduces irregularities in sunspot cycles [61]. Since
this topic has started being studied systematically only
recently [62--64], it would be premature to provide a summary
of it here. It seems that the nonlinear aspects
of the equations can also play important roles and there are some
indications that the solar dynamo may be close to a point of chaotic
bifurcation [65, 66]. We are certainly
far from a full theoretical understanding of the irregularities
of the sunspot cycle.
\def{\it Astrophys.\ J.}{{\it Astrophys.\ J.}}
\def{\it Mon.\ Notic.\ Roy.\ Astron.\ Soc.}{{\it Mon.\ Notic.\ Roy.\ Astron.\ Soc.}}
\def{\it Solar Phys.}{{\it Solar Phys.}}
\def{\it Astron.\ Astrophys.}{{\it Astron.\ Astrophys.}}
\def{\it Geophys.\ Astrophys.\ Fluid Dyn.}{{\it Geophys.\ Astrophys.\ Fluid Dyn.}}
|
\section{INTRODUCTION}
Graphene is attracting much interest not only owing to its novel physical properties,\cite{GeimRev,KatsnelsonRev,NetoRev} but also because of possible applications as a candidate material to replace silicon in future electronic devices.\cite{GeimRev2,AvourisRev,FerraiRev} In addition to its high charge carrier mobility,\cite{Morozov,Bolotin} its superior thermal properties are considered to be crucial in high-density large-scale integrated circuits where heat management is becoming more important as the density of devices grows.\cite{Pop} Balandin \textit{et al.} first reported extremely large values for the thermal conductivity ($\kappa$) in the range of $4840\pm440$ to $5300\pm480$~Wm$^{-1}$K$^{-1}$ for mechanically exfoliated single-layer graphene near room temperature.\cite{Balandin} These values are among the largest ever measured from any material so far.
Several groups since have measured $\kappa$ of mechanically exfoliated\cite{Balandin,Faugeras,Seol} or chemical vapor deposition (CVD)-grown\cite{Cai,Chen} graphene samples using different methods. For \emph{suspended}, exfoliated single-layer graphene, Faugeras \textit{et al.} reported a value of $\sim$630~Wm$^{-1}$K$^{-1}$ at 660 K,\cite{Faugeras} much lower than those of Balandin \textit{et al.}\cite{Balandin} On CVD-grown graphene, Ruoff's group\cite{Chen,Cai} reported $\kappa$ values ranging from $(2500 +1100/-1050)$~Wm$^{-1}$K$^{-1}$ to (2600$\pm900$ to 3100$\pm1000$)~Wm$^{-1}$K$^{-1}$ at 350~K. Since $\kappa$ is in principle a function of temperature and the measured values may be affected by the residual chemicals left on the samples as a result of the sample preparation processes, direct comparison of these later values with those of Balandin \textit{et al.} has been difficult. Given the importance of this key parameter for device applications, an accurate measurement and critical comparison with previous measurements are crucial. Here, we present the measurement of $\kappa$ for \emph{suspended} single-layer graphene at temperatures between 300~K and 500~K using Raman scattering spectroscopy on a clean sample prepared directly on a prepatterned substrate without involving a transfer process.
\section{EXPERIMENTAL}
\begin{figure}[b!]
\includegraphics{fig1.eps}
\caption{(Color online) (a) Raman spectrum of suspended graphene. (b) Optical microscope and (c) scanning electron microscopy images of suspended graphene sample. The scale bars are 10 $\mu$m. (d) Schematic diagram of the experiment.}
\label{fig:fig1.eps}
\end{figure}
The substrates with round holes with various diameters were prepared by photolithography and dry etching of Si substrates covered with a 300~nm-thick SiO$_{2}$ layer. The depth of the holes is $\sim$1.7~$\mu$m, deep enough to prevent interference from laser light reflected and scattered from the bottom of the holes.\cite{Yoon} The diameters were 2.6, 3.6, 4.6, and 6.6~$\mu$m. The graphene samples were prepared directly on the cleaned substrate by mechanical exfoliation from natural graphite flakes. No chemical treatment of the sample was involved in the preparation process. This ensures that the sample surface is free from chemical contaminants that may affect the measured $\kappa$ values. The sample used was a single-layer graphene flake of 35$\times$60~$\mu$m$^{2}$ dimensions identified by the line shape of the $2D$ band in the Raman spectrum\cite{Ferrari,Yoon2} (Fig.\ref{fig:fig1.eps}). The 514.5-nm (2.41~eV) beam of an Ar ion laser was focused onto the graphene sample by a 50$\times$ microscope objective lens (0.8 N.A.), and the scattered light was collected and collimated by the same objective. The scattered signal was dispersed with a Jobin-Yvon Triax 550 spectrometer (1800 grooves/mm) and detected with a liquid-nitrogen-cooled charge-coupled-device detector. The spectral resolution was about 0.7~cm$^{-1}$. The laser spot size was measured using the modified knife-edge method:\cite{Cai,Veshapidze} the Raman intensity of the Si phonon peak was monitored as the laser spot is scanned across the straight sharp edge of a Ti patch deposited on Si. By fitting the intensity to $I(r)=I_0e^{-2r^2/w^2}$, $w = 0.29~\mu$m was obtained. Figure \ref{fig:fig1.eps} shows a typical Raman spectrum of suspended single-layer graphene obtained with a laser intensity of 1.0~mW. There is no indication of the defect-induced $D$ peak, attesting to the high quality of the sample. Although all our measurements were performed on a single piece of a graphene sample, there are some hole-to-hole variations in the low-power Raman spectrum, indicating some inhomogeneities.
\section{RESULTS AND DISCUSSION}
The absorption of the laser beam by the sample induces local heating that raises the temperature in the vicinity of the laser spot. In a steady state, there exists a temperature gradient that depends on the total power supplied by the laser beam, $\kappa$, and the boundary conditions at the edge of the hole. The local temperature at the laser spot can be estimated from the shift of the Raman $G$ or $2D$ bands. The temperature dependence of the Raman spectrum of graphene has been studied by several groups.\cite{Cai,Chen,Calizo,Allen,Jo} Most of the studies were conducted on graphene samples on substrates. In those cases, the Raman spectrum may be affected by the strain induced by the difference in the thermal expansion coefficients of the substrate and graphene, in addition to the purely thermal effect. Since graphene samples suspended over a trench or a hole is less affected by the strain due to the thermal expansion coefficient difference, we used the values reported recently by Chen \textit{et al.}\cite{Chen} on \emph{suspended} graphene samples. They measured the temperature coefficients of the $G$ and $2D$ bands and found that the $2D$ band ($\partial\omega_{2D}/{\partial}T=-0.072\pm0.002$cm$^{-1}$K$^{-1}$) is more sensitive to temperature than the $G$ band ($\partial\omega_{G}/{\partial}T=-0.044\pm0.003$cm$^{-1}$K$^{-1}$). Therefore, we used the $2D$ band for the estimate of the temperature in this work.
Figure \ref{fig:fig2.eps} summarizes the shift of the $2D$ band as a function of the incident laser power. As the laser power increases, the $2D$ band frequency redshifts due to increased heating. For smaller holes, the shift is smaller because efficient heat conduction to the substrate limits the temperature rise at the laser spot. It should be noted that the largest hole (6.6~$\mu$m) shows a smaller shift than the 4.6-$\mu$m hole. This trend was confirmed by repeated measurements on several holes. This may be explained in the following way. As the hole size increases, the laser spot moves away from the edge of the hole, reducing the heat conduction to the substrate, and the temperature should increase. Beyond a certain hole size, the temperature would saturate if one ignores the thermal conduction to the ambient air. In reality, the conduction to air, however small, would decrease the temperature for larger holes, resulting in a smaller temperature rise for the larger holes.
\begin{figure}
\includegraphics{fig2.eps}
\caption{(Color online) Shift of the Raman $2D$ band as a function of the laser power.}
\label{fig:fig2.eps}
\end{figure}
\begin{table*}
\caption{\label{tab:table1} Comparison on thermal conductivity of \emph{suspended} single-layer graphene.}
\begin{ruledtabular}
\begin{tabular}{c c c c c c}
Sample type & Temp. determination & Shape & $\kappa$ (Wm$^{-1}$K$^{-1}$) & $T$ & Ref.\\ \hline
Exfoliated, pristine & $2D$ band & Circular & $\sim$1800 & $\sim$325 K & This work \\
& & & $\sim$710 & $\sim$500 K & \\
Exfoliated, pristine & $G$ band & Trench & $\sim$4840--5300 & RT & Balandin \textit{et al.}\cite{Balandin} \\
Exfoliated, transfered & Stokes/anti-Stokes & Circular & $\sim$630 & $\sim$660 K & Faugeras \textit{et al.}\cite{Faugeras} \\
CVD, transfered & $G$ band & Circular & $\sim$2500 & $\sim$350 K & Cai \textit{et al.}\cite{Cai} \\
& & & $\sim$1400 & $\sim$500 K & \\
CVD, transfered & $2D$ band & Circular & $\sim$2600--3100 & $\sim$350 K & Chen \textit{et al.}\cite{Chen} \\
\end{tabular}
\end{ruledtabular}
\end{table*}
In order to estimate the thermal conductivity, we used the heat diffusion equation ignoring the heat conduction to the ambient air. We considered heat conduction through suspended graphene and supported graphene on the substrate as well as between graphene and the substrate. The substrate is assumed to be a heat sink at the ambient temperature. With cylindrical symmetry, one can write the heat diffusion equation as:
\begin{equation}\label{eq:1}
\kappa\frac{1}{r}\frac{d}{dr}\Big(r\frac{dT_1(r)}{dr}\Big)+q(r)=0\quad \textrm{for}\ r < R,
\end{equation}
\noindent
where $R$ is the radius of the hole, $r$ is the radial position, and $\kappa$ is the thermal conductivity of \emph{suspended} graphene. $q(r)=(I\alpha/t)exp(-2r^2/w^2)$ is the heat inflow per unit volume due to laser excitation, where $I$ is the laser intensity, $\alpha$ is the absorptance of light in single layer graphene (2.3\%),\cite{Nair,Stauber,Mak,Sheehy} and $t$ is the thickness of graphene (0.335~nm). Outside the hole, where graphene is supported by the substrate, the following equation applies:
\begin{equation}\label{eq:2}
\kappa^\prime\frac{1}{r}\frac{d}{dr}\Big(r\frac{dT_2(r)}{dr}\Big)-\frac{\sigma_i}{t}(T_2(r)-T_a)=0\quad \textrm{for}\ r \geq R,
\end{equation}
\noindent
where $\kappa^\prime$ is the thermal conductivity of \emph{supported} graphene (600~Wm$^{-1}$K$^{-1}$),\cite{Seol} $T_a$ is the ambient temperature, and $\sigma_i$ is the interfacial thermal conductance between graphene and SiO$_2$ (100~MWm$^{-2}$K$^{-1}$).\cite{Chen2} The general solutions to Eqs.~(\ref{eq:1}) and (\ref{eq:2}) are
\begin{eqnarray}
\label{eq:3}&&T_1(r)=c_1+c_2 \ln(r)+c_3Ei\Big(\frac{-r^2}{r_0^2}\Big)\quad \textrm{for}\ r < R, \\
\label{eq:4}&&T_2(\gamma)=c_4I_0(\gamma)+c_5K_0(\gamma)+T_a\quad \textrm{for}\ r \geq R,
\end{eqnarray}
\noindent
where $c_i$'s are arbitrary constants, $Ei(x)$ is an exponential integral, $I_0(x)$ and $K_0(x)$ are the zero-order modified Bessel functions, and $\gamma=r(\sigma_i/(\kappa^\prime t))^\frac{1}{2}$. For a converging solution, $c_4$=0. The boundary conditions are:
\begin{eqnarray}
& &T_2(r\to\infty)=T_a, \\
& &T_1(R)=T_2(\gamma)\vert_{r=R}, \\
& &-\kappa\frac{dT_1(r)}{dr}\vert_{r=R}=-\kappa^\prime \frac{dT_2(\gamma)}{dr}\vert_{r=R}, \\
& &-2{\pi}Rt\kappa^\prime\frac{dT_2(\gamma)}{dr}\vert_{r=R}=Q,
\end{eqnarray}
\noindent
where $Q$ is the total laser power absorbed. The coefficients $c_i$'s of Eqs.~(\ref{eq:3}) and (\ref{eq:4}) are determined from these boundary conditions. On the other hand, the measured temperature ($T_m$) is an weighted average of temperature inside the beam spot and can be approximated as
\begin{equation}\label{eq:9}
T_m\approx\frac{\int_{0}^{w}T_1(r)q(r)rdr}{\int_{0}^{w}q(r)rdr}.
\end{equation}
By comparing the measured $T_m$ with Eq.~(\ref{eq:9}), one can determine the thermal conductivity $\kappa$. Figure \ref{fig:fig3.eps} shows thus determined $\kappa$ as a function of the measured temperature. The error bars are quite large for lower temperatures because $\Delta\omega_{2D}$, which determines $T_m$, is quite small in comparison to the measurement resolution. Therefore, the $\kappa$ values at temperatures below 325 K are not very reliable. It seems that $\kappa$ decreases as the temperature increases: from $\sim$1800~Wm$^{-1}$K$^{-1}$ near 325~K to $\sim$710~Wm$^{-1}$K$^{-1}$ near 500~K.
\begin{figure} [t]
\includegraphics{fig3.eps}
\caption{(Color online) Thermal conductivity of suspended graphene as a function of measured temperature. The dotted curve is guide to eye.}
\label{fig:fig3.eps}
\end{figure}
Our thermal conductivity values are somewhat lower than those reported for CVD graphene.\cite{Chen} In that work, the measured $\kappa$ is 2600 -- 3100~Wm$^{-1}$K$^{-1}$ near 350~K. The major difference in that work is that they measured the transmittance ($I_t/I_0$) of the graphene sample and took $1-I_t/I_0$ as the absorptance. Their value was 3.4\%, which is 50\% larger than the recently determined value of 2.3\%.\cite{Nair,Stauber,Mak,Sheehy} Reflection and scattering of light on the sample surface and/or other loss of the transmitted light may account for the difference. If we use the absorptance of 3.4\%, we obtain a $\kappa$ value of $\sim$2700~Wm$^{-1}$K$^{-1}$ at $\sim$325~K, similar to the value for CVD-grown graphene. On the other hand, Faugeras \textit{et al.} reported $\kappa$ of $\sim$630~Wm$^{-1}$K$^{-1}$ at $\sim$660~K for exfoliated graphene,\cite{Faugeras} which is rather close to our result at 500~K. Another source of uncertainty in the analysis is the temperature coefficient of the Raman $2D$ band. A 20\% variation in the temperature coefficient value would result in a similar variation in the obtained $\kappa$ value.
In light of the above analysis, the initially reported value of 5300~Wm$^{-1}$K$^{-1}$ by Balandin \textit{et al.} seems to be significantly overestimated. The most significant difference between their analysis and those of recent publications including our work is the value of the absorptance $\alpha$ of single layer graphene. They used $\alpha=13$\%, which is several times larger than the value of 2.3\% accurately measured and theoretically analyzed by Nair \textit{et al}.\cite{Nair} If one uses $\alpha=2.3$\%, their $\kappa$ value would reduce to $\sim940$~Wm$^{-1}$K$^{-1}$.
\section{CONCLUSIONS}
The thermal conductivity of suspended single-layer graphene was measured as a function of temperature using Raman spectroscopy on pristine graphene samples prepared directly on a patterned substrate by mechanical exfoliation. By monitoring the temperature at the laser spot using the Raman $2D$ band, the thermal conductivity was deduced by analyzing heat diffusion equations. The obtained thermal conductivity values range from $\sim$1800~Wm$^{-1}$K$^{-1}$ near 325~K to $\sim$710~Wm$^{-1}$K$^{-1}$ near 500~K. Based on our result as well as other recent reports,\cite{Faugeras,Chen} the initially reported\cite{Balandin} value of 5300~Wm$^{-1}$K$^{-1}$ seems to be significantly overestimated.
\begin{acknowledgments}
This work was supported by Mid-career Researcher Program through NRF grant funded by the MEST (No. 2008-0059038). D.~Y.~acknowledges funding from the Seoul City Government. H.~K.~and S.~L.~are supported by WCU program through the NRF funded by the MEST (No. R31-2008-000-10057-0)
\end{acknowledgments}
|
\section{Introduction}
When all the real homomorphisms defined on an algebra of real
functions defined on a space are evaluations then we say that the
space is smoothly real-compact. There are many articles stating this
property for various spaces. In \cite{prusell}, \cite{ercan} it is
shown that the spaces of real continuous functions on $\mR$ and
$\mR^n$ are smoothly real-compact. In \cite{arias} this property has
been shown for the spaces of functions of class $C^k$
$(k=1,\dots,\infty)$ on separable Banach spaces. Much information
about this topic can be found in \cite{gonz}. The most important
from the point of view of Sikorski spaces is the article
\cite{michor} since it discusses smooth real-compactness of smooth
spaces which are a wider category than the Sikorski spaces. Many
conditions for those spaces to be smoothly real-compact are given
there. In \cite{adam} and \cite{kriegl1} many important results were
obtained for a very wide class of algebras. In our article we
emphasize the concept of generators of a differential space. We use
techniques suitable for Sikorski spaces. Real valued homomorphisms
are classified by their values on generators.
\section{Basic concepts and definitions}
Let $M$ be a nonempty set and $\C$ a set of real functions on $M$.
We introduce on $M$ a topology $\tau_\C$, the weakest topology in
which the functions from $\C$ are continuous. We say that the set
$\C$ is \emph{closed with respect to superposition} if all functions
of a form $\omega\circ (f_1,\dots,f_n)$ where $f_1,\dots, f_n\in \C,
\quad \omega\in C^\infty(\mR^n)$, $n\in \mN$, are in $\C$. Adding to
$\C$ all the functions of this form we obtain what we will call the
superposition closure of $\C$, denoted by $sc\C$ following
Waliszewski \cite{wal1}. For any $A\subseteq M$ by $\C_A$ we denote
the set of all functions $f$ on $A$ such that for any $p\in A$ there
exists an open neighborhood $U\in \tau_\C$ of $p$ and a function
$g\in\C$ such that $f|_{U\cap A}=g|_{U\cap A}$. If $\C=\C_M$ then we
say that $\C$ is \emph{closed with respect to localization}
\cite{sik1}. We call the set of real functions $\C$ on a nonempty
set $M$ a
\emph{differential structure} if it is:\\
i) closed with respect to superposition, $\C=sc\C$,\\
ii) closed with respect to localization, $\C=\C_M$.\\
A differential structure is always an algebra with unity and
contains all constant functions.
\begin{defn}
A pair $(M,\C)$ is said to be a differential space if $M$ is a
nonempty set and $\C$ a differential structure on it.
\end{defn}
We define a \emph{differential subspace} of a differential space
$(M,\C)$ to be any pair $(A,\C_A)$ where $A\subseteq M$, $A\neq
\emptyset$.
\begin{defn}
The differential structure $\C$ is generated by a set of functions
$\C_0$ if\\ $\C=(sc\C_0)_M$.
\end{defn}
Thus $\C$ is the smallest differential structure that contains
$\C_0$. Sometimes we write $\C=\operatorname{Gen}\C_0$. If $\C=\operatorname{Gen}\C_0$ then for
any $f\in \C$ and any point $p\in M$ there exists an open
neighborhood $U\in \tau_{\C}$ of $p$ and functions
$f_1,\dots,f_n\in \C_0,\quad \omega\in C^\infty(\mR^n)$, $n\in \mN$
such that $f|_U=\omega\circ(f_1,\dots,f_n)|_U$. We say that the
differential space $(M,\C)$ is \emph{finitely generated} if it is
generated by a finite set of a real functions. A differential space
is \emph{countably generated} if it is generated by a countable set
of real functions but it is not finitely generated.
We denote by $(\mR^I,\V_I)$ the differential space with the
structure $\V_I$ generated by the set of projections $\C_0=\{\pi_i
:i\in I\}$, where $\pi_i:\mR^I\rightarrow \mR$ is defined by
$\pi_i(x)=x_i$ for $x=(x_i)_{i\in I}$. This is a generalization of
the Cartesian space $(\mR^n,\V_n)$ where $\V_n=C^\infty(\mR^n)$.
The \emph{spectrum} of an algebra $\C$ is the set
$$\operatorname{Spec}\C=\{\chi:\C\rightarrow \mR\},$$ where $\chi$ is a homomorphism that preserves unity. \\
Evaluation of the algebra $\C$ at a point $p\in M$ is the
homomorphism $\chi\in \operatorname{Spec}\C$ given by
\begin{equation}
\chi(f)=f(p) \quad \forall f\in \C.
\end{equation}
We will denote it by $\operatorname{ev}_p$. We define the mapping\\
$\operatorname{ev}:M\rightarrow \operatorname{Spec}\C$ by the formula:
\begin{equation}
\operatorname{ev}(p)=\operatorname{ev}_p.
\end{equation}
\begin{defn}
(\cite{michor}) We say that a differential space $(M,\C)$ is
smoothly real-compact if any $\chi\in \operatorname{Spec}\C$ is evaluation at some
point $p\in M$.
\end{defn}
From this definition it follows that the space $(M,\C)$ is smoothly
real-compact when the mapping $\operatorname{ev}$ is a surjection. For any $f\in
\C$ we define the function $\hat{f}:\operatorname{Spec}\C\rightarrow \mR$ by
\begin{equation}
\hat{f}(\chi)=\chi(f) \quad \forall \chi\in \operatorname{Spec}\C.
\end{equation}
The set of all functions of the form $\hat{f}$ will be denoted by
$\hat{\C}$. Define $\tau:\C\rightarrow \hat\C$ by
\begin{equation}
\tau(f)=\hat{f}\quad \forall f\in \C.
\end{equation}
The mapping $\tau$ is an isomorphism between the algebra $\C$ and
the algebra $\hat{\C}$.
\section{Main results}
\begin{lem} The differential space $(\mR^n,\V_n)$ is smoothly real-compact.\label{lem:rn}
\end{lem}
\begin{proof}
Let $\chi\in \operatorname{Spec}\V_n$. We define $p\in \mR^n$ by
$p_i:=\chi(\pi_i)$ for $i=1,\dots,n$. We will show that $\chi=ev_p$.
It is known that any $f\in \V_n$ can be represented as
\begin{equation}
f=f(p)+\sum_{i=1}^ng_i(\pi_i-p_i)\quad \text{for} \quad
g_1,\dots,g_n\in \varepsilon_n,
\end{equation}
where the functions $g_i$ satisfy $g_i(p)=\partial_if(p)$. Then
\[\chi(f)=\chi(f(p))+\sum_{i=1}^n\chi(g_i)(\chi(\pi_i)-\chi(p_i))=f(p)+\sum_{i=1}^n\chi(g_i)(p_i-p_i)=f(p)\].
Therefore $\chi(f)=f(p)$ for all $f\in \V_n$.
\end{proof}
Now we prove:
\begin{lem}
Every differential subspace of the differential space $(\mR^n,\V_n)$
is smoothly real-compact.\label{lem:findimsub}
\end{lem}
\begin{proof}
Let $(M,\C)$ be a differential subspace of $(\mR^n,\V_n)$. The
inclusion mapping $\iota_M:M\rightarrow \mR^n$ is smooth and
therefore $\iota_M^*:\V_n\rightarrow M$ is a homomorphism. From the
definition we know that $\iota_M^*(f)=f|_M$ for all $f\in \V_n$. For
any $\chi \in \operatorname{Spec}\C$ we have $\chi\circ\iota_M^*\in \operatorname{Spec}\V_n$.
From Lemma \ref{lem:rn} we know that there exists $p\in \mR^n$ such
that $\chi\circ\iota_M^*(f)=\chi(\iota_M^*(f))=\chi(f|_M)=f(p)$ for
all $f\in \V_n$. Suppose that $p\notin M$. Define $\omega\in \V_n$
by
\begin{equation}
\omega(x_1,\dots,x_n)=(x_1-p_1)^2+\dots+(x_n-p_n)^2.
\end{equation}
Since $\omega|_M>0$ we have $\frac{1}{\omega|_M}\in \C$. We also
know that $\chi((\omega|_M)(\frac{1}{\omega|_M}))=\chi(1)=1$, and
$(\chi\circ\iota_M^*)(\omega)=\chi(\omega|M)=\omega(p)=0$. This is a
contradiction.
We will show that $\chi=ev_p$. Let $f\in \C$. There exists an open
neighborhood $U\in \tau_{\V_n}$ of $p$ and a function $\kappa\in
\V_n$ such that $f|_{U\cap M}=\kappa|_{U\cap M}$. From \cite{sik} we
know that there exists a bump function $\phi\in \V_n$ with
$\phi(p)=1$, $\phi|_{M\cap U}>0$ and $\phi|_{\mR^n-(M\cap U)}=0$.
From these properties it follows that $(f-\kappa|_M)\phi|_M=0$. Then
$\chi((f-\kappa|_M)\phi|_M)=(\chi(f)-\chi(\kappa|_M))\chi(\phi|_M)=0$.
But $\chi(\phi|_M)=(\iota_M\circ\chi)(\phi)=\phi(p)=1$ so
$\chi(f)=\chi(\kappa|_M)=\kappa(p)=f(p)$. We have shown that
$\chi(f)=f(p)$ for all $f\in\C$.
\end{proof}
If the differential structure $\C$ of the differential space
$(M,\C)$ is generated by a set of functions $\C_0$ then we can
define a mapping $\phi:M\rightarrow \mR^{\C_0}$ by
\begin{equation}
\phi(p)(f)=f(p), \quad f\in \C_0.
\end{equation}
We will call this mapping the \emph{generator embedding}. We can
prove the following:
\begin{lem}
A differential space $(M,\C)$ with $\C=\operatorname{Gen}\C_0$ is smoothly
real-compact iff the differential space $(\phi(M),(\V_I)_{\phi(M)})$
for $I=|\C_0|$ is smoothly real-compact.\label{lem:embed}
\end{lem}
\begin{proof}
If $\C_0$ separates the points of $M$ then $\phi$ is a
diffeomorphism onto its image so the result is obvious. So assume
that $\C_0$ does not separate points. Then $\bar\phi:M\rightarrow
\phi(M)$ where $\bar\phi(p)=\phi(p)$ is surjective but not
injective. Set $F:=\bar\phi$. We know that
$F^*:(\V_I)_{\phi(M)}\rightarrow \C$ is an isomorphism of algebras.
If $(M,\C)$ is smoothly real-compact then for any $\nu\in
\operatorname{Spec}(\V_I)_{\phi(M)}$ there exists $\mu\in \operatorname{Spec}\C$ such that
$\mu=\nu\circ(F^*)^{-1}$. Then for any $g\in (\V_I)_{\phi(M)}$,
$\nu(g)=\mu(F^*(g))=\mu(g\circ F)=g(F(p))$. So if $\mu=ev_p$ then
$\nu=ev_{F(p)}$.
If $(\phi(M),(\V_I)_{\phi(M)})$ is smoothly real-compact then for
any $\mu\in \operatorname{Spec}\C$ there exists $\nu\in \operatorname{Spec}(\V_I)_{\phi(M)}$
defined by $\nu=\mu\circ F^*$, so $\mu=\nu\circ (F^*)^{-1}$.
Therefore for any $f\in \C$ we have $\mu(f)=(\nu\circ
(\phi^*)^{-1})(f)=\nu((\phi^*)^{-1}(f))=((\phi^*)^{-1}(f))(q)=f(p)$
for any $p\in F^{-1}(q)$. So if $\nu=ev_q$ then $\mu=ev_p$ for all
$p\in F^{-1}(q)$.
\end{proof}
From the last lemma we know that it is sufficient to work on
subspaces of Cartesian spaces.
\begin{cor}
Let $(M,\C)$ be a differential space with $\C=\operatorname{Gen}\C_0$ for some
finite $\C_0$. Then $(M,\C)$ is smoothly
real-compact.\label{cor:sub}
\end{cor}
\begin{proof}
By using the generators $\C_0$ we can embed $(M,\C)$ into
$(\mR^{\C_0},(\V_{\C_0})_{\phi(M)})$ and then from Lemmas
\ref{lem:findimsub}, \ref{lem:embed} we derive that $(M,\C)$ is
smoothly real-compact.
\end{proof}
From Corollary \ref{cor:sub} we obtain:
\begin{lem}
Let $(M,\C)$ be a differential space. Any $\chi\in \operatorname{Spec}\C$
satisfies the following condition:
\begin{equation}
\chi(\omega\circ (f_1,\dots,f_n))=\omega(\chi(f_1),\dots,\chi(f_n)),
\end{equation}
for all $\omega\in \V_n$ and $f_1,\dots,f_n\in \C$, $n\in
\mN$.\label{lem:comp}
\end{lem}
\begin{proof}
Let $\beta_1,\dots,\beta_n\in \C$ be arbitrary functions. We define
the mapping \\$F:(M,\C)\rightarrow (\mR^n,\V_n)$ by:
$$
F(p)=(\beta_1(p),\dots,\beta_n(p)),\quad p\in M.
$$
This mapping is smooth and it is onto its image. Therefore the
mapping\\ $F^*:(\V_n)_{F(M)}\rightarrow \C$ is a homomorphism. For
any $\chi\in \operatorname{Spec}\C$ we have $\chi\circ F^*\in
\operatorname{Spec}((\V_n)_{F(M)})$. From Corollary \ref{cor:sub} we know that
there exists $q\in F(M)$ such that $\chi\circ F^*=ev_q$ for some
$q\in F(M)$. Also there exists $p\in M$ such that
$$
(\chi\circ F^*)(\omega|_{F(M)})=ev_{F(p)}(\omega|_{F(M)}) \quad
\forall \omega\in \V_n.
$$
We can rewrite this in the form
$$
\chi(\omega\circ
F)=\omega(F(p))=\omega(\beta_1(p),\dots,\beta_n(p))\quad \forall
\omega \in \V_n. $$ By setting $\omega=\pi_i$, $i=1,\dots,n,$ we
obtain $\chi(\beta_i)=\chi(\pi_i\circ F)=\pi_i(F(p))=\beta_i(p)$ and
finally $\chi(\omega\circ
(\beta_1,\dots,\beta_n))=\omega(\chi(\beta_1),\dots,\chi(\beta_n))$
for all $\omega\in \V_n$.
\end{proof}
Now we prove the following:
\begin{lem}
Let $(M,\C)$ be a differential space such that $\C=\operatorname{Gen}\C_0$ and let
$\chi\in \operatorname{Spec}\C$. If $\chi|_{\C_0}=ev_p|_{\C_0}$ then
$\chi=ev_p$.\label{lem:5}
\end{lem}
\begin{proof}
First we will show that if $f\in sc\C_0$ then $\chi(f)=f(p)$ . From
Lemma \ref{lem:comp} we know that
$\chi(\omega\circ(\beta_1,\dots,\beta_n))=\omega(\chi(\beta_1),\dots,
\chi(\beta_n))$ for $\omega\in \V_n$ and $\beta_1,\dots,\beta_n\in
\C_0$. We also know that $\chi(\beta_i)=ev_p(\beta_i)=\beta_i(p)$.
We can write
$\chi(\omega\circ(\beta_1,\dots,\beta_n))=\omega(\beta_1(p),\dots,\beta_n(p))=\omega\circ(\beta_1,\dots,\beta_n)(p)=ev_p(\omega\circ
(\beta_1,\dots,\beta_n))$. So we see that
$\chi|_{sc\C_0}=ev_p|_{sc\C_0}$.
Now let $f\in \C$ be an arbitrary function. We know that for every
$p\in M$ there exists an open neighborhood $U\in \tau_\C$, functions
$\beta_1,\dots,\beta_n\in \C_0$ and a function $\omega\in \V_n$ such
that $f|_U=\omega\circ(\beta_1,\dots,\beta_n)|_U$. There also exists
a bump function $\psi$ which separates the point $p$ in the set $U$.
This function is constructed by compositing some function from
$\V_n$ with some generators from $\C_0$. We know that the
homomorphism $\chi$ equals evaluation at $p$ on this function, so
$\chi(\phi)=\phi(p)=1$. Now the following equality holds:
$\phi\cdot(f-\omega\circ(\beta_1,\dots,\beta_n))=0$. By applying the
homomorphism $\chi$ to this equality we obtain
$\chi(\phi)\cdot\chi(f-\omega\circ(\beta_1,\dots,\beta_n))=\chi(f)-\chi(\omega\circ(\beta_1,\dots,\beta_n))=0$
so
$\chi(f)=\chi(\omega\circ(\beta_1,\dots,\beta_n))=ev_p(\omega\circ(\beta_1,\dots,\beta_n))=f(p)=ev_p(f)$.
We see that $\chi(f)=f(p)$ for all $f\in \C$.
\end{proof}
From Lemma \ref{lem:5} we get:
\begin{cor}
The differential space $(\mR^I,\V_I)$ is smoothly real-compact.
\end{cor}
\begin{proof}
Let $\chi\in \operatorname{Spec}\V_I$ be any homomorphism. Define $p\in \mR^I$ by
$p_i=\chi(\pi_i)$ for $i\in I$. Then $\chi(\pi_i)=\pi_i(p)$ so
$\chi(\pi_i)=ev_p(\pi_i)$. Since the structure $\V_I$ is generated
by the set $\{\pi_i:i\in I\}$ we see that $\chi$ is evaluation at
$p$ on the generators. From the last Lemma we derive that $\chi$ is
an evaluation on the whole $\V_I$.
\end{proof}
By using the whole $\C$ as the set of generators we can embed $M$ in
$\mR^\C$. We denote this embedding by $\iota$ so $\iota:M\rightarrow
\mR^\C$, $\iota(p)_f=f(p)$. This is a special case of a generator
embedding. We can also map $\operatorname{Spec}\C$ into $\mR^\C$ using the mapping
$\kappa:\operatorname{Spec}\C\rightarrow \mR^\C$ defined by
$\kappa(\chi)_f=\hat{f}(\chi)=\chi(f)$. It is obvious that
$\iota=\kappa\circ ev$. In \cite{michor} Kriegl, Michor and
Schachermayer have shown that $\iota(M)$ is dense in
$\kappa(\operatorname{Spec}\C)$ in the Tikhonov topology of $\mR^\C$. Since the
mapping $\kappa$ is a homeomorphism one can easily see:
\begin{cor}
$ev(M)$ is dense in $\operatorname{Spec}\C$ in the topology
$\tau_{\hat{\C}}$.\label{cor:dense}
\end{cor}
This property will allow us to prove an interesting fact about the
space $(\operatorname{Spec}\C,\hat{\C})$.
\begin{lem} If $(M,\C)$ is a differential space
then $(\operatorname{Spec}\C,\hat\C)$ is a differential space.
\end{lem}
\begin{proof}
To prove that $(\operatorname{Spec}\C,\hat{\C})$ is a differential space, we have
to show that the set $\hat{\C}$ is closed with respect to
superposition with smooth functions from $\V_n$ and is closed with
respect to localization.
Let $g=\omega\circ (\hat{f_1},\dots,\hat{f_n})$ for some $\omega\in
\V_n$ and $\hat{f_1},\dots,\hat{f_n}\in \hat{\C}$. From Lemma
\ref{lem:comp} we know that\\
$g(\chi)=\omega\circ(\hat{f_1},\dots,\hat{f_n})(\chi)=\tau(\omega\circ
(f_1,\dots,f_n))(\chi)\quad \forall \chi\in \operatorname{Spec}\C$. We have shown
that $g\in \hat{\C}$ so $\hat\C$ is closed with respect to
superposition.
Let a function $f:\operatorname{Spec}\C\rightarrow \mR$ satisfy the localization
condition in the space $(\operatorname{Spec}\C,\hat{\C})$. For any open subset
$\hat{U}\in \operatorname{Spec}\C$ there is $\hat{g}\in \hat{\C}$ such that
$f|_{\hat{U}}=\hat{g}|_{\hat{U}}$. We can uniquely define a function
$h:M\rightarrow \mR$ by the condition $h(p)=f(ev_p)$ for all $p\in
M$. For any open set $\hat{U}\in \operatorname{Spec}\C$ the set $U=\{p\in M :
ev_p\in \hat{U}\}$ is open. From the definitions of $h$ and $U$ we
know that $h|_U=g|_U$. Because $g\in \C$ it follows that $h\in \C$.
We also know that $\hat{h}|_{evM}=f|_{evM}$. From Corollary
\ref{cor:dense} we derive that $f=\hat{h}$. This means that $f\in
\hat{\C}$ so $\hat{\C}$ is closed with respect to localization.
\end{proof}
Now one can prove the following lemmas:
\begin{lem}
If $(M,\C)$ is a differential space with the structure $\C$
generated by $\C_0$ then the differential structure $\hat{\C}$ of
the differential space $(\operatorname{Spec}\C,\hat\C)$ is generated by
$\hat{\C_0}$.
\end{lem}
\begin{proof}
Assume that $\C_0=\{f_i:i\in I\}$. We know that for any $f\in \C$
there exists an open covering of $M$ such that on each set $U$ of
this covering the function $f$ can be expressed in the form
$\omega\circ (f_1,\dots,f_n)$ where $f_1,\dots, f_n\in \C$ and
$\omega \in \V_n$. For each open set $U$ of the covering we define
$\hat{U}=\{ev_p\in \operatorname{Spec}\C: p\in U\}$. On the set $\hat{U}$ we
consider the function $\hat{f}=\tau(\omega\circ (f_1,\dots,f_n))$.
The sets of the form $\hat{U}$ might not be a covering of $\operatorname{Spec}\C$
but thair union is dense in $\operatorname{Spec}\C$. Therefore we can prolong
uniquely this representation of $\hat{f}$ on the whole $\operatorname{Spec}\C$. We
have shown that $\hat{\C}=\operatorname{Gen}\hat{\C_0}$.
\end{proof}
\begin{lem}
For any differential space $(M,\C)$ the differential space
$(\operatorname{Spec}\C,\hat{\C})$ is smoothly real-compact.
\end{lem}
\begin{proof}
We need to show that for every homomorphism $\hat\chi\in
\operatorname{Spec}\hat{\C}$ there exists a homomorphism $\psi\in \operatorname{Spec}\C$ such
that $\hat\chi=ev_\psi$. Since the algebras $\C$ and $\hat\C$ are
isomorphic we can define uniquely $\chi\in \operatorname{Spec}\C$ by the formula
$\chi(f)=\hat\chi(\hat{f})$. We will show that $\hat\chi=ev_\chi$.
Indeed $ev_\chi(\hat{f})=\hat{f}(\chi)=\chi(f)=\hat\chi(\hat{f})$.
\end{proof}
\begin{lem}\label{lem:uniq}
Let $(M,\C)$ be a differential space and $\C=\operatorname{Gen}\C_0$. If
$\chi_1,\chi_2\in \operatorname{Spec}\C$ are equal on the generators,
$\chi_1|_{\C_0}=\chi_2|_{\C_0}$, then they are equal,
$\chi_1=\chi_2$.
\end{lem}
\begin{proof}
Assume that $\chi_1|_{\C_0}=\chi_2|_{\C_0}$ and $\chi_1\neq\chi_2$.
From the last lemma we know that the differential structure
$\hat{\C}$ of the differential space $(\operatorname{Spec}\C,\hat{\C})$ is
generated by $\hat{\C}_0$. From the condition
$\chi_1|_{\C_0}=\chi_2|_{\C_0}$ we derive that
$\hat{f}(\chi_1)=\hat{f}(\chi_2)$, for all $\hat{f}\in \hat{\C}_0$.
But we know that if the generators do not separate points then all
the functions do not separate points so $\forall \hat{f}\in \hat{\C}
\hat{f}(\chi_1)=\hat{f}(\chi_2)$ and it follows that
$\chi_1(f)=\chi_2(f)$for all $f\in \C$ . This means that
$\chi_1=\chi_2$.
\end{proof}
\begin{lem}
If $(M,\C)$ is differential subspace of $(\mR^I,\V_I)$ then any
function $f\in \C$ is uniquely continuously prolongable to
$\tilde{f}:\tilde{M}\rightarrow \mR$, where $\tilde{M}=\{p\in \mR^I
: \exists \chi\in \operatorname{Spec}\C$ such that
$p_i=\chi(\pi_i|_M) \quad\forall i \in I\}$.\label{lem:prelong}
\end{lem}
\begin{proof}
We define $\tilde{f}(p)=\hat{f}(\chi)$ where $\chi\in \operatorname{Spec}\C$ is
such that $\chi(\pi_i)=p_i$ for all $i\in I$. Since a homomorphism
is uniquely defined by its value on the generators (Lemma
\ref{lem:uniq}) this definition is correct. We see that if $p\in M$
then $\chi=ev_p$ and $\tilde{f}(p)=\hat{f}(ev_p)=f(p)$ so this is
indeed a prolongation. This prolongation is continuous since the
function $\tilde{f}$ is a realization of the function $\hat{f}$ on
the set $\tilde{M}$ which is the image of $\operatorname{Spec}\C$ under the
generator embedding using the generators $\tau(\pi_i|_M), i\in I$.
Uniqueness follows from the fact that $M$ is dense in $\tilde{M}$ in
the topology of $\mR^I$.
\end{proof}
From Lemma \ref{lem:prelong} we obtain:
\begin{cor}\label{cor:prelong}
When $(M,\C)$ is a differential subspace of $(\mR^I,\V_I)$ generated
by $\C_0=\{\pi_i|_M: i\in I\}$ then the mapping
$\chi:\C_0\rightarrow\mR$ defined on generators
by$\chi(\pi_i|_M)=p_i$ for some $p\in\tilde{M}-M$ can be prolonged
to a homomorphism on the whole $\C$ iff all the functions from $\C$
are prolongable to $p$.
\end{cor}
Let $M=\mR^\mN-\{0\}$ and $\C_M=(\V_\mN)_M$. Then $(M,\C_M)$ is a
differential subspace of $(\mR^\mN,\V_\mN)$. We will show that this
space is smoothly real-compact.
\begin{lem}\label{lem:prelong1}
There exists a function $\xi\in \C_M$ which is not prolongable to
any continuous function on $\mR^\mN$.
\end{lem}
\begin{proof}
We know that there exists a function $\phi\in C^\infty(\mR)$
satisfying the
following properties:\\
1. $\forall x\in \mR\quad \phi(x)\in [0,1]$\\
2. $supp(\phi)\in (-\infty,1]$\\
3. $\phi|_{[0,\frac{1}{2}]}=1$\\
For any $k\in \mN$ we define $\tilde\rho_k:\mR^\mN\rightarrow \mR$
by the formula:
$$
\tilde{\rho}_k((x_n))=\sum_{i=1}^k x_i^2
$$
for $(x_n)\in \mR^\mN$. Then $\tilde{\rho}_k\in
\C^{\infty}(\mR^{\mN})$, and $\rho_k=\tilde{\rho}_k|_M \C_M$. We
define $\xi:M\rightarrow \mR$ by:
\begin{equation}\label{def:f}
\xi((x_n))=\sum_{k=1}^\infty \phi(k^2\rho_k((x_n))).
\end{equation}
We will show that this function belongs to the structure $\C_M$. For
any $k\in \mN$ we can define the closed subset $A_k=\{(x_n)\in M:
k^2\rho_k((x_n))\leq 1\}=\{(x_n)\in M:\rho_k((x_n))\leq
\frac{1}{k^2}\}$. We see that $\operatorname{supp}(\phi\circ(k^2\rho_k))\subseteq
A_k$. For any $(x_n)\in M$ the sequence $\rho_k((x_n))$ is
non-decreasing with respect to $k$ and there exists $k_0\in \mN$ for
which $\frac{1}{k^2}<\rho_{k_0}((x_n))$. This means that
$(x_n)\notin A_k$. Therefore $\bigcap_{k\in \mN} A_k=\emptyset$. We
also know that $A_{k+1}\subseteq A_k$. Let us define the family of
open subsets $U_k=M-A_k$. Of course $\bigcup_{k\in \mN}U_k=M$. If
$(x_n)\in U_k$ then $\phi(k^2\rho_k((x_n))=0$. Then for all $m>k$,
$x_n\in U_m$ so $\phi(m^2\rho_m((x_n)))=0$. This means that only a
finite number of elements are non-zero in the sum (\ref{def:f}) and
therefore
\begin{displaymath}
\xi((x_n))=\sum_{j=1}^{k-1}\phi(j^2\rho_j((x_n))),
\end{displaymath}
so $\xi|_{U_k}\in \C_{U_k}=(\C_M)_{U_k}$ for all $k\in \mN$. From
the localization closedness of the differential structure we derive
that $\xi\in \C_M$. Now we will define a sequence in $M$ convergent
to $0$ on which the function $\xi$ diverges. Let $z_k=(x_{n,k})$
where
\begin{displaymath}
x_{n,k}= \left\{
\begin{array}{lll}
\frac{1}{k\sqrt{2}} & \textrm{for} & n=k, \\
0 & \textrm{for} & n\neq k.
\end{array}\right.
\end{displaymath}
We can see that $ \lim_{k\rightarrow \infty}z_k=0\in \mR^{\mN} $ and
\begin{displaymath}
\rho_j((z_k))= \left\{
\begin{array}{lll}
\frac{1}{2k^2} & \textrm{for} & j\geq k, \\
0 & \textrm{for} & j<k.
\end{array}\right.
\end{displaymath}
For $j\leq k$ we obtain $\phi(j^2\rho_j((x_k)))=1$ and therefore
\begin{displaymath}
\xi((x_k))=\sum_{j=1}^{\infty}\phi(j^2\rho_j((x_k)))\geq\sum_{j=1}^k
1=k.
\end{displaymath}
This means that $\lim_{k\rightarrow{\infty}}\xi((x_k))=+\infty$. The
function $\xi$ is not prolongable to any continuous function in
$\mR^\mN$.
\end{proof}
Now we prove
\begin{lem}
The differential space $(M,\C_M)$ is smoothly real-compact.
\end{lem}
\begin{proof}
From Lemma \ref{lem:uniq} we know that the set $\operatorname{Spec}\C_M$ may
contain only one homomorphism $\chi_0$ which is not an evaluation.
This homomorphism would be defined on the generators by the formula
$\chi_0(\pi_i|_M)=0$ for all $i\in I$. So there would be only one
point $0\in \tilde{M}-M$. But it cannot be so since from Corollary
\ref{cor:prelong} we know that all the functions from $\C_M$ are
prolongable to the point $0$. From the last lemma we know that there
exists a function $\xi\in \C_M$ which is not prolongable.
\end{proof}
One can easily see
\begin{cor}
For any $p\in \mR^\mN$ the differential space
$(\mR^\mN-\{p\},(\V_\mN)_{\mR^\mN-\{p\}})$ is smoothly real-compact.
\end{cor}
\begin{proof}
This space is diffeomorphic to $(M,\C_M)$ so it is be smoothly
real-compact.
\end{proof}
\begin{defn}
The disjoint union of differential spaces $(M,\C)$ and $(N,\D)$
where $M\cap N=\emptyset$ is the differential space $(M\cup
N,\C\oplus\D)$. The structure $\C\oplus \D$ is defined by the
property $f\in\C\oplus\D \iff f|_M\in \C$ and $f|_N\in \D$.
\end{defn}
We will prove:
\begin{lem}
If differential spaces $(M,\C)$ and $(N,\D)$ are smoothly
real-compact then the differential space $(M\cup N,\C\oplus \D)$ is
smoothly real-compact.
\end{lem}
\begin{proof}
Elements of the algebra $\C\oplus \D$ are pairs $(f,g)$ where $f\in
\C$ and $g\in \D$. Let $\chi\in \operatorname{Spec}(\C\oplus\D)$. We shall show
that it is evaluation at some point $p\in M\cup N$. From the
equalities $(0,1)+(1,0)=(1,1)$ and $(0,1)(1,0)=(0,0)$ we obtain two
cases:\\
\begin{eqnarray}
1)\quad \chi((1,0))=1 \quad and \quad \chi((0,1))=0\nonumber\\
2)\quad \chi((1,0))=0\quad and \quad \chi((0,1))=1
\nonumber\end{eqnarray} Since every function from $\C\oplus \D$ can
be uniquely decomposed as $(f,g)=(f,0)(1,0)+(0,g)(0,1)$ the
homomorphism $\chi$ acts as
follows:\\
$\chi((f,g))=\chi((f,0))\chi((1,0))+\chi((0,g))\chi((0,1))$. In case
1) we will get \\ $\chi((f,g))=\chi((f,0))$ and in case 2),
$\chi((f,g))=\chi((0,f))$.\\ The algebra of functions of the form
$((f,0))\in \C\oplus\D$ is isomorphic to $\C$. Therefore a
homomorphisms $\psi\in \operatorname{Spec}\C$ can be extended to a homomorphism
from $\C\oplus\D$ by the formula $\bar\psi((f,g))=\psi(f)$. All the
homomorphisms in case 1) are of this form. Therefore in case 1) the
homomorphism $\chi((f,g))=\psi(f)$ where $\psi\in \operatorname{Spec}\C$ is such
that $\bar\psi=\chi$. But since the space $(M,\C)$ is smoothly
real-compact there exists a point $p\in M$ such that $\psi=ev_p$.
Then we can write $\chi((f,g))=ev_p((f,g))=(f,g)(p)=f(p)+g(p)$ for
$p\in M\cup N$. We have shown that in case 1) the homomorphism
$\chi$ is an evaluation. For case 2) the proof is analogous.
\end{proof}
\begin{defn}
We denote by $\tilde{\V}$ the differential structure on $\mR^\mN$
generated by the set $\C_0=\{\pi_i: i\in \mN\}\cup \{\theta_p\}$,
where $\theta_p$ is the characteristic function of the point $p\in
M$.
\end{defn}
\begin{lem}
The differential space $(\mR^\mN,\tilde{\V})$ is smoothly
real-compact.
\end{lem}
\begin{proof}
We can decompose the space $(\mR^\mN,\tilde{\V})$ into the direct
sum of the spaces $(\mR^\mN-\{p\},(\V_\mN)_{\mR^\mN-\{p\}})$ and
$(\{p\}, F(p))$ where $F(p)$ is the algebra of all functions on the
singleton space. From the definition of $(\mR^\mN,\tilde{\V})$ it is
obvious that $\mR^\mN=\{p\}\cup (\mR^\mN-\{p\})$ and
$\tilde{\V}=(\V_\mN)_{\mR^\mN-\{p\}}\oplus F(p)$. Both spaces in the
direct sum are smoothly real-compact so the space
$(\mR^\mN,\tilde{\V})$ is smoothly real-compact.
\end{proof}
Now we present the main results:
\begin{thm}
Any differential subspace of $(\mR^\mN,\V_\mN)$ is smoothly
real-compact.\label{thm:sub}
\end{thm}
\begin{proof}
Let $\iota_M:(M,\C)\rightarrow (\mR^\mN,\V_\mN)$ be the inclusion
mapping. For any $\chi\in \operatorname{Spec}\C$, $\chi\circ\iota_M^*\in
\operatorname{Spec}(\V_\mN)$. The space $(\mR^\mN,\V_\mN)$ is smoothly
real-compact so there is $p\in \mR^\mN$ such that
$\chi\circ\iota_M^*=ev_p|_{\V_\mN}$.
We need to show that $p\in M$. Assume that $p\notin M$. We can treat
the space $(M,\C)$ as a differential subspace of
$(\mR^\mN,\tilde{\V})$. Let $\nu_M:(M,\C)\rightarrow
(\mR^\mN,\tilde{\V})$ be the inclusion. Then $\chi\circ \nu_M^*\in
\operatorname{Spec}\tilde\V$. Because the space $(\mR^\mN,\tilde{\V})$ is smoothly
real-compact there exists a point $q\in \mR^\mN$ such that
$\chi\circ\nu_M^*=ev_q|_{\tilde{\V}}$. We know that on common
generators $\pi_i$ the equalities $\chi(\pi_i|_M)=ev_p(\pi_i)=p_i$
and $\chi(\pi_i|_M)=ev_q(\pi_i)=q_i$ holds for all $i\in \mN$. This
specifies all the coordinates so $p=q$. Therefore we can write
$\chi\circ\nu_M^*=ev_p|_{\tilde{\V}}$. So
$(\chi\circ\nu_M^*)(\theta_p)=ev_p(\theta_p)=1$. We have a
contradiction with the fact that
$(\chi\circ\nu_M^*)(\theta_p)=\chi(\theta_p|_M)=\chi(0)=0$. We see
that $p\in M$ and $\chi\circ \iota_M^*=ev_p|_{\V_\mN}$. So
$\chi(\pi_i|_M)=ev_p(\pi_i|M)$ for all $i\in \mN$. The set
$\{\pi_i:i\in \mN\}$ is the set of generators of the differential
space $(M,\C)$. We derive that $\chi=ev_p$.
\end{proof}
\begin{cor}
Any countably generated differential space is smoothly real-compact.
\end{cor}
\begin{proof}
Every countably generated differential space can be treated as a
subspace of $(\mR^\mN,\V_\mN)$. From Theorem \ref{thm:sub} we know
that all subspaces of this space are smoothly real-compact.
\end{proof}
\section{Conclusion}
We have shown how a real-valued homomorphism act on the algebra of
smooth functions in a differential space. It is sufficient to
examine its value on the generators of the differential structure. A
non-prolongable function on a differential space of sequences
without zero sequence is constructed. From the existence of this
function we deduced that countably generated structural algebra $\C$
of a differential space $(M,\C)$ gives all information about the set
$M$ because the set $\operatorname{Spec}\C$ contains only real valued
homomorphisms of the form $ev_p$ for $p\in M$. Therefore the
geometry of such spaces can be built on their algebras. The choice
of generators of the differential structure is not unique so it is
important to choose the smallest one. It is also shown that the pair
$(\operatorname{Spec}\C, \hat{\C})$ is a differential space for any differential
space $(M,\C)$. For countably generated differential spaces,
$(M,\C)$ and $(\operatorname{Spec}\C,\hat{\C})$ are diffeomorphic. Theorem
\ref{thm:sub} has been obtained as a result of observations about
generators. From \cite{adam} and \cite{kriegl1} it also follows as a
conclusion from a much wider theory. In \cite{ran} there is an
important theorem which gives a sufficient condition for an algebra
of functions on an arbitrary topological space to be smoothly
real-compact. We could easily show that countably generated
differential spaces satisfy this condition. But we have presented
our proof using techniques of differential space theory.
\newpage
|
\part*{Introduction}
The two basic methods to pass parameters to the main()-routine are: input
files and command line arguments. For small scale programs these input methods
allow to change parameters without having to recompile or having to create an
input file parser. Even for large programs input files and command line
arguments are a comfortable way to replace annoying graphical user
interfaces. When debugging complicated code, it is essential to be able to
isolate code fragments into a lonely main()-routine. In order to feed these
isolated code fragments with realistic data, sophisticated command line and
input file parsing becomes indispensable (excerpt from GetPot documentation).
\part{Requirements}
This document has been written to help in the choice of a configuration file
parser library to be included in Verdandi, a scientific library for data
assimilation. The main requirements are
\begin{enumerate}
\item Portability: Verdandi should compile on BSD systems, Linux, MacOS, Unix
and Windows. Beyond the portability itself, this often ensures that most
compilers will accept Verdandi. An obvious consequence is that all
dependencies of Verdandi must be portable, especially the configuration file
parser library.
\item The configuration file parser library should be written in C++, possibly
in C, to be called from Verdandi written in C++.
\item License: any dependency must have a license compatible with Verdandi
licenses (GPL and LGPL).
\item The library must provide the functionality of creating sections in the
configuration file.
\end{enumerate}
\newpage
\part{Config}
\libtab {Config} {1} {May 2008} {May 2008}
{http://www.codeproject.com/KB/files/config-file-parser.aspx} {LGPL (} {C++}
\libdesc
{Only standard C++ (code is cross platform) \\
-- MSVS Express 2005 project files are included \\
-- Manual build successfully tested with GNU g++/Linux}
{}
{-- Configuration files may be sub-structured arbitrarily deep \\
-- Configuration files support the expansion of symbolic values from
previously defined variables and environment variables}
{Sub groups less easy to create and read than with GetPot (requires an
iterator on a STL map)}
\newpage
\part{GetPot}
\libtab {GetPot} {1.1.18} {Jul. 2008} {Mar. 2007} {http://getpot.sourceforge.net/}
{LGPL (} {C++, Java, Python, Ruby}
\libdesc
{Platform independent (istream problem with carriage return/newline on
Microsoft Visual Studio should be solved)}
{-- Easy to install: contained in one single header file \\
-- Requires STL library}
{ -- Parses comand line arguments, single or multiple configuration files \\
-- Variables can be sorted into sections and subsections \\
-- Many features, such as unrecognized object detection (but I found a
bug...) and prefixes allowing to focus on variables in a given section \\
-- Easy to use thanks to detailed, organized and commented example files
(see: \url{http://getpot.sourceforge.net/example.html}) \\
-- Provides a language allowing a variety of operations on variables,
numbers and strings inside a configuration file (recursive replacements,
conglomerate variable names, dictionaries; concatenation and replacement in
strings; power expressions, comparisons and conditions in numeric
expressions (see example file:
\url{http://getpot.sourceforge.net/expand.html})) \\
-- The user can define and name its own variables (the parser finds them) \\
-- GetPot can be used to emulate trivial function calls (without syntax
checking) \\
-- Several styles of comment available for configuration files \\
-- Emacs syntax highlighting for GetPot files \\
-- Provides a Python port \\
-- Non-military usage only, but civil applications are allowed even in a
military context}
{-- No exceptions thrown (for example, no error message when a file is not
found; when numeric expressions have a wrong type, refer to non-existent
variables), a default value is returned if a key is not found in the input
file \\
-- Subsection definition is not intuitive and not flexible if './' syntax is used,
better to use only full subsection names \\
-- Header and source not fully separated}
\newpage
\part{libcfgparse}
\libtab {libcfgparse} {0.6} {Dec. 2005} {Dec. 2005} {http://freshmeat.net/projects/libcfgparse/}
{GPL (} {C}
\libdesc
{}
{Requires yacc and lex}
{-- C, C++ and Python API\\
-- Configuration file syntax is similar to C \\
-- Allows to keep data structured as lists and records; by this allows to
group variables together to handle them conveniently \\
-- Built-in variable types: string, boolean, integer and floating-point \\
-- Allows pre-definitions of types and declaration of default values \\
-- Allows for strict type-checking \\
-- Allows to parse data directly from memory without a file}
{-- Preliminary documentation \\
-- No maintenance, not in development \\
-- Requires yacc and lex (so designed for Unix operating system) \\
-- Written in C (but provides a C++ API)}
\newpage
\part{libconfig}
\libtab {libconfig} {1.3.2} {Feb. 2009} {Feb. 2009} {http://www.hyperrealm.com/libconfig/}
{LGPL (LGPL and } {C++}
\libdesc
{-- Linux, Solaris and Mac OS X \\
-- Windows 2000/XP and later: gcc in the MinGW environment or with Visual
Studio 2005}
{}
{-- A fully reentrant parser: independent configurations can be parsed in
concurrent threads at the same time \\
-- Both C and C++ bindings, as well as hooks to allow for the creation of
wrappers in other languages \\
-- A low-footprint implementation (just 38K for the C library and 66K for
the C++ library) that is suitable for memory-constrained systems \\
-- Proper documentation}
{Configuration file syntax similar to C++ (quite heavy with \{\} and ; )}
\newpage
\part{libConfuse}
\libtab {libConfuse} {2.6} {Dec. 2007} {2004} {http://www.nongnu.org/confuse/}
{LGPL (} {C}
\libdesc
{}
{}
{-- Supports sections and (lists of) values (strings, integers, floats,
booleans or other sections), as well as some other features (such as
single/double-quoted strings, environment variable expansion, functions and
nested include statements)\\
-- No myriads of features in a simple API making it easy to use and quick to
integrate with one's code}
{Written in C}
\newpage
\part{Program\_options}
\libtab {Program\_options} {1.38.0} {?} {Nov. 2007}
{http://www.boost.org/doc/libs/1_38_0/doc/html/program_options.html} {Boost
Software License (} {C++}
\libdesc
{Any platform}
{Program\_options is one of the Boost C++ Libraries}
{-- Reads key/value pairs from command line, a configuration file and
environment variables \\
-- Good documentation and tutorial }
{}
\newpage
\part{RudeConfig™}
\libtab {RudeConfig™} {5.0.5} {released in ?} {?} {http://rudeserver.com/config/}
{GPL (} {C++}
\libdesc
{Borland, Visual C++ 6.0 and Linux}
{}
{-- Independent core library designed for CGI (Common Gateway Interface) \\
-- Reads and writes configuration / .ini files \\
-- Provides fully customizable delimiters and comment characters, ensuring
compatibility with most existing configuration / .ini file formats \\
-- Allows multiline values using backslash escapes \\
-- Comments within the configuration file are fully preserved when the
contents are re-saved \\
-- Deleted values can become commented out when the object is re-saved,
preserving old data \\
-- LGPL license can be provided by the author if needed}
{}
\newpage
\part{Talos}
\libtab {Talos} {1.0} {Apr. 2007} {Oct. 2004} {http://vivienmallet.net/lib/talos/}
{GPL (} {C++}
\libdesc
{Manages Windows end of line character}
{}
{Provides miscellaneous functions and objects, such as \\
-- Functions to deal with strings and files that C++ is missing \\
-- An extended ifstream class \\
-- Two classes to read flexible configuration files: one for a single
configuration file and another for multiple configuration files \\ \\
The parser functionalities are: \\
-- Search for a word in the configuration file \\
-- Extracts the value of a field \\
-- Benefits from markup substitution \\
-- Exceptions may be thrown providing a string explaining what happened \\
-- Allows to apply a constraint on a value, or to force it to be part of a
set of values \\
-- Very clear and flexible configuration file syntax}
{No subsection}
\newpage
\part{Other Libraries}
\paragraph{A Configuration Package} :
\url{http://alumni.media.mit.edu/~rahimi/configuration/}. \limitations {Based
on Lex and Yacc, no activity since 2001.}
\paragraph{ccl}
(Customizable Configuration Library) allows the comment, key/value and string
literal delimiters to be programatically specified at runtime. Simple,
portable with a small interface consisting of five functions :
\url{http://sbooth.org/ccl/}. \limitations {Written in C.}
\paragraph{CFL}
(Configuration File Library). Portable, requires GDSL library:
\url{http://home.gna.org/cfl/}. \limitations {Written in C.}
\paragraph{ConfigFile} C++ portable code with templates:
\url{http://www-personal.umich.edu/~wagnerr/ConfigFile.html}. \limitations
{Impossible to define sections in configuration files.}
\paragraph{Configuration File Parser}
supports both shared as well as static binding of binaries. Provides API for
both manual and automatic processing of configuration file entries :
\url{http://sourceforge.net/projects/parser}. \limitations {Written in C.}
\paragraph{ConfigParser}
reads and writes configuration files with a syntax similar to C/C++; after
being read in a configuration file, the configuration data is available for
access, modification and removal, or can also be written back to a file:
\url{http://www.codedread.com/code.php#Config}. \limitations {Not maintained,
not portable (only Windows executable without source code).}
\paragraph{Confix} :
\url{http://www.flipcode.com/archives/Configuration_File_Parser.shtml}. \limitations
{IBM Public license not compatible with GPL license.}
\paragraph{dot.conf} : \url{http://www.azzit.de/dotconf/}. \limitations {Written in C, no
activity since 2003.}
\paragraph{dotconfpp} : \url{http://ostatic.com/dotconfpp/}. \limitations {Not portable (distribution FreeBSD for i386).}
\paragraph{GLib}
is a general-purpose utility library, which provides many useful data types,
macros, type conversions, string utilities, file utilities, a main loop
abstraction, and so on. It works on many UNIX-like platforms, Windows, OS/2
and BeOS. LGPL license. Provides a key-value file parser:
\url{http://library.gnome.org/devel/glib/unstable/glib-Key-value-file-parser.html}.
\limitations {Written in C.}
\paragraph{iniParser}
Simple, small, portable and robust ini file parsing library.:
\url{http://ndevilla.free.fr/iniparser/}. \limitations {Written in C.}
\paragraph{Libconfig} supports callback functions, automatic variable
assignment : \url{http://www.rkeene.org/oss/libconfig}. \limitations {Written
in C, not tested under Windows.}
\paragraph{libinifile} :
\url{http://it.bmc.uu.se/andlov/proj/libinifile/}. \limitations {Not portable
(uses make to build the library under Unix).}
\paragraph{liblcfg}
is a lightweight configuration file library. The file format supports
arbitrarily nested simple assignments, lists and maps (aka dictionaries):
\url{http://liblcfg.carnivore.it/}. \limitations {Written in C99.}
\paragraph{ParseCfg}
: \url{http://www.mentaljynx.com/CCpp/libs/parsecfg/}. \limitations {Written
in C, no good documentation, no activity since 2001.}
\paragraph{spConfig}
uses configuration files with an XML-like syntax with some additional
preprocessor-style commands:
\url{http://prj.softpixel.com/spconfig/}. \limitations {Written in C, no
activity since 2003.}
\paragraph{Templatized Configuration File Parser (TConf)}
is inspired by the TCLAP project (Templatized C++ Command Line Parser Library:
\url{http://tclap.sourceforge.net/}). TConf is a smaller configuration parser
library and can be used in combination with TCLAP. Released in Aug. 2007 under
GPL license: \url{http://tconf.sourceforge.net/}. \limitations {No sections in
configuration files.}
\paragraph{TCFP}
(Tiny Config File Parser library) is small, fast and simple:
\url{http://sourceforge.net/projects/tcfp/}. \limitations {Pre-alpha version,
written in C99.}
\end{document}
|
\section{Introduction}
In this paper we study Cohomological Hall algebra (COHA) introduced by Kontsevich and Soibelman \cite{KS}, in the case of symmetric quiver
without potential. Our main result is the proof of Kontsevich-Soibelman conjecture on the freeness of COHA of symmetric quiver.
Consider a finite quiver $Q$ with a set of vertices $I$ and with $a_{ij}$ edges from $i\in I$ to $j\in I,$ so that $a_{ij}\in\Z_{\geq 0}.$
One can choose a very degenerate stability condition on the category of complex finite-dimensional representations, so that
stable representations are precisely the simple ones, and they all have the same slope. In particular, each representation is semi-stable with the same slope. Then, for each dimension vector $$\gamma=\{\gamma^i\}_{i\in I}\in\Z_{\geq 0}^I,$$ the moduli space of representations of $Q$ is
an Artin quotient stack $M_{\gamma}/G_{\gamma},$ where $M_{\gamma}$ is an affine space of all representations in coordinate vector spaces $\C^{\gamma^i},$
$G_{\gamma}=\prod\limits_{i\in I}GL(\gamma^i,\C),$ and the action is by conjugation. One then defines a $\Z_{\geq 0}^I$-graded $\Q$-vector space
$\cH$ by the formula
$$\cH=\bigoplus\limits_{\gamma\in \Z_{\geq 0}^I}\cH_{\gamma},\quad \cH_{\gamma}:=H^{\cdot}_{G_{\gamma}}(M_{\gamma,\Q}).$$
Note that originally in \cite{KS}, one takes cohomology with integer coefficients, but we will deal only with the result of tensoring by $\Q.$
Now, for each two vectors $\gamma_1,\gamma_2\in\Z_{\geq 0}^I,$ one has natural correspondence between the stacks $M_{\gamma_1}/G_{\gamma_1}$
and $M_{\gamma_2}/G_{\gamma_2},$ which parameterizes all extensions. We get natural maps of stacks
$$(M_{\gamma_1}/G_{\gamma_1})\times (M_{\gamma_2}/G_{\gamma_2})\leftarrow M_{\gamma_1,\gamma_2}/G_{\gamma_1,\gamma_2}\to M_{\gamma_1+\gamma_2}/G_{\gamma_1+\gamma_2},$$
which allow one to define the multiplication
\begin{equation}\label{shift}H^{\cdot}_{G_{\gamma_1}}(M_{\gamma_1})\otimes H^{\cdot}_{G_{\gamma_2}}(M_{\gamma_2})\to H^{\cdot-2\chi_Q(\gamma_1,\gamma_2)}_{G_{\gamma_1+\gamma_2}}(M_{\gamma_1+\gamma_2}),\end{equation}
where $\chi_Q(\gamma_1,\gamma_2)$ is the Euler form:
$$\chi_Q(\gamma_1,\gamma_2)=\sum\limits_{i\in I}\gamma_1^i\gamma_2^i-\sum\limits_{i,j\in I}a_{ij}\gamma_1^i\gamma_2^j.$$
It is proved in \cite{KS}, Theorem 1, that the resulting product on $\cH$ is associative, so this makes $\cH$ into a $\Z_{\geq 0}^I$-graded algebra,
which is called a (rational) Cohomological Hall algebra of a quiver $Q.$
Now we restrict to the case of symmetric quiver $Q,$ i.e. to the case $a_{ij}=a_{ji}.$ In this case the Euler form $\chi_Q(\gamma_1,\gamma_2)$
is symmetric as well. One defines a $(\Z_{\geq 0}^I\times\Z)$-graded algebra structure on $\cH,$ by assigning to a subspace $H^{k}_{G_{\gamma}}(M_{\gamma})$
a bigrading $(\gamma,k+\chi_Q(\gamma,\gamma)).$ It follows from the formula \eqref{shift} that the product is compatible with this grading. We also define a parity on $\cH$ to be induced by $\Z$-grading.
In general, the algebra $\cH$ for symmetric quiver is not super-commutative, but it becomes such after twisting the product by sign. Denote by $\star$ the resulting
super-commutative product. Our main result is the following Theorem which was conjectured in \cite{KS} (Conjecture 1).
\begin{theo}\label{free_algebra_intro}For any finite symmetric quiver $Q,$ the $(\Z_{\geq 0}^I\times\Z)$-graded algebra $(\cH,\star)$
is a free super-commutative algebra generated by a $(\Z_{\geq 0}^I\times\Z)$-graded vector space $V$
of the form $V=V^{prim}\otimes\Q[x],$ where $x$ is a variable of degree $(0,2)\in\Z_{\geq 0}^I\times\Z,$
and for any $\gamma\in\Z_{\geq 0}^I$ the space $V_{\gamma,k}^{prim}$ is non-zero (and finite-dimensional) only for finitely many $k\in\Z.$\end{theo}
The second result in this paper gives explicit bounds on pairs $(\gamma,k)$ for which $V_{\gamma,k}^{prim}\ne 0.$
For a given symmetric quiver $Q,$ and $\gamma\in \Z_{\geq 0}^I\setminus\{0\},$ we put
$$N_{\gamma}(Q):=\frac12(\sum\limits_{\substack{i,j\in I,\\ i\ne j}}a_{ij}\gamma^i\gamma^j+\sum\limits_{i\in I}\max(a_{ii}-1,0)\gamma^i(\gamma^i-1))-\sum\limits_{i\in I}\gamma^i+2.$$
\begin{theo}\label{upper_bound_intro}In the notation of Theorem \ref{free_algebra_intro}, if $V_{\gamma,k}^{prim}\ne 0,$ then $\gamma\ne 0,$
$$k\equiv \chi_Q(\gamma,\gamma)\text{ mod }2,\quad\text{and}\quad \chi_Q(\gamma,\gamma)\leq k< \chi_Q(\gamma,\gamma)+2N_{\gamma}(Q).$$\end{theo}
The only non-trivial statement in Theorem \ref{upper_bound_intro} is the upper bound on $k.$ In the proofs of both Theorems, we use explicit formulas for the product in $\cH$ from \cite{KS}, Theorem 2. Namely, since the affine space $M_{\gamma}$ is $G_{\gamma}$-equivariantly
contractible, we have
$$\cH_{\gamma}\cong H^{\cdot}(\mrB G_{\gamma}),$$
and the RHS is isomorphic to the algebra of polynomials in $x_{i,\alpha},$ where $i\in I,$ $1\leq \alpha\leq \gamma^i,$
which are invariant with respect to the product of symmetric groups $S_{\gamma^i}.$ Then, given two polynomials $f_1\in \cH_{\gamma_1},$ $f_2\in\cH_{\gamma_2},$ their product $f_1\cdot f_2\in \cH_{\gamma},$ $\gamma=\gamma_1+\gamma_2,$ equals to the sum over all shuffles (for any $i\in I$) of the following rational function in variables $(x_{i,\alpha}')_{i\in I,\alpha\in \{1,\dots,\gamma_1^i\}},$
$(x_{i,\alpha}'')_{i\in I,\alpha\in \{1,\dots,\gamma_2^i\}}:$
$$f_1((x_{i,\alpha}'))f_2((x_{i,\alpha}''))
\frac{\prod\limits_{i,j\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=1}^{\gamma_2^j}(x_{j,\alpha_2}''-x_{i,\alpha_1}')^{a_{ij}}}
{\prod\limits_{i\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=1}^{\gamma_2^i}(x_{i,\alpha_2}''-x_{i,\alpha_1}')}.$$
Theorems \ref{free_algebra_intro} and \ref{upper_bound_intro} imply the corresponding results for the generating functions for Cohomological Hall algebras,
in particular, positivity for quantum Donaldson-Thomas invariants. The positivity result was used by S. Mozgovoy to prove
Kaz's conjecture for quivers with at least one loop at each vertex \cite{M}
The paper is organized as follows.
Section \ref{s:preliminaries} is devoted to some preliminaries on Cohomological Hall algebras for quivers. We follow \cite{KS}, Section 2.
In Subsection \ref{ss:COHA_definition} we give a definition of rational Cohomological Hall algebra for an arbitrary finite quiver.
Subsection \ref{ss:explicit} is devoted to explicit formulas for the product in Cohomological Hall algebras.
In Subsection \ref{ss:grading} we define an additional $\Z$-grading on COHA of symmetric quiver, so that we get a $(\Z_{\geq 0}^I\times\Z)$-graded
algebra. Then, we show how to modify a product on $\cH$ by a sign to get a super-commutative algebra $(\cH,\star),$ with parity induced
by $\Z$-grading.
Section \ref{s:main_results} is devoted to the proofs of Theorem \ref{free_algebra_intro} (Theorem \ref{free_algebra}) and
Theorem \ref{upper_bound_intro} (Theorem \ref{upper_bound}).
In Section \ref{s:DT_invariants} we discuss applications of our results to the generating function of COHA, or, in other words, to quantized Donaldson-Thomas invariants.
\smallskip
{\noindent {\bf Acknowledgements.}} I am grateful to Maxim Kontsevich and Yan Soibelman for useful discussions. I am also
grateful to Sergey Mozgovoy for pointing my attention to his paper on Kac's conjecture \cite{M}.
\section{Preliminaries on Cohomological Hall algebras}
\label{s:preliminaries}
In this Section we recall some definitions and results from \cite{KS}, Section 2.
\subsection{COHA of a quiver}
\label{ss:COHA_definition}
Let $Q$ be a finite quiver. Denote its set of vertices by $I,$ and let $a_{ij}\in\Z_{\geq 0}$ be the number of arrows from $i$ to $j,$
where $i,j\in I.$ Fix a dimension vector $\gamma=(\gamma^i)_{i\in I}\in\Z_{\geq 0}^I.$ We have an affine variety of representations of $Q$
in complex coordinate vector spaces $\C^{\gamma^i}:$
$$M_{\gamma}=\prod\limits_{i,j\in I}\C^{a_{ij}\gamma^i\gamma^j}.$$
The variety $M_{\gamma}$ is acted via conjugation by the complex algebraic group
$G_{\gamma}=\prod\limits_{i\in I}GL(\gamma^i,\C).$
Recall that infinite-dimensional Grammanian
$$\Gr(d,\infty)=\lim\limits_{\to}Gr(d,\C^n),\quad n\to+\infty,$$
is a model for the classifying space of $GL(d,\C).$ Put
$$\mrB G_{\gamma}:=\prod\limits_{i\in I}\mrB GL(\gamma^i,\C)=\prod\limits_{i\in I}\Gr(\gamma^i,\infty).$$
We have a standard universal $G_{\gamma}$-bundle $\mrE G_{\gamma}\to\mrB G_{\gamma},$ and the Artin stack $M_{\gamma}/G_{\gamma}$ gives a universal
family over $\mrB G_{\gamma}:$
$$M_{\gamma}^{univ}:=(\mrE G_{\gamma}\times M_{\gamma})/G_{\gamma}\to \mrE G_{\gamma}/G_{\gamma}=\mrB G_{\gamma}.$$
Define a $\Z_{\geq 0}^I$-graded $\Q$-vector space
$$\cH=\bigoplus_{\gamma\in\Z_{\geq 0}^I}\cH_{\gamma},$$
putting
$$\cH_{\gamma}:=H^{\cdot}_{G_{\gamma}}(M_{\gamma},\Q)=\bigoplus_{n\geq 0}H^n(M_{\gamma}^{univ},\Q).$$
Now we define a multiplication on $\cH$ which makes it into associative unital $\Z_{\geq 0}^I$-graded algebra over $\Q.$
Take two vectors $\gamma_1,\gamma_2\in \Z_{\geq 0}^I,$ and put $\gamma:=\gamma_1+\gamma_2.$ Consider the affine subspace $M_{\gamma_1,\gamma_2}\subset M_{\gamma},$ which consists of representations for which standard subspaces $\C^{\gamma_1^i}\subset \C^{\gamma^i}$ form a
subrepresentation. The subspace $M_{\gamma_1,\gamma_2}$ is preserved by the action of the subgroup $G_{\gamma_1,\gamma_2}\subset G_{\gamma}$
which consists of transformations preserving subspaces $\C^{\gamma_1^i}\subset\C^{\gamma^i}.$ We use a model for $\mrB G_{\gamma_1,\gamma_2}$
which is the total space of a bundle over $\mrB G_{\gamma}$ with fiber $G_{\gamma}/G_{\gamma_1,\gamma_2}$ (i.e. a product of infinite-dimensional partial flag varieties $\Fl(\gamma_1^i,\gamma_i,\infty)$). We have natural projection $\mrE G_{\gamma}\to \mrB G_{\gamma_1,\gamma_2}$ which is a universal
$G_{\gamma_1,\gamma_2}$-bundle.
Now define the morphism
$$m_{\gamma_1,\gamma_2}:\cH_{\gamma_1}\otimes\cH_{\gamma_2}\to \cH_{\gamma}$$
as the composition of the K\"unneth isomorphism
$$\otimes:H_{G_{\gamma_1}}^{\cdot}(M_{\gamma_1,\Q})\otimes H_{G_{\gamma_2}}^{\cdot}(M_2,\Q)\stackrel{\cong}{\to} H_{G_{\gamma_1\times G_{\gamma_2}}}^{\cdot}(M_{\gamma_1}\times M_{\gamma_2},\Q)$$
and the following morphisms:
$$H_{G_{\gamma_1\times G_{\gamma_2}}}^{\cdot}(M_{\gamma_1}\times M_{\gamma_2},\Q)\stackrel{\cong}{\to}H_{G_{\gamma_1,G_{\gamma_2}}}^{\cdot}
(M_{\gamma_1,\gamma_2},\Q)\to H^{\cdot+2c_1}_{G_{\gamma_1,\gamma_2}}(M_{\gamma},\Q)\to H^{\cdot+2c_1+2c_2}_{G_{\gamma}}(M_{\gamma}).$$
Here the first map is induced by natural surjective homotopy equivalences $$M_{\gamma_1,\gamma_2}\stackrel{\sim}{\to}M_{\gamma_1}\times M_{\gamma_2}, G_{\gamma_1,\gamma_2}\to G_{\gamma_1}\times G_{\gamma_2}.$$
The other two maps are natural pushforward morphisms, with
$$c_1=\dim_{\C} M_{\gamma}-\dim_{\C} M_{\gamma_1,\gamma_2},\quad c_2=\dim_{\C}G_{\gamma_1,\gamma_2}-\dim_{\C}G_{\gamma}.$$
\begin{theo}(\cite{KS}, Theorem 1) The constructed product $m$ on $\cH$ is associative.\end{theo}
Note that \begin{equation}\label{Euler}c_1+c_2=-\chi_Q(\gamma_1,\gamma_2),\end{equation} where
$$\chi_Q(\gamma_1,\gamma_2)=\sum\limits_{i\in I}\gamma_1^i\gamma_2^i-\sum\limits_{i,j\in I}a_{ij}\gamma_1^i\gamma_2^j$$
is the Euler form of the quiver $Q.$ That is, given two representations $R_1,R_2$ (over any field) of the quiver $Q,$ with dimension vectors $\gamma_1,\gamma_2$ respectively, one has
$$\sum\limits_i(-1)^i\dim\Ext^i(R_1,R_2)=\dim\Hom(R_1,R_2)-\dim\Ext^1(R_1,R_2)=\chi_Q(\gamma_1,\gamma_2).$$
\subsection{Explicit description of COHA of a quiver}
\label{ss:explicit}
Since the affine spaces $M_{\gamma}$ are $G_{\gamma}$-equivariantly contractible, we have natural isomorphisms
$$\cH_{\gamma}\cong H^{\cdot}(\mrB G_{\gamma},\Q)=\bigotimes_{i\in I}H^{\cdot}(\mrB GL(\gamma^i,\C),\Q).$$
Recall that
$$H^{\cdot}(\mrB GL(d,\C),\Q)\cong \Q[x_1,\dots,x_d]^{S_d}.$$
For a vector $\gamma\in\Z_{\geq 0}^{I},$ introduce variables $x_{i,\alpha},$ where $i\in I,$ $\alpha\in \{1,\dots,\gamma^i\}.$
Then, we get natural isomorphisms
$$\cH_{\gamma}\cong\Q[\{x_{i,\alpha}\}_{i\in I, \alpha\in \{1,\dots,\gamma^i\}}]^{\prod\limits_{i\in I}S_{\gamma^i}}.$$
From this moment, we identify the elements of $\cH_{\gamma}$ with the corresponding polynomials.
\begin{theo}\label{explicit}(\cite{KS}, Theorem 2)
Given two polynomials $f_1\in \cH_{\gamma_1},$ $f_2\in\cH_{\gamma_2},$ their product $f_1\cdot f_2\in \cH_{\gamma},$ $\gamma=\gamma_1+\gamma_2,$ equals to the sum over all shuffles (for any $i\in I$) of the following rational function in variables $(x_{i,\alpha}')_{i\in I,\alpha\in \{1,\dots,\gamma_1^i\}},$
$(x_{i,\alpha}'')_{i\in I,\alpha\in \{1,\dots,\gamma_2^i\}}:$
$$f_1((x_{i,\alpha}'))f_2((x_{i,\alpha}''))
\frac{\prod\limits_{i,j\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=1}^{\gamma_2^j}(x_{j,\alpha_2}''-x_{i,\alpha_1}')^{a_{ij}}}
{\prod\limits_{i\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=1}^{\gamma_2^i}(x_{i,\alpha_2}''-x_{i,\alpha_1}')}.$$
\end{theo}
\subsection{Additional grading in the symmetric case}
\label{ss:grading}
Now assume that the quiver $Q$ is symmetric, i.e. $a_{ij}=a_{ji},$ $i,j\in I.$ Then the Euler form
$$\chi_Q(\gamma_1,\gamma_2)=\sum\limits_{i\in I}\gamma_1^i\gamma_2^i-\sum\limits_{i,j\in I}a_{ij}\gamma_1^i\gamma_2^i$$
is symmetric as well.
We make $\cH$ into a $(\Z_{\geq 0}^I\times\Z)$-graded algebra as follows. For a polynomial $f\in \cH_{\gamma}$ of degree $k$
we define its bigrading to be $(\gamma,2k+\chi_Q(\gamma,\gamma)).$ It follows from either \eqref{Euler} or Theorem
\ref{explicit} that the product on $\cH$ is compatible with this bigrading.
Define the super-structure on $\cH$ to be induced by $\Z$-grading.
For two elements $a_{\gamma,k}\in\cH_{\gamma,k},$ $a_{\gamma',k'}\in\cH_{\gamma',k'},$ we have
$$a_{\gamma,k}a_{\gamma',k'}=(-1)^{\chi_Q(\gamma,\gamma')}a_{\gamma',k'}a_{\gamma,k}.$$
In general, this does not mean that $\cH$ is super-commutative. However, it is easy to twist the product by a sign, so that $\cH$ becomes super-commutative.
This can be done as follows.
Define the homomorphism of abelian groups $\epsilon:\Z^I\to \Z/2\Z$ by the formula
$$\epsilon(\gamma)=\chi_Q(\gamma,\gamma)\text{ mod }2.$$
Note that the parity of the element $a_{\gamma,k}$ equals to $\epsilon(\gamma)$ (by the definition).
We have a bilinear form
$$\Z^I\times\Z^I\to\Z/2\Z,\quad (\gamma_1,\gamma_2)\mapsto (\chi_Q(\gamma_1,\gamma_2)+\epsilon(\gamma_1)\epsilon(\gamma_2))\text{ mod }2,$$
which induces a symmetric form $\beta$ on the space $(\Z/2\Z)^I,$ such that $\beta(\gamma,\gamma)=0$ for all $\gamma\in (\Z/2\Z)^I.$
Hence, there exists a bilinear form $\psi$ on $(\Z/2\Z)^I$ such that
$$\psi(\gamma_1,\gamma_2)+\psi(\gamma_2,\gamma_1)=\beta(\gamma_1,\gamma_2).$$
Then the twisted product on $\cH$ is defined by the formula
$$a_{\gamma,k}\star a_{\gamma',k'}=(-1)^{\psi(\gamma,\gamma')}a_{\gamma,k}\cdot a_{\gamma',k'}.$$
It follows from the definition that the product $\star$ is associative, and the algebra $(\cH,\star)$ is super-commutative. From now on, we fix the choice
of bilinear form $\psi,$ and the corresponding product $\star$ on $\cH.$
\section{Freeness of COHA of a symmetric quiver}
\label{s:main_results}
\begin{theo}\label{free_algebra}For any finite symmetric quiver $Q,$ the $(\Z_{\geq 0}^I\times\Z)$-graded algebra $(\cH,\star)$
is a free super-commutative algebra generated by a $(\Z_{\geq 0}^I\times\Z)$-graded vector space $V$
of the form $V=V^{prim}\otimes\Q[x],$ where $x$ is a variable of bidegree $(0,2)\in\Z_{\geq 0}^I\times\Z,$
and for any $\gamma\in\Z_{\geq 0}^I$ the space $V_{\gamma,k}^{prim}$ is non-zero (and finite-dimensional) only for finitely many $k\in\Z.$\end{theo}
\begin{proof}Our first step is to construct the space $V.$ It will be convenient to treat $\cH_{\gamma}$ itself as a $\Z$-graded algebra (with the usual multiplication of polynomials, and the standard even grading). To distinguish the product in $\cH_{\gamma}$ and the product in $\cH,$ we will always
denote the last product by $"\star "$.
For convenience, we put $A_{\gamma}:=\Q[\{x_{i,\alpha}\}_{i\in I,1\leq \alpha\leq \gamma^i}]$ (considered as a $\Z$-graded algebra) and $P_{\gamma}:=\prod\limits_{i\in I}S_{\gamma^i}.$ Then
we have that $\cH_{\gamma}=A_{\gamma}^{P_{\gamma}}.$ Further, put
$$A_{\gamma}^{prim}:=\Q[(x_{j,\alpha_2}-x_{i,\alpha_1})_{i,j\in I,1\leq \alpha_1\leq \gamma^i,1\leq\alpha_2\leq \gamma^j}],\quad
\sigma_{\gamma}:=\sum\limits_{\substack{i\in I,\\ 1\leq \alpha\leq\gamma^i}}x_{i,\alpha}\in A_{\gamma}.$$
Then $A_{\gamma}=A_{\gamma}^{prim}\otimes\Q[\sigma_{\gamma}].$ Further, we have
$$\cH_{\gamma}=\cH_{\gamma}^{prim}\otimes\Q[\sigma_{\gamma}],\quad\cH_{\gamma}^{prim}:=(A_{\gamma}^{prim})^{P_{\gamma}}.$$
Now, for each $\gamma\in\Z_{\geq 0}^I,$ denote by $J_{\gamma}$ the smallest $P_{\gamma}$-stable $A_{\gamma}^{prim}$-submodule of the localization
$A_{\gamma}^{prim}[(x_{i,\alpha_2}-x_{i,\alpha_1})^{-1}_{i\in I,1\leq \alpha_1<\alpha_2\leq \gamma^i}],$ such that for all decompositions $\gamma=\gamma_1+\gamma_2,$ $\gamma_1,\gamma_2\in\Z_{\geq 0}^I\setminus \{0\},$ we have that
$$\frac{\prod\limits_{i,j\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=\gamma_1^j+1}^{\gamma^j}(x_{j,\alpha_2}-x_{i,\alpha_1})^{a_{ij}}}
{\prod\limits_{i\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=\gamma_1^i+1}^{\gamma^i}(x_{i,\alpha_2}-x_{i,\alpha_1})}\in J_{\gamma}.$$
Clearly, $J_{\gamma}^{P_{\gamma}}\subset\cH_{\gamma}^{prim}.$ Define $V_{\gamma}^{prim}\subset\cH_{\gamma}^{prim}$ to be a graded subspace
such that $$\cH_{\gamma}^{prim}=V_{\gamma}^{prim}\oplus J_{\gamma}^{P_{\gamma}}.$$
Further, put $$V_{\gamma}:=V_{\gamma}^{prim}\otimes\Q[\sigma_{\gamma}]\subset \cH_{\gamma},\quad V:=\bigoplus\limits_{\gamma\in\Z_{\geq 0}^I}V_{\gamma}.$$
We will prove that $V$ freely generates $\cH,$ and that all the spaces $V_{\gamma}^{prim}$ are finite-dimensional (this would imply the Theorem).
\begin{lemma}The subspace $V\subset\cH$ generates $\cH$ as an algebra.\end{lemma}
\begin{proof}Indeed, for each $\gamma\in\Z_{\geq 0}^I,$ the image of the multiplication map
$$\bigoplus\limits_{\substack{\gamma_1+\gamma_2=\gamma,\\ \gamma_1,\gamma_2\in\Z_{\geq 0}^I\setminus\{0\}}}\cH_{\gamma_1}\otimes\cH_{\gamma_2}\to\cH_{\gamma}$$
is precisely $J_{\gamma}^{P_{\gamma}}\otimes\Q[\sigma_{\gamma}]$ (this is straightforward to check). Hence, it follows by induction on $\sum\limits_{i\in I}\gamma^i$ that the subspace $\cH_{\gamma}$ is contained in the subalgebra generated by $V.$ This proves Lemma.\end{proof}
Now we will show that the spaces $V_{\gamma}^{prim}$ are finite-dimensional.
\begin{lemma}For each $\gamma\in\Z_{\geq 0}^I,$ the space $V_{\gamma}^{prim}$ is finite-dimensional.\end{lemma}
\begin{proof}In other words, we need to show that the ideal $J_{\gamma}^{P_{\gamma}}\subset\cH_{\gamma}^{prim}$
has finite codimension.
First note that if we replace $a_{ii}$ by $a_{ii}+1,$ then the fractional ideal $J_{\gamma}$ would become smaller or equal. Hence, we may and will assume,
that $a_{ii}>0$ for $i\in I,$ and so $J_{\gamma}\subset A_{\gamma}^{prim}.$
Since we have natural injective morphisms
$$\cH_{\gamma}^{prim}/J_{\gamma}^{P_{\gamma}}\hookrightarrow A_{\gamma}^{prim}/J_{\gamma},$$ it suffices to show that the ideal $J_{\gamma}\subset A_{\gamma}^{prim}$
has finite codimension. It will be convenient to treat the algebra $A_{\gamma}^{prim}$ as the algebra of functions
on the hyperplane $W\subset \A_{\Q}^{\sum\limits_{i\in I}\gamma^i},$ given by equation $\sigma_{\gamma}(x)=0.$
It suffices to show that
$$\supp(A_{\gamma}^{prim}/J_{\gamma})=\{0\}\subset W.$$
Assume the converse is true. Then there exists a point $y\in W_{\bar{\Q}},$ $y\ne 0,$ such that all the functions from $J_{\gamma}$
vanish at $y.$ Since $\sigma_{\gamma}(y)=0,$ we have that not all of the coordinates $y_{i,\alpha}$ are equal to each other. Since the ideal $J_{\gamma}$ is $P_{\gamma}$-stable, we may assume that there exists a decomposition $\gamma=\gamma_1+\gamma_2,$ $\gamma_1,\gamma_2\in\Z_{\geq 0}^I\setminus \{0\},$
such that
$$y_{i,\alpha_1}\ne y_{j,\alpha_2}\text{ for }1\leq \alpha_1\leq\gamma_1^i,\, \gamma_1^j+1\leq \alpha_2\leq \gamma^j.$$
But then the function
$$\frac{\prod\limits_{i,j\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=\gamma_1^j+1}^{\gamma^j}(x_{j,\alpha_2}-x_{i,\alpha_1})^{a_{ij}}}
{\prod\limits_{i\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=\gamma_1^i+1}^{\gamma^i}(x_{i,\alpha_2}-x_{i,\alpha_1})}\in J_{\gamma}$$
does not vanish at $y,$ a contradiction.
Lemma is proved.
\end{proof}
It remains to prove the freeness.
\begin{lemma}The subspace $V\subset \cH$ freely generates $\cH.$\end{lemma}
\begin{proof}We have already shown the generation. So we need to show freeness.
Choose an order on $I,$ and fix the corresponding lexicographical order on $\Z_{\geq 0}^I$ (denoted by $\gamma\succeq\gamma'$).
Further, denote by $e_{\gamma,\beta},$ $1\leq \beta\leq \dim V_{\gamma}^{prim},$ a homogeneous basis of $V_{\gamma}^{prim}.$
We have the lexicographical order on all of the elements $e_{\gamma,\beta}$ (for all $\gamma$ and $\beta$). Further, the elements $e_{\gamma,\beta}\sigma_{\gamma}^m$ (for all $\gamma,\beta,m$) form the basis of $V,$ and again we have a lexicographical order on them, which we
denote by $\succeq.$
Fix some $\gamma\in\Z_{\geq 0}^I.$ Consider the set $Seq_{\gamma}$ of all {\it nonincreasing} sequences $(e_{\gamma_1,\beta_1}\sigma_{\gamma_1}^{m_1},\dots,e_{\gamma_d,\beta_d}\sigma_{\gamma_d}^{m_d})$ such that
1) $\gamma_1+\dots+\gamma_d=\gamma;$
2) an equality $(\gamma_i,\beta_i,m_i)=(\gamma_{i+1},\beta_{i+1},m_{i+1})$ implies $\epsilon(\gamma_i)=0.$
Clearly, we have natural lexicographical order on $Seq_{\gamma}$ (which we again denote by $\succeq$). For a sequence $t\in Seq_{\gamma},$
we denote by $M_t\in\cH_{\gamma}$ the corresponding product.
What we need is to show non-vanishing of each non-trivial linear combination:
\begin{equation}\label{lin_comb}T=\sum\limits_{i=1}^n\lambda_iM_{t_i}\ne 0,\quad t_1,\dots,t_n\in Seq_{\gamma},\, t_1\succ\dots\succ t_n,\, \lambda_1\dots\lambda_n\ne 0.\end{equation}
Fix some $t_i$ and $\lambda_i$ as in \eqref{lin_comb}. Denote by $(\gamma_1,\dots,\gamma_k)$ the underlying sequence of elements in $\Z_{\geq 0}^I$ for the sequence $t_1\in Seq_{\gamma}.$ Then $\gamma_1+\dots+\gamma_k=\gamma,$ and $\gamma_i\ne 0.$ We have natural isomorphism
$$A_{\gamma}\cong A_{\gamma_1}\otimes\dots\otimes A_{\gamma_k}=:\widetilde{A_{\gamma}},$$ which induces an inclusion
$$\iota:\cH_{\gamma}\hookrightarrow \cH_{\gamma_1}\otimes\dots\otimes\cH_{\gamma_k}=:\widetilde{\cH_{\gamma}}.$$
Put $\widetilde{P_{\gamma}}:=P_{\gamma_1}\times\dots\times P_{\gamma_k}.$ Then we have $\widetilde{\cH_{\gamma}}=\widetilde{A_{\gamma}}^{\widetilde{P_{\gamma}}}.$ Further, take an ideal
$$(J_{\gamma_1} \cap A_{\gamma_1}^{prim}) \widetilde{A_{\gamma}}+\dots+(J_{\gamma_k}\cap A_{\gamma_k}^{prim})\widetilde{A_{\gamma}}=:\widetilde{J_{\gamma}}\subset \widetilde{A_{\gamma}}.$$
We will write $x^{(p)}_{i,\alpha}\in\widetilde{A_{\gamma}}$ for variables from the $p$-th factor $A_{\gamma_p}\subset \widetilde{A_{\gamma}}.$
{\noindent {\bf Claim.}} {\it The elements $(x^{(q)}_{j,\alpha_2}-x^{(p)}_{i,\alpha_1})\in \widetilde{A_{\gamma}},$
$1\leq p<q\leq k,$ are not zero divisors in the quotient ring
$$\widetilde{A_{\gamma}}/\widetilde{J_{\gamma}}.$$}
\begin{proof}For convenience, we may assume that the sequence $\gamma_1,\dots,\gamma_k$ is not necessarily non-increasing, and $q=k.$
Any element $g\in \widetilde{A_{\gamma}}$ can be written (in a unique way) as a sum
$$g=\sum\limits_{\nu=0}^N g_{\nu}\sigma_{\gamma_k}^{\nu},\quad g_{\nu}\in A_{\gamma_1}\otimes\dots\otimes A_{\gamma_{k-1}}\otimes A_{\gamma_k}^{prim}.$$
The following are obviously equivalent:
$(i)$ $g\not\in \widetilde{J_{\gamma}};$
$(ii)$ for some $\nu\in\{0,\dots,N\},$ $g_{\nu}\not\in\widetilde{J_{\gamma}}.$
Now suppose that $g\not\in\widetilde{J_{\gamma}}.$ We need to show that
\begin{equation}\label{not_in_ideal2}(x^{(k)}_{j,\alpha_2}-x^{(p)}_{i,\alpha_1})g\not\in\widetilde{J_{\gamma}}.\end{equation}
We may assume that $g_N\not\in \widetilde{J_{\gamma}}.$ Put
$$x^{(k)}_{av}:=\frac1{\sum\limits_{i\in I}\gamma_k^i}\sum\limits_{i,\alpha}x^{(k)}_{i,\alpha}=\frac1{\sum\limits_{i\in I}\gamma_k^i}\sigma_{\gamma_k}.$$
Then $x^{(k)}_{j,\alpha_2}-x^{(k)}_{av}\in A_{\gamma_k}^{prim},$ and we have
$$(x^{(k)}_{j,\alpha_2}-x^{(p)}_{i,\alpha_1})g=(x^{(k)}_{j,\alpha_2}-x^{(k)}_{av}-x^{(p)}_{i,\alpha_1})g+x^{(k)}_{av}g=\frac1{\sum\limits_{i\in I}\gamma_k^i}g_N\sigma_{\gamma_k}^{N+1}+
\sum\limits_{\nu=0}^Ng_{\nu}'\sigma_{\gamma_k}^{\nu}$$
for some $g_{\nu}'\in A_{\gamma_1}\otimes\dots\otimes A_{\gamma_{k-1}}\otimes A_{\gamma_k}^{prim}.$
Since $\frac1{\sum\limits_{i\in I}\gamma_k^i}g_N\not\in\widetilde{J_{\gamma}}$ by our assumption, this implies \eqref{not_in_ideal2}. Claim is proved.
\end{proof}
We put
$$\widetilde{A_{\gamma}}':=\widetilde{A_{\gamma}}[(x^{(q)}_{j,\alpha_2}-x^{(p)}_{i,\alpha_1})^{-1}_{1\leq p<q\leq k}],\quad \widetilde{\cH_{\gamma}}':=
(\widetilde{A_{\gamma}}')^{\widetilde{P_{\gamma}}}.$$
We denote by the same letter $L$ the localization maps $L:\widetilde{A_{\gamma}}\to \widetilde{A_{\gamma}}',$ $L:\widetilde{\cH_{\gamma}}\to \widetilde{\cH_{\gamma}}'.$ Also put $\widetilde{J_{\gamma}}':=\widetilde{A_{\gamma}}'L(\widetilde{J_{\gamma}}).$
It follows directly from Claim that the induced maps
\begin{equation}\label{injective}L:\widetilde{A_{\gamma}}/\widetilde{J_{\gamma}}\to \widetilde{A_{\gamma}}'/\widetilde{J_{\gamma}}',\quad
L:\widetilde{\cH_{\gamma}}/(\widetilde{J_{\gamma}})^{\widetilde{P_{\gamma}}}\to \widetilde{\cH_{\gamma}}'/(\widetilde{J_{\gamma}}')^P\end{equation}
are injective.
Now, let $r\in\{1,\dots,n\}$ be the maximal number such that the underlying sequence of elements in $\Z_{\geq 0}^I$ for $t_r$
coincides with $(\gamma_1,\dots,\gamma_k).$ Then it is straightforward to check that
$$L\iota(M_{t_l})\in (\widetilde{J_{\gamma}}')^{\widetilde{P_{\gamma}}}\text{ for }r+1\leq l\leq n.$$
Thus, it suffices to show that
\begin{equation}\label{not_in_ideal}L\iota(\sum\limits_{i=1}^r\lambda_iM_{t_i})\not\in(\widetilde{J_{\gamma}}')^{\widetilde{P_{\gamma}}}.\end{equation}
For all relevant $\beta_i,m_i$ we have the following comparison:
\begin{multline}\label{ideal1}L\iota(e_{\gamma_1,\beta_1}\sigma_{\gamma_1}^{m_1}\star\dots\star e_{\gamma_k,\beta_k}\sigma_{\gamma_k}^{m_k})\equiv\\
F_{\gamma_1,\dots,\gamma_k}\cdot \sum\limits_{\tau}s(\tau)e_{\gamma_{1},\beta_{\tau(1)}}\sigma_{\gamma_{1}}^{m_{\tau(1)}}\otimes\dots\otimes e_{\gamma_{k},\beta_{\tau(k)}}\sigma_{\gamma_{k}}^{m_{\tau(k)}}\text{ mod }(\widetilde{J_{\gamma}}')^{\widetilde{P_{\gamma}}},\end{multline}
where the sum is taken over all permutations $\tau\in S_k$ such that $\gamma_p=\gamma_{\tau(p)}$ for all $p\in\{1,\dots,k\},$
and $s(\tau)$ is the Koszul sign (recall that the parity of $e_{\gamma,\beta}\sigma_{\gamma}^k$ equals to $\epsilon(\gamma)$),
and $F_{\gamma_1,\dots,\gamma_k}\in\widetilde{\cH_{\gamma}}'$ is (up to sign) the product of some powers (positive and ($-1$)-st) of the differences
$$(x^{(q)}_{j,\alpha_2}-x^{(p)}_{i,\alpha_1})\in \widetilde{A_{\gamma}},\quad 1\leq p<q\leq k.$$
Thus, $F_{\gamma_1,\dots,\gamma_k}$ is invertible, and according to \eqref{ideal1} and injectivity of maps \eqref{injective}, we are left to check that
$$\sum\limits_{\tau}s(\tau)e_{\gamma_{1},\beta_{\tau(1)}}\sigma_{\gamma_{1}}^{m_{\tau(1)}}\otimes\dots\otimes e_{\gamma_{k},\beta_{\tau(k)}}\sigma_{\gamma_{k}}^{m_{\tau(k)}}\not\in \widetilde{J_{\gamma}}^{\widetilde{P_{\gamma}}}.$$
But this follows from the condition 2) in the above definition of the set of sequences $Seq_{\gamma},$ and from the definition of $e_{\gamma_i,\beta}.$
This proves \eqref{not_in_ideal2}, hence the desired linear independence \eqref{lin_comb}, and hence free generation. Lemma is proved.
\end{proof}
Theorem is proved.
\end{proof}
It is clear that if $V_{\gamma,k}^{prim}\ne 0$ in the notation of the above Theorem, then $k\equiv \chi_Q(\gamma,\gamma)\text{ mod }2$ and $k\geq
\chi_Q(\gamma,\gamma).$ Our next result is an upper bound on $k$ (depending on $\gamma$) for which $V_{\gamma,k}\ne 0.$
For a given symmetric quiver $Q$ and $\gamma\in\Z_{\geq 0}^I\setminus\{0\},$ we put
$$N_{\gamma}(Q):=\frac12(\sum\limits_{\substack{i,j\in I,\\ i\ne j}}a_{ij}\gamma^i\gamma^j+\sum\limits_{i\in I}\max(a_{ii}-1,0)\gamma^i(\gamma^i-1))-\sum\limits_{i\in I}\gamma^i+2.$$
\begin{theo}\label{upper_bound}In the notation of Theorem \ref{free_algebra}, if $V_{\gamma,k}^{prim}\ne 0,$ then $\gamma\ne 0,$
$$k\equiv \chi_Q(\gamma,\gamma)\text{ mod }2,\quad\text{and}\quad \chi_Q(\gamma,\gamma)\leq k< \chi_Q(\gamma,\gamma)+2N_{\gamma}(Q).$$\end{theo}
\begin{proof}According to the proof of Theorem \ref{free_algebra}, we have \begin{equation}\label{equal_dim}\dim V_{\gamma,k}^{prim}=\dim (\cH_{\gamma}^{prim}/J_{\gamma}^{P_{\gamma}})^{k-\chi_Q(\gamma,\gamma)}.\end{equation}
Recall that $P_{\gamma}=\prod\limits_{i\in I}S_{\gamma^i},$
$$A_{\gamma}^{prim}:=\Q[(x_{j,\alpha_2}-x_{i,\alpha_1})_{i,j\in I,1\leq \alpha_1\leq \gamma^i,1\leq\alpha_2\leq \gamma^j}],\quad
\cH_{\gamma}^{prim}:=(A_{\gamma}^{prim})^{P_{\gamma}},$$
and $J_{\gamma}$ is the smallest $P_{\gamma}$-stable $A_{\gamma}^{prim}$-submodule of the localization
$$A_{\gamma}^{prim}[(x_{i,\alpha_2}-x_{i,\alpha_1})^{-1}_{i\in I,1\leq \alpha_1<\alpha_2\leq \gamma^i}],$$ such that for all decompositions $\gamma=\gamma_1+\gamma_2,$ $\gamma_1,\gamma_2\in\Z_{\geq 0}^I\setminus \{0\},$ we have that
\begin{equation}\label{expr}\frac{\prod\limits_{i,j\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=\gamma_1^j+1}^{\gamma^j}(x_{j,\alpha_2}-x_{i,\alpha_1})^{a_{ij}}}
{\prod\limits_{i\in I}\prod\limits_{\alpha_1=1}^{\gamma_1^i}\prod\limits_{\alpha_2=\gamma_1^i+1}^{\gamma^i}(x_{i,\alpha_2}-x_{i,\alpha_1})}\in J_{\gamma}.\end{equation}
Recall that we take the standard even grading on $A_{\gamma}^{prim}$ with $\deg(x_{j,\alpha_2}-x_{i,\alpha_1})=2,$ and the induced grading on $\cH_{\gamma}^{prim}.$
According to \eqref{equal_dim}, it suffices to prove inclusions
\begin{equation}\label{ideal_generates}(A_{\gamma}^{prim})^d\subset J_{\gamma}\text{ for } d\geq 2N(Q).\end{equation}
For any $i,j\in I,$ put $$a_{ij}':=\begin{cases}a_{ij} & \text{ if }i\ne j;\\
\max(1,a_{ii}) & \text{ if }i=j.\end{cases}$$
Take the quiver $Q':=(I,a_{ij}').$ Note that $N_{\gamma}(Q)=N_{\gamma}(Q'),$ and if we replace $Q$ by $Q',$ then the new fractional $J_{\gamma}$ will be contained in the initial one.
Hence, in order to prove inclusions \eqref{ideal_generates}, we may and will assume that $a_{ii}\geq 1$ for $i\in I,$ and so $J_{\gamma}\subset A_{\gamma}^{prim}.$ We will deduce \eqref{ideal_generates} from the following more general result.
\begin{lemma}\label{effective_gen}Let $\mathrm{k}$ be arbitrary field, and consider the graded algebra of polynomials $B=\mathrm{k}[z_1,\dots,z_n],$ $n\geq 1,$ with grading $\deg(z_i)=1.$ Suppose that
$l_1,\dots,l_s\in B^1$ are pairwise linearly independent non-zero linear forms in $z_i.$ Take some non-empty set of polynomials $\{P_1,\dots,P_r\}\subset B$ of the form
$$P_i=l_1^{d_{i1}}\dots l_s^{d_{is}},$$
where $d_{ij}\in\Z_{\geq 0}.$ Put $d_j:=\max_{1\leq i\leq r}d_{ij},$ $1\leq j\leq s.$ Then the following are equivalent:
(i) $B^d\subset (P_1,\dots,P_r)$ for $d\geq d_1+\dots+d_s-n+1;$
(ii) the ideal $(P_1,\dots,P_r)\subset B$ has finite codimension;
(iii) For any sequence $p_1,\dots,p_r$ of numbers in $\{1,\dots,s\},$ such that $d_{i,p_i}>0$ for $1\leq i\leq r,$ the linear forms $l_{p_1},\dots,l_{p_r}$ generate the space $B^1.$\end{lemma}
\begin{proof}Both implications $(i)\Rightarrow (ii)$ and $(ii)\Rightarrow (iii)$ are evident. So we are left to prove implication
$(iii)\Rightarrow (i).$
Put $D:=d_1+\dots+d_s-n+1.$ If $D\leq 0,$ then one of the polynomials $P_i$ is constant, and there is nothing to prove. So, we assume that $D>0.$
We proceed by induction on $D+n.$ If $D+n=2,$ then $n=s=d_1=D=1,$ hence $(P_1,\dots,P_r)\supset (z_1),$ and the statement is proved.
Assume that the implication holds for $D+n<k_0>2.$ We will prove that it holds for $D+n=k_0.$ Consider the cases.
{\it The case 0.} One of $P_i$ is constant. Then, there is nothing to prove.
{\it The case 1.} We have $P_i=l_j$ for some $i,j.$ Then it suffices to show that the images of $P_{i^\prime}$ with $d_{i'j}=0$ in $B/(l_j)$ generate
$(B/(l_j))^d$ for $d\geq D.$ If $n=1,$ then this is clear, and if $n>1,$ then this follows from the induction hypothesis.
{\it The case 2.} All $P_i$ have degree at least $2.$ Take $d\geq D,$ and $f\in B^d.$ Choose some sequence $p_1,\dots,p_r$ of numbers in $\{1,\dots,s\},$ such that $d_{i,p_i}>0$ for $1\leq i\leq r.$ Then by $(iii)$ we can write
$$f=\sum\limits_{i=1}^rl_{p_i}g_i,\quad g_i\in B^{d-1}.$$
It suffices to show that for each $1\leq i\leq r,$ the polynomial $g_i$ belongs to an ideal generated by $P_{i^\prime}$ with $l_{p_i}\nmid P_{i^\prime},$
and $\frac{P_{i^{\prime\prime}}}{l_{p_i}}$ with $l_{p_i}\mid P_{i^{\prime\prime}}.$ But this follows from induction hypothesis.
In each case, we have proved the desired implication. Induction statement is proved. Lemma is proved.
\end{proof}
Now, consider the cases. If $\sum\limits_{i}\gamma^i=1,$ then $N_{\gamma}(Q)=1,$ and $A_{\gamma}^{prim}=\Q,$ hence inclusions \eqref{ideal_generates}
hold. Further, if $\sum\limits_{i}\gamma^i\geq 2,$ then we apply Lemma \ref{effective_gen} to $B=A_{\gamma}^{prim},$ the linear forms $(x_{j,\alpha_2}-x_{i,\alpha_1})$
(defined up to sign), and polynomials which are in the $P_{\gamma}$-orbit of the expressions \eqref{expr}. They generate precisely the ideal $J_{\gamma}\subset A_{\gamma}^{prim}.$ We have already shown in the proof of Theorem \ref{free_algebra} that the ideal $J_{\gamma}\subset A_{\gamma}^{prim}$ has finite codimension. Therefore, implication $(ii)\Rightarrow (i)$ from Lemma \ref{effective_gen} gives the desired inclusions \eqref{ideal_generates}. Indeed,
we have that $$d_1+\dots +d_s=\frac12(\sum\limits_{\substack{i,j\in I,\\ i\ne j}}a_{ij}\gamma^i\gamma^j+\sum\limits_{i\in I}(a_{ii}-1)\gamma^i(\gamma^i-1)),
\quad n=\sum\limits_{i\in I}\gamma^i-1,$$
hence $N_{\gamma}(Q)=d_1+\dots+d_s-n+1.$ The inclusions \eqref{ideal_generates} and Theorem are proved.
\end{proof}
\section{Applications to quantum DT invariants}
\label{s:DT_invariants}
Define the generating function for the COHA $\cH$ of symmetric quiver $Q$ by the following formula:
$$H_Q(\{t_i\}_{i\in I},q):=\sum\limits_{\gamma\in\Z_{\geq 0}^I,k\in\Z} (-1)^k\dim(\cH_{\gamma,k})t^{\gamma}q^{\frac{k}2}\in \Z((q^{\frac12}))[[\{t_i\}_{i\in I}]],$$
where $t^{\gamma}:=\prod\limits_{i\in I}t_i^{\gamma^i}.$ Note that we have an equality
\begin{equation}\label{stupid}H_Q=\sum\limits_{\gamma\in\Z_{\geq 0}^I}\frac{(-q^{\frac12})^{\chi_Q(\gamma,\gamma)}}{\prod\limits_{i\in I}(1-q)(1-q^2)\dots(1-q^{\gamma^i})}t^{\gamma}.\end{equation}
Recall the notation
$$(z;q)_{\infty}:=\prod\limits_{n\in\Z_{\geq 0}}(1-q^nz)$$
(the so-called $q$-Pochhammer symbol).
\begin{cor}\label{positive}Let $Q$ be a symmetric quiver. Then we have a decomposition
$$H_Q(\{t_i\}_{i\in I},q)=\prod\limits_{\gamma\in\Z_{\geq 0}^I,k\in\Z}(q^{\frac{k}2}x^{\gamma};q)_{\infty}^{(-1)^{k-1}c_{\gamma,k}},$$
where $c_{\gamma,k}$ are non-negative integer numbers.
Moreover, if $c_{\gamma,k}\ne 0,$ then $\gamma\ne 0,$ $$k\equiv \chi_Q(\gamma,\gamma)\text{ mod }2,\quad\text{and}\quad \chi_Q(\gamma,\gamma)\leq k< \chi_Q(\gamma,\gamma)+2N_{\gamma}(Q).$$
In particular, for a fixed $\gamma$ only finitely many of $c_{\gamma,k}$ are non-zero.\end{cor}
\begin{proof}Corollary follows immediately from Theorem \ref{free_algebra} and Theorem \ref{upper_bound} if we put $c_{\gamma,k}=\dim V_{\gamma,k}^{prim}.$
Indeed, the generating function of the free super-commutative subalgebra generated by one element of bidegree $(\gamma,k)$ equals to
$$(1-q^{\frac{k}2}t^{\gamma})^{(-1)^{k-1}}.$$ The resulting decomposition follows from free generation of $\cH$ by $V,$ and from Theorem \ref{upper_bound}.\end{proof}
In the notation of Corollary \ref{positive} and the terminology of \cite{KS}, the polynomials
$$\Omega(\gamma)(q):=\sum\limits_{k\in\Z}c_{\gamma,k}q^{\frac{k}2}\in\Z[q^{\pm \frac12}]$$
are quantum Donaldson-Thomas invariants of the quiver $Q$ with trivial potential, stability function with unique slope, and the dimension vector $\gamma.$
It follows from Corollary \ref{positive} that for $\gamma\ne 0$ we have
$$\Omega(\gamma)(q)=q^{\frac12 \chi_Q(\gamma,\gamma)} \tilde{\Omega}(\gamma)(q),$$
where $\tilde{\Omega}(\gamma)(q)$ is a polynomial with non-negative coefficients, $\tilde{\Omega}(\gamma)(0)=1,$ and $\deg(\tilde{\Omega}(\gamma)(q))<N_{\gamma}(Q).$
We would like to mention a connection with the paper of Reineke \cite{R}. In loc. cit., for each integer $m\geq 1,$ the following $q$-hypergeometric series is considered:
$$H(q,t)=H_m(q,t):=\sum\limits_{n\geq 0}\frac{q^{(m-1)\binom{n}2}}{(1-q^{-1})(1-q^{-2})\dots(1-q^{-n})}t^n\in\Z(q)[[t]].$$
Denote by $Q_m$ the $m$-loop quiver (a quiver with one vertex and $m$ loops). Since $\chi_{Q_m}(n_1,n_2)=(1-m)n_1n_2,$ the formula \eqref{stupid} implies
$$H_m(q,t)=H_{Q_m}((-1)^{m-1}tq^{\frac{1-m}2},q^{-1}).$$
Also, we have $N_n(Q_m)=(m-1)\binom{n}2-n+2.$ Therefore, Corollary \ref{positive} implies the following.
\begin{cor}\label{positive_special}
$$H_m(q,(-1)^{m-1}t)=\prod\limits_{n\geq 1,k\in\Z}(q^kt^n;q^{-1})^{-(-1)^{(m-1)n}d_{n,k}},$$
where $d_{n,k}$ are non-negative integers, and inequality $d_{n,k}>0$ implies $$n-1\leq k\leq (m-1)\binom{n}2.$$
In particular, for a fixed $n$ only finitely many of $d_{n,k}$ are non-zero.\end{cor}
This Corollary is stronger than Conjecture 3.3 in \cite{R}. According to notation of \cite{R}, the quantized Donaldson-Thomas type invariant
$DT_n^{(m)}(q)$ equals to $\sum\limits_{k\in\Z} d_{n,k}q^k.$ Thus, Corollary \ref{positive_special} implies that $DT_n^{(m)}(q)$ is a monic polynomial of degree $(m-1)\binom{n}2,$ divisible by $q^{n-1},$ with non-negative coefficients.
With above said, the numbers $d_{n,k}$ are dimensions of graded components of finite-dimensional graded algebras $\cH_n^{prim}/J_n^{S_n}.$ It would be interesting to compare this interpretation with explicit formulas for $DT_n^{(m)}(q)$ in \cite{R}, Theorem 6.8.
|
\section{Introduction}
The Cosmic Microwave Background (CMB) radiation represents a fundamental observable for Cosmology and at present the most powerful one for the investigation of several open questions, such as the nature
of inflation or primordial non-gaussianity. In two years the Planck satellite \citep{:2011ap} will provide a measure of the anisotropies of the CMB with a precision never reached before, that will allow a highly precise determination of the standard cosmological parameters as well as major constraints on some non-standard Physics processes.
The observation of CMB anisotropies is affected by several systematics and secondary effects due to the fact that the CMB is not the only source of emission in the microwave frequencies and to the formation of structures
between the observer and the last scattering surface \citep{Tegmark:1995pn,Toffolatti:1997dk,DeZotti:2004mn,Aghanim:2007bt}. The great accuracy of future data requires a compelling description of these effects, in order
to separate the different contributions to the anisotropies and to distinguish primordial and secondary effects (see \cite{Millea:2011pa}). The observable used to extract most of the cosmological information from the CMB is the angular power spectrum $C_\ell$. Secondary effects or unresolved foregrounds provide a contribution to the observed $C_\ell$. In order to obtain an unbiased determination of the cosmological parameters from CMB maps it is necessary to correctly describe possible contaminations. On the other hand, both contaminants and secondary effects themselves contain certain cosmological and astrophysical information, especially on the formation of structure at late-times and the large-scale structure of the Universe,
so that the separation of these components from primordial CMB fluctuations becomes an important science goal on its own.
The Galactic emission and radio point sources are typical examples of foreground contamination in CMB maps. While the bright sources detected in maps
can be removed with a suitable mask before the estimation of the angular power spectrum, unresolved point sources will contribute to the total anisotropy power spectrum $C_\ell$.
The Sunyaev-Zel'dovich (SZ) effect \citep{sz70}, caused by the Compton scattering of the CMB photons by the electrons in the Universe, is a well-known secondary anisotropy studied by
a variety of experiments. The SZ effect contains cosmological information, since the angular power spectrum of secondary temperature anisotropy arising
from SZ scattering depends on both the gas distribution in galaxy clusters and on the amplitude of the matter density fluctuations $\sigma_8$ \citep{Barandela_mucket,Komatsu:1999ev,Cooray00,Bond:2002tp}. Since the SZ thermal effect has a unique spectral signature relative to the CMB thermal spectrum, the SZ signal can be distinguished from
primary CMB anisotropies and other foregrounds using observations at multiple frequencies across the SZ null at $\sim 217$ GHz \citep{Cooray:2000xh}.
Such a separation, however, is not feasible with kinetic SZ effect associated with peculiar motions of the electrons scattering the CMB \citep{sz70}
as the signal has the same spectrum as that of the CMB. Even for the SZ thermal effect, in realistic experiments, the
main obstacles that limit a clear detection of the thermal signal comes from uncertainties in the modeling of the kinetic contribution and the
difficulty of separating SZ effects from clustered point sources.
The contribution of unresolved point sources and SZ effect is best seen on small angular scales where they dominate the total CMB angular power spectrum.
The use of data at these small angular scales is hence becoming decisive in the analysis of CMB data. In this work we analyze large $\ell$ data (up to $\ell\sim9000$) from the South Pole Telescope (SPT) at $150$ and $220$ GHz \citep{Shirokoff:2010cs} and from Atacama Cosmology Telescope (ACT) \citep{Dunkley:2010ge,Das:2010ga} at $150$ GHz combined with Wilkinson Microwave Anisotropy Probe (WMAP) data after $7$ years of observation \citep{Komatsu:2010fb} to put constraints on the two $SZ$ effects.
In previous studies significant limitations came from uncertainties associated with
clustering of dusty star-forming galaxies (DSFG) that contribute to high-frequency CMB data. Such clustering has now been measured with both {\it Herschel} \citep{Amblard:2011gc}
and Planck \citep{:2011ap} experiments. In the context of CMB studies, {\it Herschel} measurements are most useful as they probe the DSFG clustering down to
sub-arcminute angular scales at scales well matched to arcminute scale CMB experiments while Planck measurements are limited to scales greater than 5$'$ or $\ell < 2000$.
In this work we describe the clustering of unresolved point sources we used the same template of \cite{Amblard:2011gc}, where the authors reported a detection of both the linear clustering and the non-linear
clustering at a few arcminute scales, corresponding to $\ell \sim 4000$.
We perform a Monte Carlo Markov Chain (MCMC) analysis constraining both the thermal and the kinetic terms of the SZ effect together with the the Poisson and clustering corrections due to unresolved point
sources, including radio sources at lower frequencies such as 150 GHz. In the next Section we recall more details on the contribution of the SZ effect and foregrounds to the CMB anisotropy power spectra.
Section~3 describes the parametrization and the templates we used to model the foreground and the SZ contamination to the CMB angular anisotropy power spectrum. In Section~4
we show the results and conclude with a summary.
\section{Parametrizing SZ effect and foregrounds}
In this Section we briefly describe the adopted parametrizations and templates for the Sunyaev-Zel'dovich effect,
unresolved extragalactic point source foregrounds and lensing.
\vspace{0.3cm}
{\it Sunyaev-Zel'dovich thermal and kinetic effect.}
The SZ effect has two different contributions, one from the thermal motion of the electrons (thermal SZ effect - \emph{tSZ}) and one from the bulk motion of the electrons relative to the CMB (kinetic SZ - \emph{kSZ}). The former contribution has a distinct frequency dependence, while the kSZ effect causes only a Doppler shift of the incident CMB spectrum retaining the black body shape. The total SZ signal in a generic direction $\hat{n}$ is then given by (see for example Section~$2$ in \cite{Dunkley:2010ge}):
\begin{equation}\label{SZsignal}
\Delta T^{\rm SZ}(\nu)=\frac{f(x)}{f(x_0)}\Delta T_0^{\rm tSZ}(\hat{n})+\Delta T^{\rm kSZ}(\hat{n}) \ ,
\end{equation}
with $x=h\nu/k_BT_{CMB}$ and $f(x)=2-x/2\tanh(x/2)$. Here $\Delta T_0^{tSZ}$ is the expected thermal contribution at frequency $\nu_0$. From $f(x)$ it
can easily seen that the thermal SZ effect vanishes at $\sim 218$ GHz.
We model the SZ contributions to the anisotropy angular power spectrum, relative to a template power spectrum, as
\begin {equation}\label{SZ}
D_\ell^{{\rm SZ}, ij}=A_{\rm tSZ}\frac{f(\nu_{i})}{f(\nu_{0})}\frac{f(\nu_{j})}{f(\nu_{0})}D_{0,\ell}^{\rm tSZ}+ A_{\rm kSZ}D_{0,\ell}^{\rm kSZ} \ ,
\end {equation}
where $D_\ell=\ell(\ell+1)C_{\ell}/2\pi$ and $D_{0,\ell}^{i}$ is the template spectrum for either thermal or kinetic SZ.
In this work we consider the SZ templates from \cite{Trac:2010sp}, computed by tracing through a dark matter simulation and processed to include gas in dark matter halos and in the filamentary intergalactic
medium. The thermal SZ template describes the power from tSZ temperature fluctuations from all clusters for a Universe normalized with amplitude of matter fluctuations $\sigma_8 = 0.8$. In particular we use the 'standard' model of \cite{Trac:2010sp}, that was first described in \cite{Sehgal:2009xv}, and assuming a $\sigma_8$ scaling given by $D_{0,\ell}^{tSZ}\propto(\sigma_8/0.8)^{8.1}$ as found in \cite{Trac:2010sp}. For these templates the reference values at $\ell=3000$ are $D_{0,\ell=3000}^{tSZ}\simeq8.9 \mu K^2$ and $D_{0,\ell=3000}^{kSZ}\simeq2.1 \mu K^2$
\vspace{0.3cm}
{\it Foregrounds from unresolved extragalactic point sources.}
The foregrounds contribution to the CMB power spectrum at arcminute angular scales arises essentially from unresolved extragalactic point sources. These sources provide two contributions, a
Poisson term due to the random discrete distribution and a clustering term
accounting for the large-scale distribution of the sources. We assume the Poisson term as constant in
$C_{\ell}$, modeling it as $D_{\ell}^{\rm Poiss}=A_{\rm Poiss}D_{0,\ell}^{\rm Poiss}$ where $D_{0,\ell}^{\rm Poiss}=(\ell/3000)^2$.
The clustered term can be similarly expressed as $D_{\ell}^{{\rm clust},ij}=A_{\rm clust}(\nu_i,\nu_j)D_{0,\ell}^{\rm clust}$, where $D_{0,\ell}^{\rm clust}$ is the point sources clustering template and $A_{\rm clust}(\nu_i,\nu_j)$ encodes the frequency scaling (see section 3 for further details).
Contribution to point sources comes from radio point sources and dusty star-forming galaxies (DSFG).
At 220 GHz the main point source contribution is mainly DSFGs while at 150 GHz the point sources are primarily radio sources with a synchrotron spectrum. We therefore neglect the clustering of radio sources and assume that the contribution from radio sources is essentially described only by a Poisson behavior.
For clustered DSFGs we adopt the template from \cite{Amblard:2011gc} where the authors reported a detection of both the linear clustering and the excess of clustering associated
with the 1-halo term at arcminute scales. Those data are from the {\it Herschel} Multi-tiered
Extra-galactic survey (HerMES) \citep{Oliveretal}, taken with the Spectral and Photometric
Imaging Receiver (SPIRE) onboard the {\it Herschel} Space Observatory \citep{Pilbratt:2010mv}.
\vspace{0.3cm}
{\it CMB Lensing}
It is well known that gravitational lensing of CMB anisotropies by large-scale structure tends to increase
the power at small angular scales (see \cite{Hanson:2009kr} for a recent review). A proper calculation of this effect is hence necessary in order to prevent an incorrect estimate of the foregrounds and
SZ parameters. The calculation of lensed CMB spectra out to $\ell=9000$ is prohibitively expensive in computational time. Instead, we approximate the impact of lensing by adding a fixed lensing template
$D_{\ell}^{lens}$ computed by running \texttt{camb} \citep{Lewis:1999bs} with and without the lensing option and taking the difference between these spectra. In this run the cosmological parameters of the $\Lambda$CDM model are fixed at the best fit values WMAP7. In \cite{Shirokoff:2010cs} it has been estimated that the error due to this approximation is less than $0.5\mu$K$^2$ at $\ell>3000$ and is hence negligible with respect to secondary and foregrounds contributions.
The lensing contribution is clearly frequency independent.
\section{Analysis Method and data}
We place constraints on the cosmological parameters and on the SZ and foregrounds parameters using the $7$-years WMAP data in combination with the SPT data at $150$ GHz and at $220$ GHz, and the ACT data at $148$ GHz. The SPT and ACT datasets are necessary to analyze the smaller scales of the power spectrum where point sources and SZ are dominant. For the SPT data we select the single frequency $15\times 15$ blocks from the full $45\times 45$ covariance matrix provided by the SPT collaboration (see \cite{Shirokoff:2010cs}), neglecting the correlation between different frequencies.
We use a $6$-parameter flat-$\Lambda$CDM cosmological model to describe primary CMB anisotropies and reionization:
the baryon and dark matter physical energy densities $\Omega_{b}h^{2}$, $\Omega_{c}h^{2}$, the reionization optical depth $\tau$, the ratio of the sound horizon to the angular diameter distance at the decoupling $\theta$, the
amplitude of the curvature perturbation $A_{s}$ (with flat prior on $\log A_{s}$) and the spectral index $n_{s}$; these two last parameters are both defined at the pivot scale $k_{0}=0.002$ $\rm hMpc^{-1}$. In addition to the standard cosmological parameters we include the SZ and foreground parameters described in the previous section.
We perform a Monte Carlo Markov Chain analysis based on the publicly available package \texttt{cosmomc} \citep{Lewis:2002ah} suitably modified to account for the additional parameters, with a convergence diagnostic based on the Gelman and Rubin statistics. When estimating parameters with point sources and SZ included, the total CMB anisotropy spectra are three, one for each frequency plus the cross-correlation term, because of the frequency dependence of the secondary anisotropies.
In order to study the stability of our results on the assumed parametrization, we perform three different analysis, both with $6$ additional parameters describing SZ effect and foregrounds, but considering different parametrizations.
\vspace{0.3cm}
{\it First case: "run1".}
In the first case, that we define as "run1" in what follows, we consider the SZ effect parameters $A_{\rm tSZ}$ and $A_{\rm kSZ}$, the Poissonian contribution $A_{\rm Poiss}^{150}$ and $A_{\rm Poiss}^{220}$ and the Poisson contribution for the 150$\times$220 GHz cross-correlation $A_{\rm Poiss}^{X}$.
The use of the $A_{\rm Poiss}^{X}$ extra parameter for the cross correlation of Poisson point source can be justified from
the possibility that the contribution at one single frequency comes from more than one point source population and that the
two channels are not fully correlated. This possibility has not been considered in previous analyses and is therefore important to evaluate the impact of this assumption.
Finally we consider a {\it single} clustered point sources parameter, $A_{\rm clust}$, scaling the contribution at different frequencies using the relation of \cite{Gispert:2000np}:
\begin{eqnarray*}
I_{\nu}=8.80\times10^{-5}\left(\nu/\nu_0\right)P_\nu\left(13.6K\right) \ ,
\end{eqnarray*}
with $\nu_0 = 100 {\rm cm} ^{-1}$, following recent results from Planck \citep{:2011ap}. In what follows we
refer to this scaling as "Gispert" scaling.
In summary, the spectra in "run1" are defined as:
\begin{eqnarray*}
D_{\ell}(150)&=&D_{\ell}^{lens}+A_{\rm tSZ}D_{0,\ell}^{\rm tSZ}+A_{\rm kSZ}D_{0,\ell}^{\rm kSZ} \nonumber \\
&& +A_{\rm clust}D_{0,\ell}^{\rm clust150} +A_{\rm Poiss}^{150}D_{0,\ell}^{\rm Poiss} \\
D_{\ell}(220)& = & D_{\ell}^{lens}+A_{\rm kSZ}D_{0,\ell}^{\rm kSZ}\nonumber \\
&& +A_{\rm clust}D_{0,\ell}^{\rm clust220}+A_{\rm Poiss}^{220}D_{0,\ell}^{\rm Poiss} \\
D_{\ell}({\rm 150\times220})&=&D_{\ell}^{lens}+A_{\rm kSZ}D_{0,\ell}^{\rm kSZ}\nonumber \\
&& +A_{\rm clust}D_{0,\ell}^{\rm clustcross}+A_{\rm Poiss}^{\rm cross}D_{0,\ell}^{\rm Poiss}
\end{eqnarray*}
The thermal SZ effect is negligible at $220$ GHz. The contribution of the thermal SZ effect to the cross-correlated power spectrum may not vanish in presence of a spatial correlation between IR sources and the clusters that cause the thermal SZ. Nevertheless as showed in \cite{Shirokoff:2010cs} the effect of this correlation is negligible for the SPT data (see par. 7.4 in \cite{Shirokoff:2010cs} for further details). We hence do not consider this contribution when fitting the data.
\vspace{0.3cm}
{\it Second case: "run2".}
In the second analysis, to which in what follows we refer as "run2", we assume full correlation between the Poisson point sources signal at $150$ and $220$ GHz as done in previous analyses, i.e. we fix the cross amplitude of Poisson point sources at the square root of the product of the amplitudes at $150$ and $220$, $A_{\rm Poiss}^{X} =\sqrt{A_{\rm Poiss}^{150}A_{\rm Poiss}^{220}}$.
Moreover, we don't scale the clustered point sources template and we use instead two different parameters for $150$ and $220$ GHz.
This second analysis is more similar to the one presented in \cite{Shirokoff:2010cs}, however we point out that while
here we consider the amplitudes at different frequencies as free parameters, \cite{Shirokoff:2010cs} considered
the amplitudes at one single frequency and one common frequency spectral index for clustered and point sources
as free parameters.
In this second case the foreground spectra are defined as:
\begin{eqnarray*}
D_{\ell}(150)&=&D_{\ell,150}^{lens}+A_{\rm tSZ}D_{0,\ell}^{\rm tSZ}+A_{\rm kSZ}D_{0,\ell}^{\rm kSZ}\nonumber\\
&& +A_{\rm clust150}D_{0,\ell}^{\rm clust}+A_{\rm Poiss}^{150}D_{0,\ell}^{\rm Poiss} \\
D_{\ell}(220)& = & D_{\ell,220}^{lens}+A_{\rm kSZ}D_{0,\ell}^{\rm kSZ}\nonumber \\
&& +A_{\rm clust220}D_{0,\ell}^{\rm clust}+A_{\rm Poiss}^{220}D_{0,\ell}^{\rm Poiss} \\
D_{\ell}({\rm 150\times220})&=&D_{\ell,X}^{lens}+A_{\rm kSZ}D_{0,\ell}^{\rm kSZ} \nonumber \\
&& +\sqrt{A_{\rm clust150}A_{\rm clust220}} D_{0,\ell}^{\rm clust} \nonumber \\
&& +\sqrt{A_{\rm Poiss150}A_{\rm Poiss220}}D_{0,\ell}^{\rm cross}
\end{eqnarray*}
\vspace{0.3cm}
{\it Third case: "run3".}
Finally we combine $150$ GHz data of SPT and ACT, using separate parameters for ACT and SPT both for clustered and Poisson point sources, to account for the different masking thresholds of the point sources. In this case we have:
\begin{eqnarray*}
D_{\ell}(150)&=&D_{\ell,150}^{lens}+A_{\rm tSZ}D_{0,\ell}^{\rm tSZ}+A_{\rm kSZ}D_{0,\ell}^{\rm kSZ}\nonumber \\
&& +A_{\rm clustACT}D_{0,\ell}^{\rm clust}+A_{\rm PoissACT}^{150}D_{0,\ell}^{\rm Poiss}\nonumber \\
&& +A_{\rm clustSPT}D_{0,\ell}^{\rm clust}+A_{\rm PoissSPT}^{150}D_{0,\ell}^{\rm Poiss} \nonumber
\end{eqnarray*}
\begin{table*}[!]
\begin{center}
\begin{tabular}{lrrrr}
\hline
\hline
&WMAP7+ & WMAP7+ &WMAP7+ &WMAP7+\\
& SPT (run1, kSZ free ) & SPT (run1) &SPT(run2, kSZ free) & SPT (run2)\\
\hline
$ 10^2\Omega_b h^2$ & $2.267\pm0.049$ & $2.268\pm0.051$ & $2.264\pm0.049$ & $2.269\pm0.049$ \\
$ \Omega_c h^2$ & $0.113\pm0.005$ & $0.113\pm0.004$ & $0.1126\pm0.0052$ & $0.1127\pm0.0052$ \\
$ \tau$ & $0.090\pm0.015$ & $0.089\pm0.015$ & $0.089\pm0.015$ & $0.090\pm0.014$ \\
$ n_s$ & $0.973\pm0.013$ & $0.973\pm0.013$ & $0.972\pm0.013$ & $0.972\pm0.012$ \\
$ln(10^{10} A_s)$ & $3.18\pm0.04$ & $3.18\pm0.04$ & $3.18\pm0.045$ & $3.18\pm0.04$ \\
$ \Omega_m$ & $0.278\pm0.028$ & $0.279\pm0.029$ & $0.276\pm0.029$ & $0.276\pm0.028$ \\
$\sigma_8$ & $0.823\pm0.028$ & $0.825\pm0.028$ & $0.820\pm0.0272$ & $0.821\pm0.026$ \\
$A_{\rm tSZ}$ & $0.24\pm0.17$ & $0.25\pm0.16$ & $0.33\pm0.23$ & $0.52\pm0.22$ \\
$A_{\rm kSZ}$ & $1.3\pm0.9$ & $\left[1\right]$ & $2.7\pm1.4$ & $\left[1\right]$ \\
$A_{\rm clust}$ & $1.05\pm0.19$ & $1.08\pm0.14$ & $ - $ & $ - $ \\
$A_{\rm clust150}$ & $ - $ & $ - $ & $0.44\pm0.27$ & $0.66\pm0.26$ \\
$A_{\rm clust220}$ & $ - $ & $ - $ & $8.2\pm1.7$ & $8.7\pm1.5$ \\
\hline
$D_{\ell 3000}^{\rm tSZ}$($\mu$K$^2$) & $2.2\pm1.5$ & $2.3\pm1.4$ & $2.9\pm2.0$ & $4.7\pm2.0$ \\
$D_{\ell 3000}^{\rm kSZ}$($\mu$K$^2$) & $2.7\pm1.9$ & $\left[2.05\right]$ & $5.5\pm3.0$ & $\left[2.05\right]$ \\
$D_{\ell 3000}^{\rm clust150}$($\mu$K$^2$) & $6.05\pm1.06$ & $6.26\pm0.82$ & $2.51\pm1.60$ & $3.81\pm1.53$ \\
$D_{\ell 3000}^{\rm clust220}$($\mu$K$^2$) & $39.11\pm6.79$ & $40.63\pm5.08$ & $47.33\pm9.78$ & $50.47\pm9.17$ \\
$D_{\ell 3000}^{\rm Poiss150}$($\mu$K$^2$) & $10.03\pm0.67$ & $10.1\pm0.7$ & $10.38\pm0.63$ & $10.33\pm0.67$ \\
$D_{\ell 3000}^{\rm Poiss220}$($\mu$K$^2$) & $79.5\pm4.8$ & $80\pm5$ & $77.89\pm4.49$ & $76.5\pm4.0$ \\
$D_{\ell 3000}^{\rm Poisscross}$($\mu$K$^2$) & $26.8\pm1.4$ & $26.8\pm1.4$ & $ - $ & $ - $ \\
\hline
\hline
\end{tabular}
\caption{Mean values and $68\%$ error bars from SPT data at 150 and 220 GHz. Run1 case is with only one DSFG clustering
amplitude allowed to vary and frequency scaling fixed from \cite{Gispert:2000np}, consistent with Planck \citep{:2011ap}.
Run2 case is with two DSFG clustering amplitudes allowed to vary, without frequency scaling.}
\label{spt_tab}\vspace{1cm}
\begin{tabular}{lrr}
\hline
\hline
&WMAP7+ & WMAP7+\\
& SPT+ACT (kSZ free ) & SPT+ACT \\
\hline
$ 10^2\Omega_b h^2$ & $2.232\pm0.047$ & $2.234\pm0.046$ \\
$ \Omega_c h^2$ & $0.1121\pm0.0050$ & $0.1124\pm0.0053$ \\
$ \tau$ & $0.086\pm0.014$ & $0.086\pm0.015$ \\
$ n_s$ & $0.964\pm0.012$ & $0.964\pm0.012$ \\
$ln(10^{10} A_s)$ & $3.20\pm0.043$ & $3.19\pm0.04$ \\
$ \Omega_m$ & $0.274\pm0.027$ & $0.275\pm0.028$ \\
$\sigma_8$ & $0.812\pm0.0255$ & $0.813\pm0.027$ \\
$A_{\rm tSZ}$ & $0.34\pm0.25$ & $0.38\pm0.24$ \\
$A_{\rm kSZ}$ & $1.6\pm1.1$ & $\left[1\right]$ \\
$A_{\rm clustact}$ & $0.66\pm0.56$ & $0.75\pm0.59$ \\
$A_{\rm clustspt}$ & $0.66\pm0.43$ & $0.77\pm0.41$ \\
\hline
$D_{\ell 3000}^{\rm tSZ}$($\mu$K$^2$) & $3.1\pm2.3$ & $3.5\pm2.2$ \\
$D_{\ell 3000}^{\rm kSZ}$($\mu$K$^2$) & $3.2\pm2.3$ & $\left[2\right]$ \\
$D_{\ell 3000}^{\rm clustact}$($\mu$K$^2$) & $3.9\pm3.2$ & $4.2\pm3.2$ \\
$D_{\ell 3000}^{\rm Poissact}$($\mu$K$^2$) & $13.4\pm2.4$ & $13.5\pm2.5$ \\
$D_{\ell 3000}^{\rm clustspt}$ ($\mu$K$^2$) & $3.8\pm2.5$ & $4.5\pm2.4$ \\
$D_{\ell 3000}^{\rm Poissspt}$($\mu$K$^2$) & $10.2\pm0.8$ & $10.2\pm0.8$ \\
\hline
\hline
\end{tabular}
\caption{Mean values and $68\%$ error bars from ACT data combined with SPT data at $150$ GHz}
\label{spt_act_tab}\vspace{1cm}
\end{center}
\end{table*}
\section{Results}
In Table~\ref{spt_tab} we report the mean values of the cosmological parameters and their $68\%$ C.L. uncertainty from SPT data at $150$ and $220$ GHz for the "run1" and "run2" analyses, while
in Table~\ref{spt_act_tab} we list the mean values of the cosmological parameters and their $68\%$ C.L. uncertainty from ACT data ("run3") combined with SPT data at $150$ GHz.
In order to facilitate the comparison with other works present in the literature we also translate the constraints on
the foregrounds amplitudes in to the foreground power spectrum at $\ell=3000$, $D_{\ell=3000}$.
Since a significant correlation exists between thermal and kinetic SZ and since the kinetic SZ is predicted to be
small, we also perform an analysis by fixing $D_{\ell=3000}^{\rm kSZ}=2 \, \mu K^2$.
We find that for the ``run1'' case the thermal SZ anisotropy amplitude is $D_{\ell=3000}^{\rm tSZ}=2.2\pm1.5 \, \mu K^2$ . While a $\sim 1 \sigma$ indication for SZ is present our result is less significant than the one reported by \cite{Shirokoff:2010cs} with $D_{\ell=3000}^{\rm tSZ}=3.2\pm1.3 \, \mu K^2$ i.e. with a thermal SZ detection at at more than two standard deviations. The result on the kinetic SZ component are compatible, with
$D_{\ell=3000}^{\rm kSZ}=2.7\pm1.9 \, \mu K^2$ at $68 \%$ c.l. from our analysis to be compared with
$D_{\ell=3000}^{\rm kSZ}=2.4\pm2.0 \, \mu K^2$ from \cite{Shirokoff:2010cs}.
\begin{figure} [h!]
\includegraphics[width=\columnwidth]{fig1.eps}
\caption{Joint two-dimensional posterior probability contours showing $68\%$ and $95\%$ C.L. constraints on $D_{\ell=3000}^{\rm ksz}$ and $D_{\ell=3000}^{\rm tSZ}$ from ACT $150$ GHz data (red) and SPT all frequencies data (blue) for the run2 case.}
\label{sz}
\end{figure}
Although the point sources and SZ parameters do not show significant degeneracies with cosmological parameters (see also \cite{Serra08}), a strong correlation exists between $A_{\rm tSZ}$ and $A_{\rm kSZ}$ and to a smaller extent between $A_{\rm tSZ,kSZ}$ and $A_{\rm clust}$. This can be seen in Figure~\ref{sz} where we show the $2-D$ likelihood constraints in the plane $D_{\ell=3000}^{\rm kSZ}-D_{l=3000}^{\rm tSZ}$ for the "run2" and "run3" case. Fixing the kSZ term sligthly improves the detection for the
thermal SZ with $D_{\ell=3000}^{\rm tSZ}=2.3\pm1.4 \, \mu K^2$ in ``run1'' but still with less significance than the one in
\cite{Shirokoff:2010cs} where a value of $D_{\ell=3000}^{\rm tSZ}=3.5\pm1.0 \, \mu K^2$ is reported.
Based on the degeneracy direction of Figure~\ref{sz}, we constrain the sum of the SZ effects at
$\ell=3000$ to be $D^{\ell=3000}_{\rm tSZ}+0.5 D^{\ell=3000}_{\rm kSZ}= 3.5 \pm 1.8$ $\mu$K$^2$ to be compared $4.5 \pm 1.0$ $\mu$K$^2$ of \cite{Shirokoff:2010cs} .
These amplitudes are consistent but, again, the significance of the detection is worse than \cite{Shirokoff:2010cs} who found this sum to be $4.5 \pm 1.0$ $\mu$K$^2$.
The small discrepancy with the results presented in \cite{Shirokoff:2010cs} comes essentially
from the different parametrization used.
Adopting a more similar parametrization as in the case of "run2" we found $D_{\ell=3000}^{\rm tSZ}=2.9\pm2.0 \, \mu K^2$, $D_{\ell=3000}^{\rm kSZ}=5.5\pm3.0 \, \mu K^2$ at $68 \%$ c.l., $D^{\ell=3000}_{\rm tSZ}+0.5 D^{\ell=3000}_{\rm kSZ}= 5.6 \pm 2.6$ $\mu$K$^2$, yielding a detection for the thermal SZ with higher significance.
In case of fixed kSZ we obtain $D_{\ell=3000}^{\rm tSZ}=4.7\pm2.0 \, \mu K^2$, again a more significant detection in better agreement with \cite{Shirokoff:2010cs}.
The different assumptions in the frequency scaling of the clustered point sources component in "run1" and "run2"
is the main explanation for the difference in the results.
In ``run1'', taking into account Gispert scaling \citep{Gispert:2000np}, we find $D_{\ell=3000}^{\rm clust220}=39.11\pm6.79 \, \mu K^2$, while in ``run2'', when the amplitude of the clustering point sources is allowed to vary, we have $D_{\ell=3000}^{\rm clust220}=47.33\pm9.78 \, \mu K^2$, that is more consistent with the corresponding value of $D_{\ell=3000}^{\rm clust220}=57\pm9$ reported in \cite{Shirokoff:2010cs}. A small tension therefore exists between the Gispert scaling and the data at $220$ GHz, resulting also in a worse determination of the thermal SZ signal.
The use of a different parametrization of the point source (just amplitudes in our
case while \cite{Shirokoff:2010cs} varies one amplitude and one spectral index per component) can explain
the remaining differences.
Concerning the Poisson point sources component at $150$ GHz, our results are different when
directly compared to those from \cite{Shirokoff:2010cs}, both in ``run1'' and ``run2''. At $150$ GHz we find $D_{\ell=3000}^{\rm Poiss150}=10.03\pm0.67 \, \mu K^2$ in ``run1'' and $D_{\ell=3000}^{\rm Poiss150}=10.38\pm0.63 \, \mu K^2$ in ``run2'', while the value in \cite{Shirokoff:2010cs} is $D_{\ell=3000}^{\rm Poiss150}=7.4\pm0.6 \, \mu K^2$.
This difference is explained in straightforward terms if we take in account that in \cite{Shirokoff:2010cs} radio galaxies are included in their ``baseline model'' with an amplitude $D_{\ell=3000}^{\rm r}=1.28 \, \mu K^2$ with a $15\%$ uncertainty. Clustering of radio galaxies is negligible, so this radio galaxies term is a Poisson like term of the form $ \propto \ell^2$. Adding this component, our Poisson amplitudes are consistent with those reported by \cite{Shirokoff:2010cs} within $1\sigma$. In ``run1'' at $150$ GHz we find $D_{\ell=3000}^{\rm Poiss150}=10.03\pm0.67 \, \mu K^2$, while in \cite{Shirokoff:2010cs} the correspondent total Poisson contribution at $\ell=3000$ is about $(8.68 \pm 0.69) \, \mu K^2$. At $220$ GHz we find $D_{l=3000}^{\rm Poiss220}=79.5\pm4.8 \, \mu K^2$ in ``run1'' and $D_{\ell=3000}^{\rm Poiss220}=77.89\pm4.49 \, \mu K^2$ in ``run2'', while the value in \cite{Shirokoff:2010cs} is $D_{\ell=3000}^{\rm Poiss220}=71\pm5 \, \mu K^2$.
We can therefore conclude that the current results presented in the literature on the
amplitude of the secondary anisotropies should be considered with great care since there is
a clear dependence on the parametrization used, on the frequency scaling adopted and on the
assumed templates. We stress that, a part for small discrepancies imputable to differences in the parameterization, all our results for the SZ amplitudes from the analysis of SPT data both for our "run1" and "run2" cases, are substantially consistent with the analysis of the same data made by \cite{Shirokoff:2010cs}, even if we are finding less tight constraints. Our results hence compare in the same way to the recent predictions of tSZ power made by the models of \cite{Shaw2010}, \cite{Trac:2010sp} and \cite{Battaglia2010} confirming that these models overestimate the power of the tSZ signal, as already found in \cite{Shirokoff:2010cs}.
\begin{figure}[h!]
\includegraphics[width=\columnwidth]{fig3.eps}
\caption{Planck DSFG clustering data (red points) at $217$ GHz and SPT data at $220$ GHz (white points) compared with the combination (solid black line) of the Poisson term (green line) and clustering term (red) of unresolved point sources by scaling the best-fit model to measurements made with {\it Herschel} at 350 $\mu$m to 217 GHz using the frequency scaling of
\cite{Gispert:2000np}. We show the 1-halo (pink) and 2-halo (orange) contributions to the clustering term following \cite{Amblard:2011gc}.
The dashed lines are the $220$ GHz SPT DSFG power spectrum components from \cite{Hall:2009rv}, which resulted in an overestimate of Planck DSFG clustering at $\ell < 3000$.
}\label{CIB}
\end{figure}
In Figure \ref{CIB} we show the recent CIB power spectra data of the Planck collaboration \citep{:2011ap} at 217 GHz and small angular scale CMB power spectrum data from SPT, at 220 GHz, with a comparison to scaled measurements from \cite{Amblard:2011gc}. The {\it Herschel} model is shown in terms of the 1-halo and 2-halo contributions to the total power spectrum. For reference, we also show the model used by \cite{Hall:2009rv} at 220 GHz to describe the clustering of DSFGs, which overestimated the power at tens of arcminute angular scales and above relative to {\it Herschel} and Planck DSFG clustering measurements. \cite{Hall:2009rv} used a linear model to analyze their data. At small angular scales non-linear effects are not negligible and using a linear model to interpret the data may lead to a wrong determination of the bias and hence to an overestimation of the power at larger angular scales (see also discussion in \cite{:2011ap}). Instead our model shows a good fit of both Planck and SPT CIB.
In Figure~\ref{spectra} we show the best fit models for each component compared with the SPT and WMAP7 data. In the ``run1'', when only one amplitude of clustered DSFGs is allowed to vary when fitting the all frequencies SPT data combined with WMAP7 data, we find that $A_{\rm clust}=1.05\pm0.19$. This suggests that the combination of \cite{Amblard:2011gc} model and the frequency scaling for the mean CIB is a good fit of the DSFG clustering at lower CMB frequencies. Higher precision CMB power spectra at $150$, $220$ and $350$ GHz and a direct cross-correlation of {\it Herschel}-SPIRE maps against the CMB will be necessary to study if fluctuations scale with frequency as the mean CIB intensity and to improve overall constraints on secondary anisotropies.
\begin{figure*}[ht!]
\begin{center}
\includegraphics[width=2 \columnwidth]{fig2.eps}
\end{center}
\caption{Contribution to the angular anisotropy power spectrum from point sources and from SZ effect for the best fit model of the WMAP7+SPT analysis. Left panel is $150$ GHz, middle $220$ GHz and right panel shows the cross spectra. kSZ term is the orange solid line and the tSZ term at $150$ GHz is the purple line. Green lines are the Poisson terms and blue lines are the clustering contributions. The black lines are the total best fit power spectra. Black dots are SPT data and red squares are WMAP7 data. The bottom panels show the residual relative to the total model, including primordial CMB and best-fit secondary anisotropy amplitudes.}\label{spectra}
\end{figure*}
\section{Conclusions}
In conclusion, we provided a new analysis of the foreground contribution to the CMB data making use of the
latest ACT and SPT results. Our work is complementary to those presented by the SPT and ACT experimental teams
since we use a different parametrization of the foregrounds contribution and different templates.
The foreground contribution from Poisson point sources at $220$ and $150$ GHz is detected with very high
significance (at more than $\sim 15$ standard deviations) with no particular dependence on the
parametrization used ("run1" and "run2" cases are giving very consistent results).
The contribution from clustered point sources is also well detected at $220$ GHz. We have found that current CMB
data favours a larger contribution at this frequency than the one expected by the Gispert frequency scaling once
the data is normalized at $150$ GHz.
The thermal SZ component is detected at a level slightly above the two standard deviations. However a different
parametrization of the components and the assumption of the Gispert scaling could bring this detection to about
one standard deviation. The correlation with the kinetic SZ term is present in the data despite the
multi-frequency approach. More data at more frequencies are clearly needed to establish a strong detection
of the SZ term.
While our constraints do not improve the results in the literature, we have made a significant addition to prior studies by firmly establishing the power spectrum of DSFGs that dominate the arcminute scale CMB anisotropies at 220 GHz and higher frequencies. This comes from the recent {\it Herschel} results combined with Planck-confirmed frequency spectrum for the CIB mean intensity. In future, additional improvements will come from directly cross-correlating the CMB maps against high-resolution CIB maps from {\it Herschel}; for this a {\it Herschel}-SPIRE survey at the same large areas as CMB surveys will become useful \citep{Cooray10}.
\section{Acknowledgments}
It is a pleasure to thank Erminia Calabrese for useful suggestions and comments. MA thanks the group at UCI for
hospitality while this research was conducted. We thank NSF AST-0645427 and NASA NNX10AD42G for support.
|
\section{Introduction}
Searches for events with large missing transverse energy (MET) are a major focus of experiments at the Large Hadron Collider (LHC), motivated mainly by the fact that such events are generically predicted in supersymmetric extensions of the Standard Model (SM). The CMS and ATLAS collaborations recently published the results of the first such search, in the jets+MET channel, using the 35 pb$^{-1}$ of data collected during the 2010 run of the LHC with $\sqrt{s}=7$ TeV~\cite{CMS,ATLAS}. No excess over the expected SM background was observed, allowing to place bounds on the supersymmetric models which are already stronger than the previous Tevatron bounds. However, supersymmetry is not the only extension of the SM which predicts anomalous events with large MET. In fact, the prediction is very generic: all that's required of a model is that the heavy TeV-scale states carry a new conserved quantum number not carried by any SM particles, and that the lightest particle carrying this quantum number (which is automatically stable) be electrically neutral and color-singlet. Non-supersymmetric examples with these properties include models with universal extra dimensions (UED)~\cite{ued} and Little Higgs with T-parity (LHT)~\cite{LH,LHT,LHreview}. As in supersymmetry, the TeV-scale particles in these models can be paired with SM states of the same gauge quantum numbers; unlike supersymmetry, these theories predict that the heavy particles have the {\it same} spin as their SM partners. Searches for jets+MET events put constraints on these theories as well. In this paper, we will reinterpret the null result of the CMS search~\cite{CMS} in terms of bounds on a simple extension of the SM containing vector-like fermion partners for the light quarks, and an additional massive gauge boson, a partner of the SM photon (or, more precisely, the hypercharge gauge boson). This extension can be thought of as a ``simplified model"\footnote{See {\tt www.lhcnewphysics.org} for philosophy and examples of the simplified model approach.} corresponding to a limit of UED or LHT where all exotic particles other than the ones included in the model are either too heavy or too weakly coupled to be produced in significant numbers at the LHC (at least with current luminosity). In the LHT, such a situation is quite natural, and in fact this simplified version of the model has already been used as a basis for the Tevatron search for the T-odd quarks~\cite{LHT_TeV,LHT_D0}. We will thus refer to the heavy vector-like fermions as ``T-quarks". In UED, the Kaluza-Klein (KK) partners tend to be approximately degenerate in mass, and thus a simple truncation to just a handful of states is less likely to capture the phenomenology correctly; still, there is a limit of the model (which can be taken by adjusting brane-localized kinetic terms~\cite{ued2}) where our simple analysis would apply.
Following Ref.~\cite{LHT_TeV}, we introduce four new vector-like (Dirac) fermions, $\tilde{Q}_i=(\tilde{U}_i, \tilde{D}_i)$, with the same gauge quantum numbers as the SM left-handed quarks $Q_i$, namely $({\bf 3}, {\bf 2})_{1/3}$. Here $i=1, 2$ is the generation index. There's also a discrete $Z_2$ symmetry, which we will call the T-parity, under which $\tilde{Q}_i\rightarrow-\tilde{Q}_i$. All SM states are invariant under this symmetry. We also introduce a massive vector boson $\tilde{B}^\mu$, which has no SM gauge charges, is odd under T-parity, and is coupled to fermions via
\begin{equation}
{\cal L}_{\rm int} = c g^\prime \left( \bar{Q}_i \gamma^\mu P_L \tilde{Q}_i + {~\rm h.c.}\right) \,B_\mu\,,
\eeq{Lint}
where sum over $i$ is implicit and $c$ is an order-one number whose exact value will play no role in our discussion, as long as the interaction is perturbative. We will assume that the four T-quarks have the same mass, $\tilde{M}$. This is motivated by constraints from flavor-changing neutral currents in the LHT model~\cite{LHTflavor}, as well as desire for maximal simplicity. We will further assume that $M(\tilde{B})<\tilde{M}$, so that the $\tilde{B}$ is the lightest T-odd particle (LTP), and is therefore stable. As discussed in~\cite{LHT_TeV}, this set of particles and mass hierarchies (with a few additional states above $\tilde{M}$ which will play no role in this analysis) arises naturally in the LHT model; moreover, $\tilde{B}$ can play the role of dark matter candidate~\cite{LHT_DM}, further motivating this spectrum.
At the LHC, the dominant production mechanism in this model is pair-production of T-quarks via strong interactions:
\begin{equation}
q\bar{q}\rightarrow\tilde{U}_i\bar{\tilde{U}}_i,~~~gg\rightarrow\tilde{U}_i\bar{\tilde{U}}_i,
\eeq{process}
and same for $\tilde{D}_i$. The produced T-quarks decay promptly via the interaction in Eq.~\leqn{Lint}, {\it e.g.}
\begin{equation}
\tilde{U}_i\rightarrow u_i \tilde{B},
\eeq{decay}
producing two high-$p_T$ jets and MET carried by the pair of $\tilde{B}$'s.
Since no other decays are possible, the branching ratio of the decay~\leqn{decay} is one throughout the parameter space, independent of $c$. The phenomenology of the model is completely described by two parameters, $\tilde{M}$ and $M_{\rm LTP}\equiv M(\tilde{B})$. The main goal of this paper is to estimate the current and near-future reach of the LHC search for jets+MET in this parameter space, taking the published CMS search~\cite{CMS} as the benchmark analysis.
\section{Simulations, Analysis and MSSM Validation}
We implemented the model described above in {\tt MadGraph/MadEvent}~\cite{MG}. The process~\leqn{process} was simulated using {\tt MadEvent}, and the decays~\leqn{decay} of the T-quarks were simulated using {\tt BRIDGE}~\cite{bridge}. (We checked, for a few representative points in the model parameter space, that the results are identical to simulating the production and decay together as a $2\rightarrow 4$ process in {\tt MadEvent}, but this approach is less computationally efficient and was not used in our scan over the model parameters.) The CTEQ6M PDF set~\cite{cteq} was used. Simulations were performed for a single T-quark flavor. We ignored the electroweak contribution to the production process, which is suppressed by a factor of order $\alpha^2/\alpha_s^2\sim 0.01$ with respect to the leading QCD contribution. In this approximation, production cross section and kinematic distributions are identical for the 4 T-quark flavors. The parton-level events created by {\tt MadEvent} were passed on to {\tt Pythia}~\cite{pythia} for hadronization and showering. A fast detector simulation was then performed using the {\tt PGS} (Pretty Good Simulation) package~\cite{PGS}, with the parameters set to approximate the CMS detector.\footnote{We used the file {\tt pgs\_card\_CMS.dat} included in the
{\tt MadGraph/MadEvent} package to set the PGS parameters.} The resulting event sample was then subjected to the following set of cuts, which were chosen to approximate those used by the CMS study~\cite{CMS} as closely as possible:
\begin{enumerate}
\item Lepton veto: An event is rejected if a lepton (including an identified $\tau$) is present, with $p_T>10$ GeV, $|\eta|\leq 3$, and separated by $\Delta R\geq 0.4$ from every other object in the event (not including MET).
\item Jet acceptance: Jets with $p_T<50$ GeV or $|\eta|\geq 3$ are deleted. If the resulting event has $<2$ jets, it is discarded.
\item Jet selection: An event is accepted if and only if it has at least two jets with $p_T>100$ GeV, and the leading (highest-$p_T$) jet is at $|\eta|\leq 2.5$.
\item $H_T$ cut: An event is rejected if $H_T<350$ GeV, where $H_T$ is defined as the scalar sum of $p_T$'s of all jets in the event.
\item $\alpha_T$ cut: For a 2-jet event, we define $\alpha_T=E_{T2}/M_T$, where $E_{T2}$ is the energy of the least energetic of the jets, and $M_T$ is the transverse mass of the dijet system~\cite{alpha}. For an event with $\geq 3$ jets, we first combine the jets into 2 pseudo-jets, choosing the partition which minimizes the $E_T$ difference between the two pseudo-jets. We then define $\alpha_T$ in the same way as before, with pseudo-jets replacing the jets. In either case, the event is rejected if $\alpha_T<0.55$.
\end{enumerate}
We simulated 10000 events for each of 391 points in the LHT parameter space, spanning the ranges $\tilde{M}=220\ldots 700$ GeV, $M_{\rm LTP}=20\ldots \tilde{M}-20$ GeV, with most of this area covered with a (20 GeV)x(20 GeV) grid. (We also explored additional points in the region of quasi-degenerate T-quark and LTP, see below.) For each point, we obtain the total cross section for production of the 4 flavors of T-quarks (which of course only depends on $\tilde{M}$), and the combined efficiency of the cuts. We then apply a K-factor to set the total cross section to its NLO value, with renormalization scale set to $\tilde{M}$~\cite{NLO_xsec}. (We used the simple parametrization of the NLO cross section given in the Appendix B of Ref.~\cite{berger}.) Multiplying these numbers yields the total signal cross section after cuts, which can then be compared with present or extrapolated bounds inferred from the CMS analysis.
We estimate that the combined renormalization scale and PDF uncertainty on the total cross section to be about 20\% (see, for example, Fig.~2~(c) in Ref.~\cite{berger}). By comparing cut efficiencies computed from simulations with fixed renormalization and factorization scales (which we use for our central-value estimates) to simulations with renormalization and factorization scales adjusted on event-by-event basis according to jet $p_T$'s, for a few representative points in the model parameter space, we estimate the Monte Carlo uncertainty in the efficiency calculation to be of order 10\%. Combining the two, we conservatively assign an uncertainty of 25\% to the theoretical prediction of the signal cross section after cuts.
\begin{table}[t]
\begin{center}
\begin{tabular}{|c||c|c||c|c|} \hline
& \multicolumn{2}{|c||}{Total Efficiency} & \multicolumn{2}{|c|}{Signature Efficiency} \\
\hline
channel & CMS & ~~~PS~~~ & CMS & PS \\
\hline
$\tilde{q}\tilde{q}$ & $16.0\pm0.1$ & 18.8 & $22.2\pm0.4$ & 21.7 \\
$\tilde{q}\tilde{g}$ & $14.4\pm0.1$ & 20.3 & $23.0\pm0.5$ & 23.9 \\
$\tilde{g}\tilde{g}$ & $12.0\pm0.4$ & 18.5 & $22.5\pm2.0$ & 22.6 \\
\hline
\end{tabular}
\caption{Signal efficiencies, in \%, for the cMSSM, LM1 point. ``CMS" denotes values reported by the CMS collaboration~\cite{CMS}; ``PS" refers to the analysis described in this section.}
\label{tab:valid}
\end{center}
\end{table}
To check that this simple procedure yields sensible results, we validated it by comparing the cut efficiencies for the supersymmetric signal at the point LM1 in the constrained minimal supersymmetric model (cMSSM), estimated using exactly the same procedure we use for T-quarks, with the corresponding efficiencies presented by CMS~\cite{CMS}. The results of this comparison are shown in Table~\ref{tab:valid}. Here, ``total" efficiencies denote the number of events passing all cuts normalized to the total number of events, while ``signature" efficiencies are normalized to the number of events after the lepton veto. It is clear that our procedure gives excellent match to the CMS results for the signature efficiencies, which are within 2 standard deviations for all 3 production channels. Our procedure does not work nearly as well for the total efficiencies. This is most likely due to our extremely simple-minded treatment of $\tau$'s, which are very common in the LM1 events since staus are light at that cMSSM point. Since the model we're studying in this paper does not predict any events with $\tau$'s (or indeed any SM leptons) at the parton level, we do not expect this issue to affect our results. Indeed, the lepton veto only removes 1-3\% of events in our T-quark event samples, so that even a factor of 2 error in the efficiency of this cut would only have a marginal effect on our results. On the other hand, the excellent agreement of our procedure with the CMS numbers at the level of signature efficiencies indicates that the procedure works very well when purely hadronic+MET events are considered. Since the vast majority of the T-quark events are of this kind, we expect our efficiencies for T-quark searches to be valid to a good approximation. (Of course, ideal agreement is not expected, since our modeling of the detector is quite crude.)
\section{Current and Projected Experimental Reach}
The sensitivity of an experiment to the new physics signal described above is limited by the SM backgrounds. Dominant SM sources of multijet+MET events are pure QCD events with mis-measured jet momenta, $Z$+jets events with $Z\rightarrow\nu\bar{\nu}$, $W$+jets events with $W\rightarrow \ell\nu$ and $\ell$ either not detected or misidentified, and $t\bar{t}$ events with one of the tops decaying leptonically. The CMS collaboration estimated the rate of SM background events passing the analysis cuts, by a combination of Monte Carlo simulations and data-based techniques. For example, an inclusive background estimate for the 35 pb$^{-1}$ data set considered in~\cite{CMS} is
\begin{equation}
{\rm Bg} = 9.4^{+4.8}_{-4.0}~{\rm (stat)}\pm1.0~{\rm (syst)}.
\eeq{bg_rate}
This number was obtained by measuring the rate of multijet+MET events in the ``control region", {\it i.e.} the region with lower $H_T$ than required in the analysis ($H_T\in [250, 350]$ GeV), and extrapolating to higher $H_T$. Alternative estimates for electroweak and $t\bar{t}$ processes were obtained by measuring related processes: for example, the $Z$+jets rate was estimated using the $\gamma$+jets sample. These estimates agree, within errors, with the inclusive estimate quoted above.
The CMS collaboration observed 13 events passing the selection cuts in the data, consistent with the estimate~\leqn{bg_rate}. This measurement can be used to place a limit on the new physics contribution to the rate. A systematic derivation of such a limit should take into account the possibility of contamination of the control region by new physics. The procedure for doing so is rather complicated, and the result is model-dependent. On the other hand, given the shape of the SM $H_T$ distribution and the existing constraints on new physics models under consideration, the contamination is expected to be a minor effect, and at the level of phenomenological analysis of this paper, it is reasonable to ignore it. In this approximation, a new physics contribution after all cuts is limited to at most 13.4 events at 95\% confidence level (CL)~\cite{CMS}, corresponding to a cross section limit of 0.383 pb.
\begin{figure}[t]
\centering%
\includegraphics[width=5in]{excl_plot.pdf}
\caption{Solid black line: Estimated exclusion contour based on the published CMS analysis~\cite{CMS}. Solid red line: Estimated reach for the same analysis with 1 fb$^{-1}$ of data at 7 TeV. Dashed black/red lines indicate the variation of the limits assuming a 25\% uncertainty on the cross section prediction.
Lightly shaded parameter region below the dash-dotted black line is excluded by the D{\O} search at the Tevatron~\cite{LHT_D0}. The region below the dotted line is ruled out by precision electroweak constraints in the LHT model~\cite{LHT_PEW}, but may be allowed in a more general simplified model context. In the darkly shaded band, the jets that pass the analysis cuts are predominantly from initial-state radiation.}
\label{fig:excl}
\end{figure}
The impact of this cross section limit on the parameter space of our model is shown in Fig.~\ref{fig:excl}.
The region to the left and below the solid black line is excluded. Notice that the bound is already stronger than the only published Tevatron bound on this model by the D{\O} collaboration~\cite{LHT_D0}, shown as the lightly shaded region on the figure. (To be fair, it should be noted that the D{\O} analysis was based on a 2.5 fb$^{-1}$ of the Tevatron data and has not been updated.) The LHC search sensitivity can be rapidly improved with larger data sets, expected to be collected during the 2011-12 run. As an example, we present an estimate of the reach expected with 1 fb$^{-1}$ of integrated luminosity (solid red line in Fig.~\ref{fig:excl}). To obtain this estimate, we rescaled the background rate from~\leqn{bg_rate}, assuming that the fractional statistical error will scale as $1/\sqrt{L_{\rm int}}$ while the fractional systematic error will remain unchanged at approximately $10$ \%. We further assumed that the measured rate will coincide with the central value of the background estimate. This procedure yields the expected cross section bound (after all cuts) of 74.6 fb, which was used to compute the reach. According to this simple extrapolation, the analysis becomes systematics-dominated around 1 fb$^{-1}$ of integrated luminosity, and further data does not significantly improve the reach: the expected cross section limit improves only to about 60 fb for 10 fb$^{-1}$.
\begin{figure}[t]
\centering%
\includegraphics[width=3in]{efficiency_fit_log.pdf}
\includegraphics[width=3in]{lht_400_380_pt_comb.pdf}
\caption{Left: Combined analysis cut efficiency for the T-quark signal, as a function of the T-quark/LTP mass difference, for four sample values of the T-quark mass. Right: $p_T$ distribution of the second-hardest jet in each event, at the parton level (dotted histogram) and at the level of {\tt PGS} output (solid histogram). In this simulation, $\tilde{M}=400$ GeV and $M_{\rm LTP}=380$ GeV.}
\label{fig:eff}
\end{figure}
A somewhat surprising feature of Fig.~\ref{fig:excl} is the apparent sensitivity of the analysis in the region of quasi-degenerate T-quark and LTP. Naively, the cut efficiency should become vanishingly small in that region: The jet energy in the T-quark rest frame is proportional to $\tilde{M}-M_{\rm LTP}$, and the T-quarks are not ultrarelativistic in the lab frame (the typical velocity is about 0.5), so the jets should be soft and should fail the acceptance and $H_T$ cuts. In fact, the cut efficiency decreases as expected for $\tilde{M}-M_{\rm LTP}\lower.7ex\hbox{$\;\stackrel{\textstyle>}{\sim}\;$} 100$ GeV, but then approximately flattens out (at about 1\% for the typical T-quark mass in our analysis) at $\tilde{M}-M_{\rm LTP} \lower.7ex\hbox{$\;\stackrel{\textstyle<}{\sim}\;$} 100$ GeV (see the left panel of Fig.~\ref{fig:eff}). Our interpretation is that in this region, the hard jets passing the acceptance and $H_T$ cuts are due to initial-state radiation (ISR). This is further confirmed by comparing parton-level and {\tt Pythia}-level events for a sample parameter point in the quasi-degenerate region, shown on the right panel of Fig.~\ref{fig:eff}. The parton-level sample has no jets passing acceptance cuts, while {\tt Pythia}-level $p_T$ distribution has a long tail extending above the 100 GeV threshold required in this analysis. Of course, the treatment of ISR in {\tt Pythia} is approximate, and large corrections are expected at large $p_T$. In our events the typical ISR jet passing the cut has $p_T\sim 100$ GeV, which is hierarchically smaller than the typical hard scale of the T-quark production process
$\sqrt{\hat{s}}\sim 2\tilde{M}\sim 1$ TeV, so the {\tt Pythia} predictions should not be grossly incorrect. Still, a more detailed treatment (with careful jet matching, or, ideally, with fully differential NLO cross sections in addition to resummation of large logs) is needed to obtain fully reliable efficiency estimates in the quasi-degenerate region. To emphasize this point, in Fig.~\ref{fig:excl} we shaded the region where most of the jets passing the analysis cuts are from ISR.
\section{Conclusions}
In this paper, we estimated the current and near-future LHC sensitivity to ``T-quarks", exotic vector-like quarks carrying a $Z_2$ charge, which decay to an ``LTP", a weakly-interacting massive particle invisible in the detector, and an SM quark. We used the published CMS search in the jets+MET channel~\cite{CMS} as the benchmark analysis. We found that this search, based on the 35 pb$^{-1}$ of data at $\sqrt{s}=7$ TeV collected in 2010, already improves on the existing Tevatron bounds. Significant future improvement can be obtained with 1 fb$^{-1}$ data set, expected to be collected in 2011-12. Beyond 1 fb$^{-1}$, we find that the analysis becomes systematics-limited, assuming that the fractional systematic error on the background estimate reported by CMS (about 10\%) does not improve with more data.
While we used the terminology of the Little Higgs model with T-parity (LTP) to describe the new particles present in our model, the analysis is in fact quite model-independent, and can be applied to any new physics model whose LHC phenomenology can be approximated by the simple Lagrangian in~\leqn{Lint}.
In this sense, our study is in the spirit of the ``simplified model" approach to describing LHC signals of new physics. In particular, our analysis applies (at least approximately) to the UED model, in the limit when KK partner of the gluon is significantly heavier than the KK quarks.
An interesting feature that was noted is that in the region of nearly-degenerate T-quark and LTP, the cut efficiencies do not vanish, due to the presence of ISR jets passing the acceptance and analysis cuts in a non-vanishing fraction of events. This can be especially important in the UED context, where such near-degenerate spectrum is generic (at least in the absence of large brane-localized kinetic terms). A more careful analysis of ISR jets is needed in this region. (See Ref.~\cite{Jay} for similar observations in the context of supersymmetry searches.)
On a more technical note, we found that, reassuringly, the Monte Carlo tools widely used by phenomenologists, including {\tt PGS}, do a decent job of reproducing experimental efficiencies for fully hadronic events. However, we did not find a good agreement with CMS lepton veto efficiencies, most likely due to our overly simplistic treatment of $\tau$'s. This was not an issue in our analysis, since our signal model predicted essentially no leptons, but it does indicate that caution is warranted in general.
To conclude, our analysis shows that the 7 TeV LHC runs in 2010-2012 have a potential to either discover the T-quarks or substantially improve the existing bounds. We encourage the LHC experiments to perform this search. If an excess above the SM background is observed in this channel, we stress that alternative interpretations, such as the model studied here, must be considered before it is attributed to supersymmetry. (For a recent analysis of model discrimination power of the LHC in this channel based on jet $p_T$ and angular distributions, see Ref.~\cite{CMS_MD}.)
\section{Acknowledgements}
We are grateful to Jay Hubisz for many useful discussions, and to David Curtin for help with software. MP is supported by the U.S. National Science Foundation through grant PHY-0757868 and CAREER award PHY-0844667. JS is supported by the Syracuse University College of Arts and Sciences.
|
\section{Introduction}
The array variate random variable up to $2$ dimensions has been studied intensively in \cite{gupta2000matrix} and by many others. For arrays observations of $3,$ $4$ or in general $i$ dimensions a suitable normal probablity model has been recently proposed in \cite{DenizGupta}. In this paper we will generalize the notion of elliptical random variables to the array case.
In Section 2, we first study the algebra of arrays. In Section 3, we introduce the concept of an array variable random variable. In section 4, the density of the elliptical array random variable is provided. Then, in Section 5, we define the array random variable with elliptical density.
\section{Array Algebra}
In this paper we will only study arrays with real elements. We will write $\widetilde{X}$ to say that $\widetilde{X}$ is an array. When it is necessary we can write the dimensions of the array as subindices, e.g., if $\widetilde{X}$ is a $m_1 \times m_2\times m_3 \times m_4$ dimensional array in $R^{m_1\times m_2 \times \ldots \times m_i}$, then we can write $\widetilde{X}_{m_1 \times m_2\times m_3 \times m_4}.$ Arrays with the usual elementwise summation and scalar multiplication operations can be shown to be a vector space.
To refer to an element of an array $\widetilde{X}_{m_1 \times m_2\times m_3 \times m_4},$ we write the position of the element as a subindex to the array name in parenthesis, $(\widetilde{X})_{r_1r_2r_3r_4}.$ If we want to refer to a specific column vector obtained by keeping all but an indicated dimension constant, we indicate the constant dimensions as before but we will put '$:$' for the non constant dimension, e.g., for $\widetilde{X}_{m_1 \times m_2\times m_3 \times m_4},$ $(\widetilde{X})_{r_1r_2:r_4}$ refers to the the column vector $(({X})_{r_1r_21r_4},({X})_{r_1r_22r_4}, \ldots, ({X})_{r_1r_2m_3r_4})'.$
We will now review some basic principles and techniques of array algebra. These results and their proofs can be found in Rauhala \cite{rauhala1974array}, \cite{rauhala1980introduction} and Blaha \cite{blaha1977few}.
\begin{dfn}\emph{Inverse Kronecker product} of two matrices $A$ and $B$ of dimensions $p\times q$ and $r \times s$ correspondingly is written as $A\otimes^i B$ and is defined as $A\otimes^i B=[A(B)_{jk}]_{pr\times qs}=B\otimes A,$ where $'\otimes'$ represents the ordinary Kronecker product.\end{dfn}
The following properties of the inverse Kronecker product are useful:\begin{itemize}\item $\bzero \otimes^i A= A \otimes^i \bzero=\bzero.$ \item $(A_1+A_2)\otimes^i B=A_1\otimes^i B+ A_2 \otimes^i B.$ \item $A \otimes^i (B_1+ B_2)=A \otimes^i B_1+ A \otimes^i B_2.$ \item $\alpha A \otimes^i \beta B= \alpha \beta A \otimes^i B.$ \item $(A_1 \otimes^i B_1)(A_2 \otimes^i B_2)= A_1A_2 \otimes^i B_1B_2.$ \item $(A \otimes^i B)^{-1}=(A^{-1} \otimes^i B^{-1}).$ \item $(A \otimes^i B)^{+}=(A^{+} \otimes^i B^{+}),$ where $A^{+}$ is the Moore-Penrose inverse of $A.$ \item $(A \otimes^i B)^{-}=(A^{-} \otimes^i B^{-}),$ where $A^{-}$ is the $l$-inverse of $A$ defined as $A^{-}=(A'A)^{-1}A'.$\item If $\{\lambda_i\}$ and $\{\mu_j\}$ are the eigenvalues with the corresponding eigenvectors $\{\bx_i\}$ and $\{\by_j\}$ for matrices $A$ and $B$ respectively, then $A\otimes^i B$ has eigenvalues $\{\lambda_i\mu_j\}$ with corresponding eigenvectors $\{\bx_i\otimes^i\by_j\}.$\item Given two matrices $A_{n\times n}$ and $B_{m\times m}$ $|A\otimes^i B|=|A|^m|B|^n,$ $tr(A\otimes^i B)=tr(A)tr(B).$\item $A\otimes^i B=B\otimes A=U_1 A\otimes B U_2,$ for some permutation matrices $U_1$ and $U_2.$
\end{itemize}
It is well known that a matrix equation $$AXB'=C$$ can be rewritten in its monolinear form as \begin{equation}\label{eqmnf}A\otimes^i B vec(X)=vec(C).\end{equation} Furthermore, the matrix equality $$A\otimes^i B XC'=E$$ obtained by stacking equations of the form (\ref{eqmnf}) can be written in its monolinear form as $$(A\otimes^i B \otimes^i C) vec(X)=vec(E).$$ This process of stacking equations could be continued and R-matrix multiplication operation introduced by Rauhala \cite{rauhala1974array} provides a compact way of representing these equations in array form:
\begin{dfn}\emph{R-Matrix Multiplication} is defined elementwise:
$$((A_1)^1 (A_2)^2 \ldots (A_i)^i\widetilde{X}_{m_1 \times m_2 \times \ldots \times m_i})_{q_1q_2\ldots q_i}$$ $$=\sum_{r_1=1}^{m_1}(A_1)_{q_1r_1}\sum_{r_2=1}^{m_2}(A_2)_{q_2r_2}\sum_{r_3=1}^{m_3}(A_3)_{q_3r_3}\ldots \sum_{r_i=1}^{m_i}(A_i)_{q_ir_i}(\widetilde{X})_{r_1r_2\ldots r_i}.$$ \end{dfn}
R-Matrix multiplication generalizes the matrix multiplication (array multiplication in two dimensions)to the case of $k$-dimensional arrays. The following useful properties of the R-Matrix multiplication are reviewed by Blaha \cite{blaha1977few}:
\begin{enumerate}
\item $(A)^1B=AB.$
\item $(A_1)^1(A_2)^2C=A_1CA'_2.$
\item $\widetilde{Y}=(I)^1(I)^2\ldots (I)^i \widetilde{Y}.$
\item \small $((A_1)^1 (A_2)^2\ldots (A_i)^i)((B_1)^1(B_2)^2\ldots (B_i)^i)\widetilde{Y}= (A_1B_1)^1(A_2B_2)^2\ldots(A_iB_i)^i\widetilde{Y}.$ \normalsize
\end{enumerate}
The operator $rvec$ describes the relationship between $\widetilde{X}_{m_1 \times m_2 \times \ldots m_i}$ and its monolinear form $\bx_{m_1m_2\ldots m_i\times 1}.$
\begin{dfn}\label{def:rvec} $rvec( \widetilde{X}_{m_1 \times m_2 \times \ldots m_i})=\bx_{m_1m_2\ldots m_i\times 1}$ where $\bx$ is the column vector obtained by stacking the elements of the array $\widetilde{X}$ in the order of its dimensions; i.e., $(\widetilde{X})_{j_1 j_2 \ldots j_i}=(\bx)_j$ where $j=(j_i-1)n_{i-1}n_{i-2}\ldots n_1+(j_i-2)n_{i-2}n_{i-3}\ldots n_1+\ldots+(j_2-1)n_1+j_1.$\end{dfn}
\begin{thm}\label{rmultvec}Let $\widetilde{L}_{m_1 \times m_2 \times\ldots m_i}=(A_1)^1(A_2)^2\ldots(A_i)^i\widetilde{X}$ where $(A_j)^j$ is an $m_j\times n_j$ matrix for $j=1,2,\ldots,i$ and $\widetilde{X}$ is an $n_1\times n_2\times\ldots\times n_i$ array. Write $\mathbf{l}=rvec(\widetilde{L})$ and $\bx=rvec(\widetilde{X}).$ Then, $\mathbf{l}=A_1\otimes^iA_2\otimes^i\ldots\otimes^i A_i\bx.$ \end{thm}
Therefore, there is an equivalent expression of the array equation in monolinear form.
\begin{dfn}{} The square norm of $\widetilde{X}_{m_1 \times m_2 \times\ldots m_i}$ is defined as $$\|\widetilde{X}\|^2=\sum_{j_1=1}^{m_1}\sum_{j_2=1}^{m_2}\ldots\sum_{j_i=1}^{m_i}((\widetilde{X})_{j_1j_2\ldots j_i})^2.$$ \end{dfn}
\begin{dfn}{} The distance of $\widetilde{X_1}_{m_1 \times m_2 \times\ldots m_i}$ from $\widetilde{X_2}_{m_1 \times m_2 \times\ldots m_i}$ is defined as $\sqrt{\|\widetilde{X_1}-\widetilde{X_2}\|^2}.$ \end{dfn}
\begin{ex} Let $\widetilde{Y}=(A_1)^1 (A_2)^2\ldots (A_i)^i\widetilde{X}+\widetilde{E}.$ Then $\|\widetilde{E}\|^2$ is minimized for $\widehat{\widetilde{X}}=(A_1^{-})^1(A_2^{-})^2\ldots(A_i^{-})^i\widetilde{Y}.$ \end{ex}
\section{Array Variate Random Variables}
Arrays can be constant arrays, i.e. if $(\widetilde{X})_{r_1r_2\ldots r_i}\in\mathbf{R}$ are constants for all $r_j,$ $j=1,2,\ldots,m_j$ and $j=1,2,\ldots, i$ then the array $\widetilde{X}$ is a constant array.
Array variate random variables are arrays with all elements $(\widetilde{X})_{r_1r_2\ldots r_i}\in\mathbf{R}$ random variables. If the sample space for the random outcome $s$ is $\mathbb{S}$, $(\widetilde{X})_{r_1r_2\ldots r_i}=(\widetilde{X}(s))_{r_1r_2\ldots r_i}$ where each of $(\widetilde{X}(s))_{r_1r_2\ldots r_i}$ is a real valued function from $\mathbb{S}$ to $\mathbb{R}.$
If $\widetilde{X}$ is an array variate random variable, its density (if it exists) is a scalar function $f_{\widetilde{X}}(\widetilde{X})$ such that:
\begin{itemize}
\item $f_{\widetilde{X}}(\widetilde{X})\geq 0;$
\item $\int_{\widetilde{X}} f_{\widetilde{X}}(\widetilde{X})d\widetilde{X}=1;$
\item $P(\widetilde{X}\in A)=\int_{A} f_{\widetilde{X}}(\widetilde{X})d\widetilde{X},$ where A is a subset of the space of realizations for $\widetilde{X}.$
\end{itemize}
A scalar function $f_{\widetilde{X},\widetilde{Y}}(\widetilde{X},\widetilde{Y})$ defines a joint (bi-array variate) probability density function if
\begin{itemize}
\item $f_{\widetilde{X},\widetilde{Y}}(\widetilde{X},\widetilde{Y})\geq 0;$
\item $\int_{\widetilde{Y}}\int_{\widetilde{X}} f_{\widetilde{X},\widetilde{Y}}(\widetilde{X},\widetilde{Y})d\widetilde{X}d\widetilde{Y}=1;$
\item $P((\widetilde{X},\widetilde{Y})\in A)=\int \int_{A} f_{\widetilde{X},\widetilde{Y}}(\widetilde{X},\widetilde{Y})d\widetilde{X}d\widetilde{Y},$ where A is a subset of the space of realizations for $(\widetilde{X},\widetilde{Y}).$
\end{itemize}
The marginal probability density function of $\widetilde{X}$ is defined by $$f_{\widetilde{X}}(\widetilde{X})=\int_{\widetilde{Y}} f_{\widetilde{X},\widetilde{Y}}(\widetilde{X},\widetilde{Y})d\widetilde{Y},$$ and the conditional probability density function of $\widetilde{X}$ given $\widetilde{Y}$ is defined by $$f_{\widetilde{X}|\widetilde{Y}}(\widetilde{X}|\widetilde{Y})=\frac{f_{\widetilde{X},\widetilde{Y}}(\widetilde{X},\widetilde{Y})}{f_{\widetilde{Y}}(\widetilde{Y})},$$ where $f_{\widetilde{Y}}(\widetilde{Y})>0.$
Two random arrays $\widetilde{X}$ and $\widetilde{Y}$ are independent if and only if $$f_{\widetilde{X},\widetilde{Y}}(\widetilde{X},\widetilde{Y})= f_{\widetilde{X}}(\widetilde{X})f_{\widetilde{Y}}(\widetilde{Y}).$$
\begin{thm}{} Let $(A_1)^1,$ $(A_2)^2,$ $\ldots,$ $(A_i)^i$ be $m_1,$ $m_2,$ $\ldots,$ $m_i$ dimensional positive definite matrices. The Jacobian $J(\widetilde{X}\rightarrow \widetilde{Z})$ of the transformation $\widetilde{X}=(A_1)^1 (A_2)^2 \ldots (A_i)^i\widetilde{Z}+\widetilde{M}$ is $$(|A_1|^{\prod_{j\neq 1}{m_j}} |A_2|^{\prod_{j\neq 2}{m_j}} \ldots |A_i|^{\prod_{j\neq i}{m_j}})^{-1}.$$
\begin{proof} The result is proven using the equivalence of monolinear form obtained through the $rvec(\widetilde{X})$ and array $\widetilde{X}.$ Let $\widetilde{L}_{m_1 \times m_2 \times\ldots m_i}=(A_1)^1(A_2)^2\ldots(A_i)^i\widetilde{Z}$ where $(A_j)^j$ is an $m_j\times n_j$ matrix for $j=1,2,\ldots,i$ and $\widetilde{X}$ is an $n_1\times n_2\times\ldots\times n_i$ array. Write $\mathbf{l}=rvec(\widetilde{L})$ and $\bz=rvec(\widetilde{Z}).$ Then, $\mathbf{l}=A_1\otimes^iA_2\otimes^i\ldots\otimes^i A_i\bz.$ The result follows from noting that $J(\mathbf{l}\rightarrow \bz)=|A_1\otimes^iA_2\otimes^i\ldots\otimes^i A_i|^{-1},$ and using induction with the rule $|A\otimes^i B|=|A|^m|B|^n$ for ${n\times n}$ matrix $A$ and ${m\times m}$ matrix $B$ to show that \small $|A_1\otimes^iA_2\otimes^i\ldots\otimes^i A_i|^{-1}=(|A_1|^{\prod_{j\neq 1}{m_j}} |A_2|^{\prod_{j\neq 2}{m_j}} \ldots |A_i|^{\prod_{j\neq i}{m_j}})^{-1}.$ \normalsize \end{proof}\end{thm}
\begin{cor}{\label{jacobian}} Let $\widetilde{Z}\sim f_{\widetilde{Z}} (\widetilde{Z}).$ Define $\widetilde{X}=(A_1)^1 (A_2)^2 \ldots (A_i)^i\widetilde{Z}+\widetilde{M}$ where $(A_1)^1, (A_2)^2, \ldots, (A_i)^i$ be $m_1, m_2, \ldots, m_i$ dimensional positive definite matrices. The pdf of $\widetilde{X}$ is given by \small $$f_{\widetilde{X}} (\widetilde{X}; (A_1)^1, (A_2)^2, \ldots, (A_i)^i, \widetilde{M})=\frac{f(A_1^{-1})^1 (A_2^{-1})^2 \ldots (A_i^{-1})^i(\widetilde{X}-\widetilde{M}))}{|A_1|^{\prod_{j\neq 1}{m_j}} |A_2|^{\prod_{j\neq 2}{m_j}} \ldots |A_i|^{\prod_{j\neq i}{m_j}}} .$$ \normalsize \end{cor}
The main advantage in choosing a Kronecker structure is the decrease in the number of parameters.
\section{Array Variate Normal Distribution}
By using the results in the previous section on array algebra, mainly the relationship of the arrays to their monolinear forms described by Definition \ref{def:rvec} , we can write the density of the standard normal array variable.
\begin{dfn}{} If $$\widetilde{Z}\sim N_{m_1 \times m_2\times \ldots \times m_i}(\widetilde{M}=\bzero, \Lambda=I_{m_1m_2\ldots m_i}),$$ then $\widetilde{Z}$ has array variate standard normal distribution. The pdf of $\widetilde{Z}$ is given by \begin{equation}\label{eq:densitystarn}f_{\widetilde{Z}} (\widetilde{Z})=\frac{\exp{(-\frac{1}{2}\|\widetilde{Z}\|^2)}}{(2\pi)^{m_1m_2\ldots m_i/2}}.\end{equation}\end{dfn}
For the scalar case, the density for the standard normal variable $z \in \mathbf{R}^1$ is given as \[\phi_1(z)=\frac{1}{(2\pi)^\frac{1}{2}}exp(-\frac{1}{2}z^2).\] For the $m_1$ dimensional standard normal vector $\bz\in\mathbf{R}^{m_1},$ the density is given by \[\phi_{m_1}(\bz)=\frac{1}{(2\pi)^\frac{m_1}{2}}exp(-\frac{1}{2}\bz'\bz).\] Finally the $m_1\times m_2$ standard matrix variate variable $Z\in\mathbf{R}^{m_1\times m_2}$ has the density \[\phi_{m_1\times m_2}(Z)=\frac{1}{(2\pi)^\frac{m_1m_2}{2}}exp(-\frac{1}{2}trace(Z'Z)).\] With the above definition, we have generalized the notion of normal random variable to the array variate case.
\begin{dfn}{\label{stdnorm}} We write $$\widetilde{X}\sim N_{m_1 \times m_2\times \ldots \times m_i}(\widetilde{M}, \Lambda_{m_1m_2\ldots m_i})$$ if $rvec(\widetilde{X})\sim N_{m_1m_2\ldots m_i}(rvec(\widetilde{M}),\Lambda_{m_1m_2\ldots m_i}).$ Here, $\widetilde{M}$ is the expected value of $\widetilde{X}$, and $\Lambda_{m_1m_2\ldots m_i}$ is the covariance matrix of the ${m_1m_2\ldots m_i}$-variate random variable $rvec(\widetilde{X}).$\end{dfn}
The family of normal densities with Kronecker Delta Covariance Structure are obtained by considering the densities obtained by the location-scale transformations of the standard normal variables. This kind of model is defined in the next.
\begin{thm}{}\label{modkroncov} Let $\widetilde{Z}\sim N_{m_1 \times m_2\times \ldots \times m_i}(\widetilde{M}=\bzero, \Lambda=I_{m_1m_2\ldots m_i}).$ Define $\widetilde{X}=(A_1)^1 (A_2)^2 \ldots (A_i)^i\widetilde{Z}+\widetilde{M}$ where $A_1, A_2,\ldots,A_i$ are non singular matrices of orders $m_1, m_2,\ldots, m_i$. Then the pdf of $\widetilde{X}$ is given by \small \begin{equation}\label{eq:densityarn}\phi(\widetilde{X}; \widetilde{M},A_1,A_2,\ldots A_i)=\frac{\exp{(-\frac{1}{2}\|{(A_1^{-1})^1 (A_2^{-1})^2 \ldots (A_i^{-1})^i(\widetilde{X}-\widetilde{M})}\|^2)}}{(2\pi)^{m_1m_2\ldots m_i/2}|A_1|^{\prod_{j\neq 1}{m_j}} |A_2|^{\prod_{j\neq 2}{m_j}} \ldots |A_i|^{\prod_{j\neq i}{m_j}}}.\end{equation} \normalsize \begin{proof} The result is easily obtained using the density in Definition \ref{stdnorm} with Corollary \ref{jacobian}.\end{proof}\end{thm}
\section{Array Variate Elliptical Distribution}
A spherically symmetric random vector $\bx$ has a density of the form $f(\bx'\bx)$ for some kernel pdf $f(t),$ defined for $t\in \mathbf{R}^+.$
A stochastic representation for the random vector $\bx$ is given by $\bx\sd r\bu,$ where $r$ is a nonnegative random variable with cdf $K(r)$ that is independent of $\bu$ which is uniformly distributed over the unit sphere in $\mathbf{R}^k,$ denoted by $\mathbf{S}^{k},$ where $r\sd \sqrt(\bx'\bx),$ $\bu\sd\frac{\bx}{\sqrt(\bx'\bx)}.$ If both $K'(r)=k(r)\in\mathbf{K}$ the pdf of $r,$ and $f(\bx'\bx)$ the pdf of $\bx$ exist, then they are related as follows: \begin{equation}k(r)=\frac{2\pi^{\frac{k}{2}}}{\Gamma(\frac{k}{2})}r^{k-1}f(r^2).\end{equation}
Some examples follow:
\begin{enumerate}
\item The standard multivariate normal $N_k(\bzero_k, I_k)$ distribution with pdf $$\phi_k(\bx)=\frac{1}{(2\pi)^{k/2}}e^{-\frac{1}{2}\bx'\bx}.$$
\item The multivariate spherical $t$ distribution with $v$ degrees of freedom with density function $$f(\bx)=\frac{\Gamma(\frac{1}{2}(v+k))}{\Gamma(\frac{1}{2}v)(v\pi)^{k/2}}\frac{1}{(1+\frac{1}{v}\bx'\bx)^{(v+k)/2}}.$$
\item When $v=1$ the multivariate $t$ distribution is called spherical Cauchy distribution.
\end{enumerate}
The array variate random variable with spherical contours is defined using the relationship between the array and its monolinear form. This amounts to saying that the density of the array variable $\widetilde{X}$ is spherical if and only if it can be written in the form $f(\|\widetilde{X}\|^2)$ for some kernel pdf $f(t),$ defined for $t\in \mathbf{R}^+.$ The elliptically contoured random variables with unstructured covariance matrices can be obtained by applying a linear transformation to the monolinear form of an array variate spherical random variable.
In general we can define an array variable random variable with unstructured covariance matrix, i.e. we can say $\widetilde{X}$ has array variate distribution if $rvec(\widetilde{X})$ has elliptical distribution. Array variate densities with elliptical contours that have Kronecker delta covariance structure are also easily constructed using Corrolary \ref{jacobian}.
\begin{dfn}\label{elliptical} Let $A_1, A_2,\ldots,A_i$ be non singular matrices with orders $m_1,$ $m_2,$ $\ldots,$ $m_i$ and $\widetilde{M}$ be a $m_1 \times$ $m_2$ $\times \ldots$ $\times m_i$ constant array. Also, let $m=m_1m_2\ldots m_i.$ Then the pdf of an $m_1 \times$ $m_2$ $\times \ldots$ $\times m_i$ elliptically contoured array variate random variable $\widetilde{X}$ with kernel pdf $f(t),$ $t>0$ is \small $$f_{\widetilde{X}} (\widetilde{X}; (A_1)^1, (A_2)^2, \ldots, (A_i)^i, \widetilde{M})=\frac{f(\|(A_1^{-1})^1 (A_2^{-1})^2 \ldots (A_i^{-1})^i(\widetilde{X}-\widetilde{M})\|^2)}{|A_1|^{\prod_{j\neq 1}{m_j}} |A_2|^{\prod_{j\neq 2}{m_j}} \ldots |A_i|^{\prod_{j\neq i}{m_j}}} .$$ \normalsize
\end{dfn}
For example, the following definition provides a generalization of the multivariate $t$ distribution to the array variate case.
\begin{dfn} Let $A_1, A_2,\ldots,A_i$ be non singular matrices with orders $m_1,$ $m_2,$ $\ldots,$ $m_i$ and $\widetilde{M}$ be a $m_1 \times$ $m_2$ $\times \ldots$ $\times m_i$ constant array. Also, let $m=m_1m_2\ldots m_i.$ Then the pdf of an $m_1 \times$ $m_2$ $\times \ldots$ $\times m_i$ array variate t random variable $\widetilde{T}$ with degrees of freedom $v$ is given by \small \begin{equation}\label{eq:densityart}f(\widetilde{T}; \widetilde{M},A_1,A_2,\ldots A_i)=c\frac{(1+\|{(A_1^{-1})^1 (A_2^{-1})^2 \ldots (A_i^{-1})^i(\widetilde{T}-\widetilde{M})\|^2)^{-(v+m)/2}}}{|A_1|^{\prod_{j\neq 1}{m_j}} |A_2|^{\prod_{j\neq 2}{m_j}} \ldots |A_i|^{\prod_{j\neq i}{m_j}}}\end{equation} \normalsize where $c=\frac{(v\pi)^{m/2}\Gamma((v+m)/2)}{\Gamma(v/2)}.$
\end{dfn}
Distributional properties of a array normal variable with density in the form of given by Theorem \ref{modkroncov} or a array variate elliptical random variable with density of the form in (\ref{elliptical}) can obtained by using the equivalent monolinear representation of the random variable. The moments, the marginal and conditional distributions, independence of variates can be studied considering the equivalent monolinear form of the array variable and the well known properties of the multivariate normal random variable and the multivariate elliptical random variable.
\bibliographystyle{plain}
|
\section{Introduction}
In this note, we study the Nonlinear Schr\"odinger Equation (NLS) on a
compact, Riemannian manifold with a periodic elliptic (or stable)
geodesic, which we define and discuss in more detail in Appendix \ref{quasi}. Specifically, we study solutions to
\begin{eqnarray}
\label{nls}
\left\{ \begin{array}{c}
i \partial_t \psi = \Delta_g \psi + \sigma |\psi|^{p} \psi, \\
\psi (x,0) = \psi_0,
\end{array} \right.
\end{eqnarray}
where $\Delta_g$ is the standard (negative semidefinite) Laplace-Beltrami operator and the solution is of the form
\begin{eqnarray*}
\psi (x,t) = e^{-i \lambda t} u(x).
\end{eqnarray*}
To solve the resulting nonlinear elliptic equation, which can be
analyzed using constrained variations, we will use Fermi
coordinates to construct a nonlinear quasimode similar to the one
presented in \cite{Tho} for an arbitrary manifold with a periodic,
elliptic orbit. To do so, we must analyze the metric geometry in a
neighbourhood of a periodic orbit, for which we use the presentation in
\cite{Tubes}. For further references to construction of quasimodes
along elliptic geodesics, see the seminal works Ralston \cite{Ral-beam},
Babi{\v{c}} \cite{Bab-geo}, Guillemin-Weinstein \cite{GuWe-geo}, Cardoso-Popov \cite{CaPo-quasi}, as well as the
thorough survey on spectral theory by Zelditch
\cite{Zel-survey}. For experimental evidence of the existence of such
Gaussian beam type solutions in nonlinear optics as solutions to
nonlinear Maxwell's equations, see the recent paper by Schultheiss et
al \cite{optics}.
Henceforward, we assume that $(M,g)$ is a compact Riemannian manifold
of dimension $d \geq 2$ without boundary, and that it admits an elliptic periodic geodesic $\Gamma \subset M$. For $\beta>0$, we introduce the notation
\begin{equation*}
U_\beta= \{ x \in M : \text{dist}\,_g(x, \Gamma) < \beta \}.
\end{equation*}
Let $L>0$ be the period of $\Gamma$. Throughout the paper, each time we refer to the small parameter $h>0$, this means that $h$ takes the form
\begin{equation}\label{formeh}
h=\frac L{2\pi N},
\end{equation}
for some large integer $N\in {\mathbb N}$.
Our first result is the existence, in the case of a smooth nonlinearity, of quasimodes which are highly
concentrated near the elliptic periodic geodesic $\Gamma$. More precisely :
\begin{theorem}
\label{T:0}
Let $(M,g)$ be a compact Riemannian manifold of dimension $d \geq 2$, without boundary and which has an elliptic periodic geodesic. Let $p$ be an even integer, let $s\geq 0$ and assume that
\begin{equation*}
p \big(\frac{d-1}4-s\big)<1.
\end{equation*}
Then for $h \ll 1$ sufficiently small, and any $\delta>0$, there exists
$\varphi_h(x) \in H^{\infty}(M)$ satisfying
\begin{itemize}
\item[(i)] Frequency localization : For all $r\geq 0$, there exist $C_{1}, C_{2}$ independent of $h$ such that
\begin{equation*}
C_{1} h^{s-r} \leq \| \varphi_h \|_{H^r(M)} \leq C_{2} h^{s-r}.
\end{equation*}
\item[(ii)] Spatial localization near $\Gamma$ :
\begin{equation*}
\|\varphi_h \|_{H^s(M \setminus U_{h^{1/2-\delta}})} = {\mathcal O}(h^{\infty}) \| \varphi_h \|_{H^s(M)}.
\end{equation*}
\item[(iii)] Nonlinear quasimode : There exists $\lambda(h) \in \mathbb{R}$ so that $\varphi_{h}$ satisfies the equation
\begin{equation*}
-\Delta_{g} \varphi_h=\lambda(h)\varphi_{h} -\sigma | \varphi_h|^{p}\varphi_h+\mathcal{O}(h^{\infty}).
\end{equation*}
\end{itemize}
\end{theorem}
The quasimode $\varphi_{h}$ and the nonlinear eigenvalue $\lambda(h)$ will be constructed thanks to a WKB method, this will give a precise description of these objects (see Section \ref{sec:nthorder}). For instance, $\lambda(h)$ reads $\lambda(h)=h^{-2}-E_{0}h^{-1}+o(h^{-1})$, for some $E_{0}\in \mathbb{R}$.
As a consequence of Theorem \ref{T:0}, and under the same assumptions, we can state :
\begin{theorem}
\label{T:1}
Consider the function $\varphi_{h}$ given by Theorem \ref{T:0}. There exists $c_{0}>0$, such that if we denote by $\displaystyle T_{h}=c_{0}h^{p(\frac{d-1}4-s)}\ln(\frac1h)$, the solution $u_{h}$ of the Cauchy problem \eqref{nls} with initial condition $\varphi_{h}$ satisfies
\[
\| u_h\|_{L^\infty ([0,T_h]; H^s(M \setminus
U_{h^{1/2-\delta}})) } \leq C h^{(d+1)/4}.
\]
\end{theorem}
To construct a quasimode with $\mathcal{O}(h^{\infty})$ error, we need that the nonlinearity is of polynomial type, that is $p\in 2{\mathbb N}$. However, in the special case where $d=2$ and $s=0$, we are able to obtain a result for any $0<p<4$, which is a weaker version of Theorem \ref{T:1}.
\begin{theorem}
\label{T:2}
Let $(M,g)$ be a compact Riemannian surface without boundary which admits an elliptic periodic geodesic and let $0<p<4$. Then for $h \ll 1$ sufficiently small, and any $\delta,\epsilon>0$, there exist
$\varphi_h(x) \in H^{\infty}(M)$ and $\nu>0$ satisfying
\begin{itemize}
\item[(i)] Frequency localization : For all $r\geq 0$, there exist $C_{1}, C_{2}$ independent of $h$ such that
\begin{equation*}
C_{1} h^{-r} \leq \| \varphi_h \|_{H^r(M)} \leq C_{2} h^{-r}.
\end{equation*}
\item[(ii)] Spatial localization near $\Gamma$ :
\begin{equation*}
\|\varphi_h \|_{L^{2}(M \setminus U_{h^{1/2-\delta}})} = {\mathcal O}(h^{\infty}) \| \varphi_h \|_{L^{2}(M)},
\end{equation*}
\end{itemize} such that the corresponding solution $u_h$ to \eqref{nls}
satisfies
\[
\| u_h \|_{L^\infty([0,T_h] ; L^{2}(M \setminus
U_{h^{1/2-\delta}})) } \leq C h^{\nu},
\]
for $T_{h}=h^{p}$ in the case $p\in (0,4)\backslash \{1\}$, and $T_{h}=h^{1+\epsilon}$
in the case $p=1$.
\end{theorem}
\begin{rema}
In our current notations, the $\dot{H}^1$ critical exponent for \eqref{nls} is
\[
p = \frac{4}{d-2} = \frac{4}{n-1},
\]
where $n$ is the dimension of the geodesic normal hypersurface to
$\Gamma$. The $L^2$ critical exponent is $p = 4/d$, and the $\dot{H}^{1/2}$ critical exponent is $p = 4/(d-1)$.
In general we construct highly concentrated
solutions along an elliptic orbit, which is effectively a $d-1$
dimensional soliton. Since stable soliton solutions exist on
$\mathbb{R}^d$ for monomial Schr\"odinger of the type in \eqref{nls} precisely
for $p < \frac{4}{d}$, this numerology matches well that required for the expected local
existence of stable nonlinear states on these lower dimensional
manifolds.
In addition, solving \eqref{nls} on compact manifolds with the power
$p = 4/(d-2)$ plays an important role in the celebrated Yamabe
problem, see \cite{SchYau}. The authors hope that the
techniques here can give insight into the energy minimizers of such a problem
related to the geometry.
\end{rema}
\begin{rema}
In \cite{ChMa-Hyp}, the existence of nonlinear bound states on
hyperbolic space, $\mathbb{H}^d$, was explored. In that case, it was
found that the geometry made compactness arguments at $\infty$ rather
simple, hence the only driving force for the existence of a nonlinear
bound state was the local behavior of the nonlinearity. Work in progress with the second author,
third author, Michael Taylor and Jason Metcalfe is attempting to
generalize this observation to any rank one symmetric space. In this note,
we find nonlinear quasimodes based purely on the local geometry, where
here the nonlinearity becomes a lower order correction, which is a
nicely symmetric result.
\end{rema}
\begin{rema}
In a private conversation with Nicolas Burq, he has pointed out that
in settings where one assumes radial symmetry of the manifold, it is
possible to construct exact nonlinear bound states.
\end{rema}
\begin{Notations}
In this paper $c$, $C$ denote constants the value of which may change from line to line. We use the notations $a\sim b$,
$a\lesssim b$ if $\frac1C b\leq a\leq Cb$, $a\leq Cb$ respectively.
\end{Notations}
{\sc Acknowledgments.} P. A. was supported by an NSF joint institutes postdoctoral
fellowship (NSF grant DMS-0635607002) and a postdoctoral fellowship
of the Foundation Sciences Math\'ematiques de Paris. J.L.M. was
supported in part by an NSF postdoctoral fellowship (NSF grant DMS-0703531) at Columbia
University and a Hausdorff Center Postdoctoral Fellowship at the
University of Bonn. He would also like to thank both the Institut Henri
Poincar\'e and the Courant Institute for being generous hosts during
part of the completion of this work. H.C. was supported in part by
NSF grant DMS-0900524, and would also like to thank the University of
Bonn, Columbia University, the Courant Institute, and the Institut Henri Poincar\'e for their hospitality during
part of this work.
L.T. was supported in part by the
grants ANR-07-BLAN-0250 and ANR-10-JCJC 0109 of the Agence Nationale
de la Recherche. In addition, the authors with to thank Rafe Mazzeo
and especially Colin Guillarmou for helpful conversations throughout
the preparation of this work.
\section{Outline of Proof}
The proof of Theorem \ref{T:1} is to solve approximately the
associated stationary equation. That is, by separating variables in the $t$ direction, we
write
\[
\psi (x,t) = e^{-i \lambda t} u(x),
\]
from which we get the stationary equation
\[
(\lambda+ \Delta_g ) u = \sigma |u|^{p} u.
\]
For $h\ll1$ of the form \eqref{formeh}, the construction in the proof
finds a family of functions
$$u_h(x) = h^{-(d-1)/4} g(h^{-1/2} x)$$ such that $g$ is rapidly decaying away from
$\Gamma$, ${\mathcal C}^\infty$, $g$ is normalized in $L^2$, and
\[
(\lambda+ \Delta_g ) u_h = \sigma |u_h|^{p} u_h +
Q(u_h),
\]
where $\lambda\sim h^{-2}$ and where the error $Q(u_h)$ is expressed by a truncation of an
asymptotic series similar to that in \cite{Tho} and is of lower order in $h$.
The point is that Theorem \ref{T:1} (as well as Theorem \ref{T:2}) is an improvement over the trivial approximate
solution.
It is well known that there exist quasimodes for the linear equation
localized near $\Gamma$ of the form
\[
v_h(x) = h^{-(d-1)/4} e^{is/h }f(s,h^{-1/2}x), \,\, (0<h\ll1),
\]
with $f$ a function rapidly decaying away from $\Gamma$, and $s$
a parametrization around $\Gamma$, so that
$v_h(x)$ satisfies
\[
(\lambda+\Delta_g )v_h = {\mathcal O}(h^{\infty}) \| v_h \|
\]
in any seminorm, see \cite{Ral-beam}. Then
\[
(\lambda+ \Delta_g ) v_h= \sigma |v_h|^{p} v_h +
Q_2(v_h),
\]
where the error $Q_2(v_\lambda) = | v_h|^p v_h$ satisfies
\[
\| Q_2(v_h) \|_{\dot{H}^s} = {\mathcal O}(h^{-s - p(d-1)/4} ).
\]
\section{A toy model}
\label{S:toy}
In this section we consider a toy model in two dimensions, and we give an idea of the proof of Theorem \ref{T:2}. For simplicity, we moreover assume that $2\leq p<4$. As it is a
toy model, we will not dwell on error analysis, and instead make
Taylor approximations at will without remarking on the error terms.
Consider the manifold
\[
M = {\mathbb R}_x / 2 \pi {\mathbb Z} \times {\mathbb R}_\theta / 2 \pi
{\mathbb Z},
\]
equipped with a metric of the form
\[
ds^2 = d x^2 + A^2(x) d \theta^2,
\]
where $A \in {\mathcal C}^\infty$ is a smooth function, $A \geq \epsilon>0$ for some
$\epsilon$.
From this metric, we get the volume form
\[
d \text{Vol} = A(x) dx d \theta,
\]
and the Laplace-Beltrami operator acting on $0$-forms
\[
\Delta_g f = (\partial_x^2 + A^{-2} \partial_\theta^2 + A^{-1}
A' \partial_x) f.
\]
We observe that we can conjugate
$\Delta_g$ by an isometry of metric spaces and separate variables so
that spectral analysis of $\Delta_g$ is equivalent to a one-variable
semiclassical problem with potential. That is, let $S : L^2(X, d
\text{Vol}) \to L^2(X, dx d \theta)$ be the isometry given by
\[
Su(x, \theta) = A^{1/2}(x) u(x, \theta).
\]
Then $\widetilde{\Delta} = S \Delta S^{-1}$ is essentially self-adjoint on $L^2 (
X, dx d \theta)$ with mild assumptions on $A$. A simple calculation gives
\[
-\widetilde{\Delta} f = (- \partial_x^2 - A^{-2}(x) \partial_\theta^2 + V_1(x) )
f,
\]
where the potential
\[
V_1(x) = \frac{1}{2} A'' A^{-1} - \frac{1}{4} (A')^2 A^{-2}.
\]
We are interested in the nonlinear Schr\"odinger equation \eqref{nls},
so we make a separated Ansatz:
\[
u_\lambda(t, x, \theta) = e^{-i t \lambda} e^{i k \theta} \psi(x),
\]
where $k \in {\mathbb Z}$ and $\psi$ is to be determined (depending on both
$\lambda$ and $k$).
Applying the Schr\"odinger operator (with
$\widetilde{\Delta}$ replacing $\Delta$) to
$u_\lambda$ yields the equation
\[
(D_t + \widetilde{\Delta}) e^{i t \lambda} e^{i k \theta} \psi(x) = (\lambda +
\partial_x^2 - k^2 A^{-2}(x) + V_1(x) ) e^{i t \lambda} e^{i k \theta}
\psi(x) = \sigma | \psi|^p e^{i t \lambda} e^{i k \theta} \psi(x),
\]
where we have used the standard notation $D =- i \partial$.
We are interested in the behaviour of a solution or approximate
solution near an elliptic periodic geodesic, which occurs at a maximum
of the function $A$. For simplicity, let
\[
A(x) = \sqrt{ (1 + \cos^2(x))/2 },
\]
so that in a neighbourhood of $x = 0$, $A^2 \sim 1 - x^2$ and $A^{-2}
\sim 1 + x^2$. The function $V_1(x) \sim \text{const.}$ in a
neighbourhood of $x = 0$, so we will neglect $V_1$. If we assume
$\psi(x)$ is localized near $x=0$, we get the stationary reduced equation
\[
(-\lambda + \partial_x^2 - k^2 ( 1 + x^2 ) ) \psi =- \sigma | \psi |^p
\psi .
\]
Let $h = |k|^{-1}$ and use the rescaling operator $T \psi(x) = T_{h,0}
\psi(x) = h^{-1/4} \psi(h^{-1/2} x )$ (see Lemma
\ref{L:T-lemma} below with $n=1$) to conjugate:
\[
T^{-1} (-\lambda + \partial_x^2 - k^2 ( 1 + x^2 ) ) T T^{-1} \psi =
T^{-1} (\sigma | \psi |^p
\psi )
\]
or
\[
( -\lambda + h^{-1} \partial_x^2 - k^2(1 + h x^2 ) \varphi = \sigma h^{-p/4} |
\varphi |^p \varphi,
\]
where $\varphi = T^{-1} \psi$. Let us now multiply by $h$:
\[
( - \partial_x^2 + x^2 - E ) \varphi = \sigma h^{q} | \varphi |^p \varphi,
\]
where
\[
E := \frac{1 - \lambda h^2}{h}
\]
and
\[
q: = 1 - p \frac{d-1}{4} = 1 - \frac{p}{4}.
\]
Observe the range restriction on $p$ is precisely so that
\[
0 < q \leq 1/2.
\]
We make a WKB type Ansatz, although in practice we will only take
two terms (more is possible if the nonlinearity is smooth as in the context of Theorems \ref{T:0} and \ref{T:1}):
\[
\varphi =\varphi_0 + h^q \varphi_1, \,\,\, E =E_0 + h^q E_1.
\]
The first two equations are
\begin{align*}
& h^0: \quad ( - \partial_x^2 + x^2 - E_0 ) \varphi_0 = 0, \\
& h^{q}: \quad ( - \partial_x^2 + x^2 - E_0 -h^q E_1) \varphi_1 = E_1 \varphi_0 +
\sigma | \varphi_0 |^p \varphi_0.
\end{align*}
Observe we have included the $h^q E_1 \varphi_1$ term on the left hand
side.
The first equation is easy:
\[
\varphi_0(x) = e^{-x^2/2}, \,\,\, E_0 = 1.
\]
For the second equation, we want to project away from $\varphi_0$ which
is in the kernel of the operator on the left hand side. That is,
choose $E_1$ satisfying
\[
\left\langle E_1 \varphi_0 + \sigma | \varphi_0 |^p \varphi_0 , \varphi_0 \right\rangle = 0,
\]
so that the right hand side is in $\varphi_0^\perp \subset L^2$.
Then since the spectrum of the one-dimensional harmonic oscillator is
simple (and of the form $(2m+1)$, $m \in {\mathbb Z}$), the operator $(
- \partial_x^2 + x^2 - E_0 -h^q E_1)$ is invertible on
$\varphi_0^\perp \subset L^2 $ with inverse bounded by $(2 - h^q)^{-1}$.
Hence for $h>0$ sufficiently small, we can find
$\varphi_1 \in L^2$ satisfying the second equation above (here we have
used that $\varphi_0$ is Schwartz with bounded $H^s$ norms). Further,
since $\varphi_0$ is Schwartz and strictly positive, so is $|\varphi_0 |^p \varphi_0$, so by
propagation of singularities, $\varphi_1$ is also Schwartz. In particular, both
$\varphi_0$ and $\varphi_1$ are rapidly decaying away from $x = 0$.
Let now $\psi(x) = T(\varphi_0(x) + h^{q} \varphi_1(x) )$, and observe
that by the above considerations, $\psi(x)$ is ${\mathcal O}(h^\infty)$ in any
seminorm, outside an $h^{1/2 - \delta}$ neighbourhood of $x = 0$. Let
$u = e^{i t \lambda} e^{i k \theta} \psi(x)$ so that $\| u \|_{L^2(dx
d \theta)} \sim 1$, and $u$ is ${\mathcal O}(h^\infty)$ outside an $h^{1/2 -
\delta}$ neighbourhood of $x = 0$. Furthermore, $u$ satisfies the
equation (again neglecting smaller terms)
\begin{align*}
(D_t + \widetilde{\Delta}) u & = h^{-1} T h T^{-1} (-\lambda +
\partial_x^2 - k^2 A^{-2}(x) ) T T^{-1} u\\
& = h^{-1} T (\sigma h^{q} | T^{-1} u|^p T^{-1} u + {\mathcal O}(h |
T^{-1} u |^{p+1}) \\
& = \sigma | u |^{p} u + Q,
\end{align*}
where $Q$ satisfies the pointwise bound
\[
Q = {\mathcal O}( h^{q} | u |^{p+1}).
\]
We now let $\tilde{u}$ be the actual solution to \eqref{nls} with the
same initial profile:
\[
\begin{cases}
(D_t + \Delta ) \tilde{u} = \sigma | \tilde{u} |^p \tilde{u} \\
\tilde{u} |_{t = 0 } = e^{ik \theta} \psi(x).
\end{cases}
\]
Set $T_{h}=h^{p}$, then with the Strichartz estimates of Burq-G\'erard-Tzvetkov \cite{BGT-comp} we prove (see Proposition \ref{properreur2}) that there exists $\nu>0$ so that
\begin{equation*}
\| u - \tilde{u} \|_{L^\infty([0,T_h];L^{2}(M))} \leq Ch^{\nu},
\end{equation*}
and therefore we can compute:
\begin{align*}
\| \tilde{u} \|_{L^\infty([0,T_h]; L^2( M \setminus U_{h^{1/2-\delta
}}))} & \leq \| u \|_{L^\infty([0,T]; L^2( M \setminus U_{h^{1/2-\delta}}))} + \| \tilde{u} - u \|_{L^\infty([0,T]; L^2( M \setminus U_{h^{1/2-\delta}}))} \\
& = {\mathcal O}(h^{\infty}) + {\mathcal O}(h^{\nu})={\mathcal O}(h^{\nu}),
\end{align*}
which gives the result.
\begin{rema}
It is very important to point out that the sources of additional error
in this heuristic exposition have been ignored, and indeed, to apply a
similar idea in the general case, a microlocal reduction to a {\it
tubular neighbourhood} of $\Gamma$ in cotangent space is employed.
The function $\varphi_0$ is no longer so simple, and the nonlinearity $|
\varphi_0 |^p \varphi_0$ is no longer necessarily smooth. Because of this,
the semiclassical wavefront sets are no longer necessarily compact, so
a cutoff in frequency results in a {\it fixed loss}.
\end{rema}
\begin{rema}
We remark that the Strichartz estimates from \cite{BGT-comp} are sharp
on the sphere ${\mathbb S}^d$ for a particular Strichartz pair, but this
is not necessarily true on a generic Riemannian manifold. See
\cite{BSS-Sch} for a thorough discussion of this fact.
\end{rema}
\section{Preliminaries}
\label{preliminaries}
\subsection{Symbol calculus on manifolds}
This section contains some basic definitions and results from semiclassical and microlocal
analysis which we will be using throughout the paper. This is essentially standard, but we
include it for completeness. The techniques presented have been
established in multiple references, including but not limited to the
previous works of the second author \cite{Chr-NC,Chr-QMNC},
Evans-Zworski \cite{EvZw}, Guillemin-Sternberg
\cite{GuSt-Geom,GuSt-Semi}, H\"ormander \cite{Hor-v1,Hor-v2},
Sj\"ostrand-Zworski \cite{SjZw-mono}, Taylor \cite{Tay-pdo},
and many more.
To begin we present results from \cite{EvZw}, Chapter $8$ and Appendix
$E$. Let $X$ be a smooth, compact manifold. We will be operating on half-densities,
\begin{eqnarray*}
u(x)|dx|^{\frac{1}{2}} \in {\mathcal C}^\infty\left(X, \Omega_X^{\frac{1}{2}}\right),
\end{eqnarray*}
with the informal change of variables formula
\begin{eqnarray*}
u(x)|dx|^{\frac{1}{2}} = v(y)|dy|^{\frac{1}{2}}, \,\, \text{for}\,\, y = \kappa(x)
\Leftrightarrow v(\kappa(x))|\kappa'(x)|^{\frac{1}{2}} = u(x),
\end{eqnarray*}
where $|\kappa'(x)|$ has the canonical interpretation as the Jacobian $| \det (\partial
\kappa) |$.
By symbols on $X$ we mean the set
\begin{eqnarray*}
\lefteqn{{\mathcal S}^{k,m} \left(T^*X, \Omega_{T^*X}^{\frac{1}{2}} \right):= } \\
& = &\left\{ a \in {\mathcal C}^\infty(T^*X \times (0,1], \Omega_{T^*X}^{\frac{1}{2}}): \left| \partial_x^\alpha
\partial_\xi^\beta a(x, \xi; h) \right| \leq C_{\alpha \beta}h^{-m} \langle \xi \rangle^{k - |\beta|} \right\}.
\end{eqnarray*}
Essentially this is interpreted as saying that on each coordinate
chart, $U_\alpha$, of $X$, $a \equiv a_\alpha$ where is $a_\alpha \in
{\mathcal S}^{k,m}$ in a standard symbol on ${\mathbb R}^d$.
There is a corresponding class of pseudodifferential operators $\Psi_h^{k,m}(X, \Omega_X^{\frac{1}{2}})$
acting on half-densities defined by the local formula (Weyl calculus)
in ${\mathbb R}^n$:\footnote{We use the semiclassical, or rescaled {\it
unitary} Fourier transform throughout: \[\mathcal{F}_h u( \xi) =
(2 \pi h)^{-d/2} \int e^{-i \langle x, \xi \rangle/h} u(x) d x.\]}
\begin{eqnarray*}
\mathrm{Op}\,_h^w(a)u(x) = \frac{1}{(2 \pi h)^n} \int \int a \left( \frac{x + y}{2}, \xi; h \right)
e^{i \langle x-y, \xi \rangle / h }u(y) dy d\xi.
\end{eqnarray*}
We will occasionally use the shorthand notations $a^w := \mathrm{Op}\,_h^w(a)$ and $A:=\mathrm{Op}\,_h^w(a)$ when
there is no ambiguity in doing so.
We have the principal symbol map
\begin{eqnarray*}
\sigma_h : \Psi_h^{k,m} \left( X, \Omega_X^{\frac{1}{2}} \right) \to {\mathcal S}^{k,m} \left/ {\mathcal S}^{k-1, m-1}
\left(T^*X, \Omega_{T^*X}^{\frac{1}{2}} \right) \right.,
\end{eqnarray*}
which gives the left inverse of $\mathrm{Op}\,_h^w$ in the sense that
\begin{eqnarray*}
\sigma_h \circ \mathrm{Op}\,_h^w: {\mathcal S}^{k,m} \to {\mathcal S}^{k,m}/{\mathcal S}^{k-1, m-1}
\end{eqnarray*}
is the natural projection. Acting on half-densities in the Weyl calculus, the principal
symbol is actually well-defined in ${\mathcal S}^{k,m} / {\mathcal S}^{k-2,m-2}$, that is, up to ${\mathcal O}(h^2)$ in
$h$ (see, for example \cite{EvZw}, Appendix E).
We will use the notion of semiclassical wave front sets for
pseudodifferential operators on manifolds, see H\"ormander
\cite{Hor-v2}, \cite{GuSt-Geom}.
If $a \in {\mathcal S}^{k,m}(T^*X, \Omega_{T^*X}^{\frac{1}{2}})$, we define the singular support or essential support for $a$:
\begin{eqnarray*}
\text{ess-supp}\,_h a \subset T^*X \bigsqcup {\mathbb S}^*X,
\end{eqnarray*}
where ${\mathbb S}^*X = (T^*X \setminus \{0\}) / {\mathbb R}_+$ is the cosphere bundle (quotient taken with
respect to the usual multiplication in the fibers), and the union is
disjoint. The $\text{ess-supp}\,_h a$ is defined using complements:
\begin{eqnarray*}
\lefteqn{\text{ess-supp}\,_h a := } \\
&& \complement \left\{ (x, \xi) \in T^*X : \exists \epsilon >0, \,\,\, (\partial_x^\alpha
\partial_\xi^\beta a) (x', \xi') = {\mathcal O}(h^\infty), \,\,\, d(x, x') + |\xi - \xi'| < \epsilon \right\} \\
&& \bigcup \complement \{ (x, \xi) \in T^*X \setminus 0 : \exists \epsilon > 0, \,\,\, (\partial_x^\alpha
\partial_\xi^\beta a) (x', \xi') = {\mathcal O} (h^\infty \langle \xi \rangle^{-\infty}), \\
&& \quad \quad \quad d(x, x') + 1 / |\xi'| + | \xi/ |\xi| - \xi' /
|\xi'| | < \epsilon \} / {\mathbb R}_+.
\end{eqnarray*}
We then define the wave front set of a pseudodifferential operator $A \in \Psi_h^{k,m}( X, \Omega_X^{\frac{1}{2}} )$:
\begin{eqnarray*}
\mathrm{WF}_h(A) : = \text{ess-supp}\,_h(a), \,\,\, \text{for} \,\,\, A = \mathrm{Op}\,_h^w(a).
\end{eqnarray*}
If $u(h)$ is a family of distributional half-densities, $u \in {\mathcal C}^\infty( (0, 1]_h, \mathcal{D}'(X,
\Omega_X^{\frac{1}{2}}))$, we say $u(h)$ is $h$-{\it tempered} if there is an $N_0$ so that $h^{N_0}u$ is bounded in $\mathcal{D}'(X, \Omega_X^{\frac{1}{2}})$.
If $u = u(h)$ is an $h$-tempered family of distributions, we can
define the semiclassical wave front set of $u$, again by complement:
\begin{eqnarray*}
\lefteqn{\mathrm{WF}_h (u) := } \\
&& \complement \{(x, \xi) : \exists A \in \Psi_h^{0,0}, \,\, \text{with} \,\, \sigma_h(A)(x,\xi) \neq 0, \,\, \\
&& \quad \text{and} \,\,Au \in h^\infty {\mathcal C}^\infty((0,1]_h, {\mathcal C}^\infty(X, \Omega_X^{\frac{1}{2}})) \}.
\end{eqnarray*}
For $A = \mathrm{Op}\,_h^w(a)$ and $B = \mathrm{Op}\,_h^w(b)$, $a \in {\mathcal S}^{k,m}$, $b \in {\mathcal S}^{k',m'}$ we have the composition
formula (see, for example, \cite{DiSj})
\begin{eqnarray}
\label{Weyl-comp}
A \circ B = \mathrm{Op}\,_h^w \left( a \# b \right),
\end{eqnarray}
where
\begin{eqnarray}
\label{a-pound-b}
{\mathcal S}^{k + k', m+m'} \ni a \# b (x, \xi) := \left. e^{\frac{ih}{2} \omega(Dx, D_\xi; D_y, D_\eta)}
\left( a(x, \xi) b(y, \eta) \right) \right|_{{x = y} \atop {\xi = \eta}} ,
\end{eqnarray}
with $\omega$ the standard symplectic form.
We record some useful Lemmas.
\begin{lemma}
Suppose $(x_0, \xi_0) \notin \mathrm{WF}_h (u)$. Then $\forall b \in {\mathcal C}^\infty_c(T^*
{\mathbb R}^n)$ with support sufficiently close to $(x_0, \xi_0)$ we have
\[
b(x, hD) u = {\mathcal O}_{\mathcal{S}} ( h^\infty ).
\]
\end{lemma}
Here $ {\mathcal O}_{\mathcal{S}} ( h^\infty )$ means ${\mathcal O}(h^\infty)$ in any
Schwartz semi-norm. The proof of this Lemma follows similarly to that
of Theorem $8.9$ in \cite{EvZw}.
\begin{theorem}
\label{T:wf-properties}
\begin{itemize}
\item[(i)] Suppose $a \in {\mathcal S}(m)$ and $u(h)$ is $h$-tempered. Then
\[
\mathrm{WF}_h (a^w u) \subset \mathrm{WF}_h (u) \cap \text{ess-supp}\,_h (a).
\]
\item[(ii)] If $a \in {\mathcal S}(m)$ is real-valued, then also
\[
\mathrm{WF}_h ( u ) \subset \mathrm{WF}_h (a^w u) \cup \overline{\complement \{\text{ess-supp}\,_h a \}}.
\]
\end{itemize}
\end{theorem}
\begin{proof}
Assertion $1$ is straightforward. The proof of assertion $2$ is
standard, however we present it here so we can use it for the
analogous result for the blown-up wavefront set.
We will show if $a(x_0, \xi_0) \neq 0 $ and $a^w u =
{\mathcal O}_{L^2}(h^\infty)$ then there exists $b$, $b(x_0, \xi_0) \neq 0$ so
that $b^w u = {\mathcal O}_{L^2}(h^\infty)$. There exists a neighbourhood $U
\ni (x_0, \xi_0)$, a real-valued function $\chi$, and a positive
number $\gamma >0$ such that $\mathrm{supp}\, \chi \cap U = \emptyset$ and
\[
| a + i \chi| \geq \gamma \text{ everywhere}.
\]
Then $P = a^w + i \chi^w$ has an approximate left inverse $c^w$ so
that
\[
c^w P = \,\mathrm{id}\, + R^w,
\]
where $R^w = {\mathcal O}_{L^2 \to L^2}(h^\infty)$. Choose $b \in {\mathcal S}$ so that
$\mathrm{supp}\, (b) \subset U$ and $b(x_0, \xi_0) \neq 0$. Then $b^w \chi^w =
{\mathcal O}(h^\infty)$ as an operator on $L^2$. Hence
\begin{align*}
b^w u = & b^w c^w P u - b^w R^w u \\
= & b^w c^w a^w u + i b^w c^w \chi^w u - b^w R^w u \\
= & {\mathcal O}(h^\infty),
\end{align*}
where we have used the Weyl composition formula to conclude
$\text{ess-supp}\,_h( c \# \chi) \cap \mathrm{supp}\, b = \emptyset$.
\end{proof}
\subsection{Exotic symbol calculi}
Following ideas from \cite{SjZw-mono}, since rescaling often means dealing with symbols with bad decay
properties, we introduce weighted wave front sets as well. Let us
first recall the non-classical symbol classes:
\begin{eqnarray*}
\lefteqn{{\mathcal S}^{k,m}_{\delta, \gamma} \left(T^*X, \Omega_{T^*X}^{\frac{1}{2}} \right):= } \\
& = &\left\{ a \in {\mathcal C}^\infty(T^*X \times (0,1], \Omega_{T^*X}^{\frac{1}{2}}): \left| \partial_x^\alpha
\partial_\xi^\beta a(x, \xi; h) \right| \leq C_{\alpha
\beta}h^{-\delta|\alpha| -\gamma | \beta|-m} \langle \xi \rangle^{k - |\beta|} \right\}.
\end{eqnarray*}
That is, symbols which lose $\delta$ powers of $h$ upon
differentiation in $x$ and $\gamma$ powers of $h$ upon differentiation
in $\xi$. Note the simplest way to achieve this is to take a
symbol $a(x, \xi) \in {\mathcal S}$ and rescale $(x, \xi) \mapsto (h^{-\delta} x, h
^{-\gamma} \xi)$, which then localizes on a scale $h^{\delta + \gamma}$ in
phase space. We thus make the restriction that $0 \leq \delta, \gamma \leq
1$, $0 \leq \delta + \gamma \leq 1$, and to gain powers of $h$ by integrations by parts, we usually
also require $\delta + \gamma < 1$. We can define wavefront sets using
${\mathcal S}_{\delta, \gamma} = {\mathcal S}^{0,0}_{\delta,\gamma}$ symbols, but the localization of the wavefront sets is stronger.
\begin{definition}
If $u(h)$ is $h$-tempered and $0 \leq \delta , \gamma \leq 1$, $0
\leq \delta + \gamma \leq 1$,
\begin{align*}
\mathrm{WF}_{h,\delta, \gamma}(u) = & \complement \{ (x_0 ,\xi_0) : \exists a \in
{\mathcal S}_{\delta,\gamma} \cap {\mathcal C}^\infty_c, \, a(x_0, \xi_0)
\neq 0, a(x,\xi) = \tilde{a}(h^{-\delta}x, h^{-\gamma}\xi) \\
& \text{ for some } \tilde{a}
\in {\mathcal S} \text{ and }\, a^w u = {\mathcal O}_{L^2}(h^\infty) \}.
\end{align*}
\end{definition}
We have the following immediate corollary.
\begin{corollary}
\label{C:dWF-cor}
If $a \in {\mathcal S}_{\delta, \gamma}(m)$, $0 \leq \delta + \gamma< 1$, and $a$ is real-valued, then
\[
\mathrm{WF}_{h,\delta, \gamma} ( u ) \subset \mathrm{WF}_{h,\delta, \gamma} (a^w u) \cup \overline{\complement \{ \text{ess-supp}\,_h a \}}.
\]
\end{corollary}
The proof is exactly the same as in Theorem \ref{T:wf-properties} only
all symbols must scale the same, so they must be in ${\mathcal S}_{\delta,
\gamma}$.
In
order to conclude the existence of approximate inverses, we need the
restriction $\delta + \gamma < 1$, and the rescaling
operators from \S \ref{S:prelim-rescaling} which can be used to reduce to the familiar $h^{-\delta'}$
calculus, where $\delta' = (\delta + \gamma)/2$, by replacing $h$ with
$h^{\delta -\gamma}$.
\subsection{Rescaling operators}
\label{S:prelim-rescaling}
We would like to introduce $h$-dependent rescaling operators. The rescaling operators should be unitary
with respect to natural Schr\"odinger energy norms, namely the
homogeneous $\dot{H}^s$ spaces. Let us recall in ${\mathbb R}^n$, the $\dot{H}^s$
space is defined as the completion of ${\mathcal S}$ with respect to the
topology induced by the inner product
\[
\left\langle u , v \right\rangle_{\dot{H}^s} = \int_{{\mathbb R}^n} | \xi |^{2s} \hat{u}(\xi)
\overline{\hat{v}}(\xi) d \xi,
\]
where as usual $\hat{u}$ denotes the Fourier transform. The $\dot{H}^0$
norm is just the $L^2$ norm, and the $\dot{H}^1$ norm is
\[
\| u \|_{\dot{H}^1} \simeq \| |\nabla u | \|_{L^2}.
\]
The purpose in taking the homogeneous norms instead of the usual
Sobolev norms is to make the rescaling operators in the next Lemma
unitary.
\begin{lemma}
\label{L:T-lemma}
For any $s \in {\mathbb R}$, $h>0$, the linear operator $T_{h,s}$ defined by
\[
T_{h,s} w(x) = h^{s/2-n/4}w(h^{-1/2} x)
\]
is unitary on $\dot{H}^s( {\mathbb R}^n)$, and for any $r \in {\mathbb R}$,
\begin{align*}
\| T_{h,s} w \|_{\dot{H}^r} & = h^{(s-r)/2} \| w \|_{\dot{H}^r}, \\
\| T_{h,0} w \|_{\dot{H}^s} & = h^{-s/2} \| w \|_{\dot{H}^s}.
\end{align*}
Moreover, for any pseudodifferential
operator $P(x, D)$ in the Weyl calculus,
\[
T_{h,s}^{-1} P(x,D) T_{h,s} = P(h^{1/2} x, h^{-1/2} D).
\]
\end{lemma}
\begin{rema}
Observe that for this lemma, the usual assumption that $h$ be small is not
necessary.
\end{rema}
\begin{proof}
The proof is simple rescaling, but we include it here for the
convenience of the reader. To check unitarity, we just change variables:
\begin{align*}
\left\langle u, T_{h,s} w \right\rangle_{\dot{H}^s} & = \left\langle | \xi |^s \hat{u}, | \xi |^s
\widehat{ T_{h,s} w } \right\rangle \\
& = h^{s/2+n/4} \int | \xi |^{2s} \hat{u}( \xi )
\overline{\hat{w}(h^{1/2} \xi)} d \xi \\
& = h^{-n/4-s/2} \int | \xi |^{2s} \hat{u}( h^{-1/2} \xi )
\overline{\hat{w}(\xi)} d \xi \\
& = h^{n/4 -s/2} \int | \xi |^{2s} \widehat{ u(h^{1/2}\cdot)}(\xi)
\overline{\hat{w}(\xi)} d \xi \\
& = \left\langle T_{h,s}^{-1} u, w \right\rangle_{\dot{H}^s}.
\end{align*}
To check the conjugation property, we again compute
\begin{align*}
T_{h,s}^{-1} P(x,D) T_{h,s} w(x) & = T_{h,s}^{-1} (2 \pi )^{-n}
h^{s/2-n/4} \int
e^{i\left\langle x-y , \xi \right\rangle} P(\frac{x+y}{2}, \xi) w(h^{-1/2} y) dy d \xi
\\
& = T_{h,s}^{-1} (2 \pi )^{-n}
h^{s/2+n/4} \int e^{i \left\langle x-h^{1/2} y, \xi \right\rangle } P(\frac{x + h^{1/2}
y}{2}, \xi ) w(y) d y d \xi \\
& = T_{h,s}^{-1} (2 \pi )^{-n}
h^{s/2-n/4} \int e^{i \left\langle x-h^{1/2} y, h^{-1/2} \xi \right\rangle } P(\frac{x + h^{1/2}
y}{2}, h^{-1/2} \xi )w(y) d y d \xi \\
& = (2 \pi )^{-n}\int e^{i \left\langle x- y, \xi \right\rangle } P(\frac{h^{1/2} ( x +
y)}{2}, h^{-1/2} \xi ) w(y) d y d \xi \\
& = P(h^{1/2} x, h^{-1/2} D) w(x).
\end{align*}
\end{proof}
The purpose of using the rescaling operators $T_{h,s}$ is that if $u
\in \dot{H}^s$ has $h$-wavefront set
\[
\mathrm{WF}_{h,\delta, \gamma} (u) \subset \{ | x | \leq \alpha(h) \text{ and } | \xi | \leq
\beta(h) \},
\]
where, according to the uncertainty principle, $\alpha(h) \beta(h)
\geq h$, then $T_{h,s} u$ has $h$-wavefront set
\[
\mathrm{WF}_{h, \delta-1/2, \gamma + 1/2} (T_{h,s} u) \subset \{ | x | \leq \alpha(h) h^{1/2} \text{ and } | \xi | \leq
\beta(h) h^{-1/2} \},
\]
provided, of course, that $\delta \geq 1/2$.
To see this, we just observe that for any $\psi \in
{\mathcal C}^\infty_c({\mathbb R}^{2n})$, $\psi \equiv 0$ on $\{ | x | \leq 1, | \xi | \leq
1 \}$, we have
\[
\psi( x/\alpha(h) , D / \beta(h) ) u = {\mathcal O}(h^\infty ) \| u \|_{L^2},
\]
or any other semi-norm, and hence
\begin{align*}
\psi( h^{-1/2} x/\alpha(h) , h^{1/2} D / \beta(h) ) T_{h,s} u & =
T_{h,s} \psi( x/\alpha(h) , D / \beta(h) ) T_{h,s}^{-1} T_{h,s} u \\
& = T_{h,s} \psi( x/\alpha(h) , D / \beta(h) ) u \\
& = h^{s/2} {\mathcal O}(h^\infty) \| u \|_{L^2} \\
& = {\mathcal O}(h^\infty) \| u \|_{L^2}.
\end{align*}
Finally, we note that the symbol of the operator on the left-hand side
is zero on the set
\[
\{ | x | \leq h^{1/2} \alpha(h), \text{ and } | \xi | \leq h^{-1/2}
\beta(h) \},
\]
and any such symbol in ${\mathcal S}_{\delta - 1/2, \gamma + 1/2}$, elliptic at a
point outside this set, can be locally
obtained in this fashion.
\section{Geometry near an elliptic geodesic}
Suppose $\Gamma$ is a geodesic in a $n+1$ dimensional Riemannian manifold.
Following \cite[\S 2.1]{Mazzeo-Pacard} (cf. \cite{Tubes}) we fix an arclength parametrization $\gamma(t)$ of $\Gamma$, and a parallel orthonormal frame $E_1, \ldots, E_n$ for the normal bundle $N\Gamma$ to $\Gamma$ in $M$. This determines a coordinate system
\begin{equation*}
x= (x_0, x_1, \ldots, x_n) \mapsto \exp_{\gamma(x_0)}(x_1 E_1 + \ldots x_n E_n) = F(x).
\end{equation*}
We write $x' = (x_1, \ldots, x_n)$ and use indices $j,k, \ell \in \{ 1, \ldots, n\}$, $\alpha, \beta, \delta \in \{ 0, \ldots, n \}$.
We also use $X_\alpha = F_*(\partial_{x_{\alpha}})$.
Note that $r(x) = \sqrt{x_1^2 + \ldots x_n^2}$ is the geodesic distance from $x$ to $\Gamma$ and $\partial_r$ is the unit normal to the geodesic tubes $\{ x: d(x, \Gamma)= \text{cst} \}$.
Let $p = F(x_0, 0)$, $q = F(x_0, x')$, and $r = r(q) = d(p,q)$ then we
have \cite[Proposition 2.1]{Mazzeo-Pacard}, which states
\begin{equation}\label{LocalMet}
\begin{split}
g_{jk}(q) &= \delta_{ij} + \frac13 g(R(X_s, X_j) X_\ell, X_k
)_p x_s x_\ell + \mathcal{O}(r^3) , \\
g_{0k}(q) &= \mathcal{O}(r^2) , \\
g_{00}(q) &= 1 - g(R(X_k, X_0) X_0, X_{\ell} )_p x_k x_\ell +
\mathcal{O}(r^3) , \\
\Gamma_{\alpha\beta}^{\delta} &= \mathcal{O}(r) , \\
\Gamma_{00}^{k} &= -\sum_{j=1}^n g(R(X_k, X_0)X_j, X_0)_p x_j
+ \mathcal{O}(r^2) ,
\end{split}
\end{equation}
where
$\Gamma^\delta_{\alpha\beta} = \frac{1}{2} g^{\delta\eta} ( X_\alpha g_{\eta\beta} + X_{\beta} g_{\alpha\eta} - X_\eta g_{\alpha\beta}).$
In these coordinates, the Laplacian has the form
\begin{equation}\label{LocalLap}
\begin{split}
\Delta_g & := \frac{1}{\sqrt{\det g}} \text{div} (\sqrt{\det g} g^{-1} \nabla) \\
&= g^{jk} X_j X_k - g^{jk}\Gamma_{jk}^\ell X_\ell \\
&= g^{00} X_0X_0 + 2 g^{k0} X_kX_0 - g^{jk}\Gamma^0_{jk} X_0
+2 g^{0j}\Gamma_{0j}^k X_k + \Delta_{\Gamma^{\perp}} ,
\end{split}
\end{equation}
where $\Delta_{\Gamma^{\perp}}$ is the Laplacian in the directions transverse to $\Gamma$.
Denote the geodesic flow on $SM$, the unit tangent bundle, by $\varphi_t$.
Let $\gamma(0) = p \in \Gamma$ and $\zeta = \gamma'(0) \in SM$.
Associated to $\Gamma$ is a periodic orbit $\varphi_t \zeta$ of the geodesic flow on $SM$.
This orbit $\varphi_t \zeta$ is called {\em stable} if, whenever $\mathcal{V}$ is a tubular neighbourhood of $\varphi_t \zeta$, there is a neighbourhood $\mathcal{U}$ of $\zeta$ such that $\zeta' \in \mathcal{U}$ implies $\varphi_t\zeta' \subseteq \mathcal{V}$.
Given a hypersurface $\Sigma$ in $SM$ containing $\zeta$ and transverse to $\varphi_t \zeta$, we can define a {\em Poincar\'e map} $\mathcal{P}$ near $\zeta$, by assigning to each $\zeta'$ the next point on $\varphi_t(\zeta')$ that lies in $\Sigma$.
Two Poincar\'e maps of $\varphi_t\zeta$ are locally conjugate and hence the eigenvalues of $d\mathcal{P}$ at $\zeta$ are invariants of the periodic orbit $\varphi_t\zeta$.
The Levi-Civita connection allows us to identify $TTM$ with the sum of a horizontal space and a vertical space, each of which can be identified with $TM$.
Thus we can choose $\Sigma$ so that $T_\zeta \Sigma$ is equal to $E \oplus E$, where $E$ is the orthogonal complement of $\zeta$ in $T_pM$.
The linearized Poincar\'e map is then given by
\begin{equation*}
\xymatrix @R=1pt
{ E \oplus E \ar[r]^{d\mathcal{P}} & E \oplus E , \\
(V, W) \ar@{|->}[r] & (J( \mathrm{length}(\gamma) ), \nabla_{X_0} J( \mathrm{length}(\gamma) ) ) },
\end{equation*}
where $J$ is the unique Jacobi field along $\gamma$ with $J(0) = V$ and $\nabla_{X_0} J(0) = W$, i.e., $J$ solves
\begin{equation*}
\begin{cases}
\nabla_{X_0} \nabla_{X_0} J + R(J, X_0) X_0 = 0 \\
J(0) = V, \quad \nabla_{X_0} J(0) = W.
\end{cases}
\end{equation*}
The linearized map $d\mathcal{P} = \left.d\mathcal{P}\right|_{\zeta}$ preserves the symplectic form on $E \oplus E$,
\begin{equation*}
\omega( (V_1, W_1), (V_2, W_2) ) = \langle V_1, W_2 \rangle - \langle W_1, V_2 \rangle,
\end{equation*}
and so its eigenvalues come in complex conjugate pairs.
We say that $\Gamma$ is a {\em non-degenerate elliptic closed geodesic} if the eigenvalues of $d\mathcal{P}$ have the form $\{ e^{\pm i \alpha_j} : j = 1, \ldots, n \}$ where each $\alpha_j$ is real and the set $\{ \alpha_1, \ldots, \alpha_n, \pi \}$ is linearly independent over $\mathbb{Q}$.
From \eqref{LocalMet}, the Hessian of the function $g_{00}(q)$ as a function of $x'$ is (minus) the transformation appearing in the Jacobi operator
\begin{equation*}
E \ni V \mapsto R(V, X_0) X_0 \in E.
\end{equation*}
Notice that if $V$ is a normalized eigenvector for this operator, with eigenvalue $\lambda$, then
\begin{equation*}
\mathrm{sec}_p(X_0, V) = g_p(R(V, X_0)X_0, V) = \lambda g_p(V, V) = \lambda.
\end{equation*}
The very useful property
\begin{equation}\label{Assumption}
{ p \in \Gamma, V \in E \implies \mathrm{sec}_p(X_0, V) > 0 },
\end{equation}
holds for any elliptic closed geodesic.
\begin{rema}
One can verify \eqref{Assumption} by means of the Birkhoff normal form (see \cite[\S10.3]{Zel-survey}).
Indeed, if one of the sectional curvatures were negative, then the Birkhoff normal form of the linearized Poincare map (in $T^*M$) must have an eigenvalue off the unit circle. If so, then there is at least one nearby orbit which does not stay nearby, hence the periodic geodesic is not elliptic. Similarly, if one of the curvatures vanishes, then the linearized Poincare map has a zero eigenvalue, and hence the logs of the eigenvalues are not independent from $\pi$ over the rationals (in other words, the Poincare map is degenerate, and not even symplectic).
\end{rema}
\section{Compact Solitons: The nonlinear Ansatz}
We are interested in finding quasimodes for the non-linear Schr\"odinger equation
\begin{eqnarray*}
-\Delta_g u = \lambda u - \sigma | u |^{p} u,
\end{eqnarray*}
where $\sigma = \pm 1$ determines if we are in the focussing or
defocussing case. We will construct $u$ approximately solving this
equation with $u$ concentrated near $\Gamma$ in a sense to be made
precise below.
We take as Ansatz
\begin{equation*}
F(x, h) = e^{ix_0/h} f(x, h), \,\,\, h^{-1} = \frac{2 \pi N}{L},
\end{equation*}
where $L>0$ is the period of $\Gamma$,
and assume for the time being that the function $f$ is concentrated in
an $h$-dependent neighbourhood of $\Gamma$. We are going to employ
a semiclassical reduction, and we are interested in fast oscillations
($h \to 0$), so we assume $\mathrm{WF}_{h,1/2-\delta, 0} f(x,h) \subset \{ | x'| \leq \epsilon
h^{1/2 - \delta }, | \xi' | \leq \epsilon \}$ for
some $\epsilon, \delta >0$\footnote{The reason for the weaker
concentration in frequency $| \xi'|$ is that the nonlinearity forces
working with non-smooth functions, so some decay at infinity in
frequency is lost.}. The localization property of $f$ as constructed
later will be verified
in Section \ref{S:quasimodes}.
We compute from \eqref{LocalLap}, with $\Delta$ the non-positive Laplacian,
\begin{multline}
\label{E:DeltaF}
\Delta F
= e^{ix_0/h} \left[
g^{00} \left( -\frac1{h^2} f + \frac{2i}h X_0 f + X_0X_0 f \right)
+ 2 g^{k0} X_k \left( \frac ih f + X_0 f \right)
\right. \\ \left.
- g^{kj} \Gamma^0_{kj} \left( \frac ih f + X_0 f \right)
+2 g^{0j}\Gamma_{0j}^k X_k f + \Delta_{\Gamma^{\perp}}f \right].
\end{multline}
\begin{rema}
One may initially be inclined to use the Ansatz of the original
Gaussian beam from Ralston
\cite{Ral-beam}, which is
\begin{eqnarray*}
e^{i \psi (x)/h} (a_0 + a_1 h + \dots + a_N h^N),
\end{eqnarray*}
the standard geometric optics quasimode construction.
After all, Ralston is able to make very nice use of the geometry to construct a phase function of the form $i/h (x_0 + \tfrac12 x' B(x_0) x')$
with $\,\mathrm{Im}\, B(x_0) >0$ (for $x_0 \neq 0,$ vanishing otherwise).
In such a regime, however, the non-linear term in the Schr\"odinger equation \eqref{nls}
vanishes to infinite order in $h.$
Thus while such a solution always exists, it fails to capture the effects of the nonlinearity that we are interested in.
\end{rema}
We analyze \eqref{E:DeltaF} by applying the operator $T_{h,s}^{-1}$ in
the variables transversal to $\Gamma$. We normalize everything in the
$L^2$ sense, so we take here $s=0$. Let $z = h^{-1/2}x'$ and set
$v(x_0, z, h) = T_{h,0}^{-1} f(x_0, z, h) = h^{n/4 } f(x_0, h^{1/2}z, h)$.
Notice that the distance to the geodesic $r = |x'|$ is scaled to $\rho
= |z| = h^{-1/2} r$, as described above. In particular, now
\begin{equation}
\label{E:v-WF}
\mathrm{WF}_{h,0, 1/2} v \subset \{ | z | \leq h^{-\delta} \epsilon, \,\, | \zeta | \leq
h^{1/2} \epsilon \},
\end{equation}
if $\zeta$ is the (semiclassical) Fourier dual to $z$.
We conjugate \eqref{E:DeltaF} to get
\begin{multline*}
T_{h,0}^{-1} \Delta T_{h,0} T_{h,0}^{-1} F
= e^{ix_0/h} \left[
g^{00} \left( -\frac1{h^2} v + \frac{2i}h X_0 v + X_0X_0 v \right)
+ 2 g^{k0} h^{-1/2} \partial_{z_k} \left( \frac ih v + X_0 v \right)
\right. \\ \left.
- g^{kj} \Gamma^0_{kj} \left( \frac ih v + X_0 v \right)
+2 g^{0j}\Gamma_{0j}^k h^{-1/2}\partial_{z_k} v + h^{-1} \Delta_{\Gamma^{\perp}}v \right],
\end{multline*}
where the metric components and Christoffel symbols are evaluated at
$(x_0, h^{1/2}z )$.
On the other hand, from \eqref{LocalMet}, expanding in Taylor
polynomials, we know that
\begin{equation}
\begin{split}
g^{00} (x) &= 1 + R_2(x) + R_3(x) + R_4(x) + \mathcal{O}(r^5)
\notag \\
& = 1 + hR_2(z) + h^{3/2}R_3(z) + h^2 R_4(z) + \mathcal{O}(h^{5/2}\rho^5) \\
g^{0k}(x) &= h \tilde{g}^{0k}_2(z) + {\mathcal O}(h^{3/2} \rho^3) \notag
\\
g^{jk} (x) &= \delta_{jk} +
\mathcal{O}(h\rho^2) \label{geomexps} \\
\Gamma^0_{jk}(x) &= h^{1/2}\widetilde\Gamma^0_{jk1} (z)+ h
\widetilde\Gamma^0_{jk2}(z) + \mathcal{O}(h^{3/2} \rho^3)
\notag \\
\Gamma^k_{0j}(x) & = h^{1/2} \widetilde\Gamma^k_{0j1} (z) +
{\mathcal O}(h \rho^2), \notag
\end{split}
\end{equation}
for some smooth functions $R_\ell$, $\tilde{g}^{0k}_\ell$,
$\widetilde{\Gamma}^\alpha_{jk\ell}$ homogeneous of degree $\ell$ respectively. Hence
\begin{multline}
\label{E:TDelta}
T_{h,0}^{-1} \Delta T_{h,0} T_{h,0}^{-1} F
= e^{ix_0/h} \left[
- \frac 1{h^2}v - \frac1h R_2(z)v - \frac1{h^{1/2}} R_3(z) v -
R_4(z) v
+ \frac{2i}h X_0 v
+ \frac {2i}{h^{1/2}} \tilde{g}^{k0}_2 (z)\partial_{z_k}v
\right. \\ \left.
- \frac {i}{h^{1/2}} g^{jk} \widetilde\Gamma^0_{jk1} (z) v
-i g^{jk} \widetilde\Gamma^0_{jk2} (z) v
+ \frac1h \Delta_{\Gamma^{\perp}}v \right] + Pv,
\end{multline}
where $P$ contains the remaining terms from the Taylor expansion. Let us record the following
Lemma.
\begin{lemma}
\label{L:P-error-est}
The operator $P$ has the following expansion:
\begin{eqnarray*}
P & =& {\mathcal O}(h^{1/2} | z |^5) + X_0 X_0 + {\mathcal O}(h | z |^2 )
h^{-1/2} \partial_{z_k} X_0 + {\mathcal O}(|z|^3) \partial_{z_k} + {\mathcal O}(h^{1/2} |
z |^3) \\
&& + {\mathcal O}(h^{1/2} | z |) X_0 + {\mathcal O}(h^{3/2} | z |^{3} )
h^{-1/2} \partial_{z_k}.
\end{eqnarray*}
In particular, if $v$ satisfies \eqref{E:v-WF}, then
\[
\| P v \|_{L^2} \leq C \| v \|_{L^2} + C \| X_0 X_0 v \|_{L^2} + C
h^{1/2-\delta} \| X_0 v \|_{L^2}.
\]
\end{lemma}
\begin{rema}
We will show later that for the particular choice of $v$ we construct,
the operators $X_0$ and $X_0^2$ are bounded operators, so that the
error $Pv$ is bounded in $L^2$ by $v$ (see Remark \ref{rem:x0bdd}).
\end{rema}
For the purposes of exposition, let us then assume for now that the
term $Pv$ is bounded and proceed (this will be justified later). Applying $T_{h,0}^{-1}$ to the equation $-\Delta F = \lambda F -
\sigma |F|^p F$ yields
\begin{align*}
-T_{h,0}^{-1}\Delta T_{h,0} T_{h,0}^{-1} F & = \lambda T_{h,0}^{-1} F -
\sigma T_{h,0}^{-1} (|F|^p F) \\
& = \lambda T_{h,0}^{-1} F -
\sigma h^{-pn/4}|T_{h,0}^{-1} F|^p T_{h,0}^{-1} F,
\end{align*}
so that multiplying \eqref{E:TDelta} by $h e^{-ix_0/h}$, we get
\begin{multline} \label{EqToSolve}
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z))v \\
= \frac{1-h^2\lambda}{h}v
+h^{1/2} \left( i g^{jk} \widetilde\Gamma^0_{jk1} -
2i\tilde{g}^{k0}(z)\partial_{z_k} + R_3(z) \right)v
+ h (R_4(z) +i g^{jk} \tilde{\Gamma}^0_{jk2}) v \\
+ h^{1-pn/4}\sigma |v|^p v
+ he^{-ix_0/h} Pv,
\end{multline}
where $P$ is the same as above.
\begin{rema}
In order to ensure that the nonlinearity appears here as a lower order
term, we require
\begin{equation}
\label{E:p-rest1}
q:=1-pn/4>0, \text{ or } p < \frac{4}{n} = \frac{4}{d-1}
\end{equation}
as stated in the theorems.
\end{rema}
We want to think of the left hand side as similar to a time-dependent harmonic oscillator where $x_0$ plays the role of the time variable.
Let $ q = 1-pn/4$, $0 < q < 1$. We would like to assume that $v$ and
$\lambda$ have expansions in $h^{q}$, however the spreading of
wavefront sets due to the nonlinearity allows us to only take the
first two terms when $0 < q \leq 1/2$ and the first three terms
otherwise.
{\bf Case 1: $0 < q \leq 1/2$.} Assume that $v$ has a two-term expansion
\[
v = v_0 + h^q v_1
\]
and moreover that
there exist
$E_k$, $k=0,1$, satisfying
\begin{equation*}
\frac{1-h^2\lambda}{h} = E_0 + h^{q}E_1 + {\mathcal O}(h^{2q}).
\end{equation*}
Since $q \leq
1/2$, then the ${\mathcal O}(h^{1/2})$
term in \eqref{EqToSolve} is of equal or lesser order than the nonlinear term,
and substituting into \eqref{EqToSolve} we get the following equations
according to powers of $h$:
\begin{equation}\label{SystemEqs}
\begin{split}
h^0: \quad (2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0) v_0 &= 0,\\
h^q : \quad (2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0-h^q E_1) v_1
&= E_1v_0 + \sigma | v_0 |^p v_0 + h^{1/2-q}L v_0 ,
\end{split}
\end{equation}
where
\[
L v_0 = \left( i g^{jk} \widetilde\Gamma^0_{jk1} -
2i\tilde{g}^{k0}(z)\partial_{z_k} + R_3(z) + h^{1/2} R_4(z)+ih^{1/2}g^{jk}\tilde{\Gamma}^{0}_{jk2}
+ h^{1/2} e^{-ix_0/h} P \right) v_0 .
\]
We will show the error terms are ${\mathcal O}(h^{2q})$ in the appropriate
$H^s$ space. See \S \ref{S:q-leq-half}.
{\bf Case 2: $1/2 < q < 1$.} In the case $q > 1/2$, the ${\mathcal O}(h^{1/2})$ term becomes potentially
larger than the nonlinearity, so we take three terms in the expansions
of $v$ and $E=\frac{1-h^2\lambda}{h}$:
\[
v = v_0 + h^{1/2} v_1 + h^q v_2, \,\,\, E = E_0 + h^{1/2} E_1 + h^q
E_2 + \mathcal{O} (h).
\]
We then want to solve
\begin{equation}
\label{SystemEqs2}
\begin{split}
h^0: \quad (2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0) v_0
&= 0,\\
h^{1/2}: \quad (2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0) v_1 &
=
E_1 v_0 + L v_0 \\
h^q : \quad (2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0 -
h^{1/2} E_1 -h^q E_2) v_2
&= E_2 v_0 + h^{1-q} E_1 v_1 + h^{1/2} E_2 v_1 \\
& + \sigma | v_0 |^p v_0 + h^{1-q} L v_1.
\end{split}
\end{equation}
In this case we will show the error is ${\mathcal O}(h^{1/2 + q})$ in the
appropriate $H^s$ space. See \S \ref{S:q-g-half}.
\section{Quasimodes}
\label{S:quasimodes}
We begin by approximately solving the $h^0$ equation by undoing our
previous rescaling. That is, let $w_0 (x_0, x, h) = T_{h.0} v_0(x_0,
x, h) = h^{-n/4} v_0(x_0, h^{-1/2} x, h)$, and conjugate the $h^0$
equation by $T_{h,0}$ to get:
\begin{align*}
0 & = T_{h,0} (2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0) T_{h,0}^{-1}
T_{h,0} v_0 \\
& = (2iX_0 + h\Delta_{\Gamma^{\perp}} - h^{-1} R_2(x) - E_0) w_0,
\end{align*}
where now the coefficients in $\Delta_{\Gamma^\perp}$ are independent
of $h$, and we have used the homogeneity of $R_2$ in the $x$
variables. Multiplying by $h$, we have the following equation:
\begin{equation}
\label{E:sc-orbit}
(2ihX_0 + h^2 \Delta_{\Gamma^{\perp}} - R_2(z) - hE_0) w_0 = 0,
\end{equation}
Hence, \eqref{E:sc-orbit} is a semiclassical equation in a fixed
neighbourhood of $\Gamma$ with symbols in the $h^0$ calculus (i.e. no
loss upon taking derivatives). The principal symbol of the operator
in \eqref{E:sc-orbit} is
\[
p = \tau - | \zeta |^2_{\tilde{g}} - R_2(z),
\]
where $\tilde{g}$ is the metric in the transversal directions to
$\Gamma$, and $\tau$ is the dual variable to the $x_0$ direction. If
we let $\tilde{\Gamma}$ be the (unit speed) lift of $\Gamma$ to
$T^*M$, and if $\exp(s H_p )$ is the Hamiltonian flow of $p$, then
\[
\Gamma = \{ \zeta = z = 0, x_0 = s \in {\mathbb R}/ {\mathbb Z} \}.
\]
Since, in the transversal directions, $p$ is a negative definite
quadratic form, the linearization of the Poincar\'e map $S$ is easy to
compute:
\[
S = \exp (H_{q}),
\]
where
\[
q = -|\zeta|^2_{\tilde{g}} - R_2(z),
\]
so that
\[
H_q = -2 a(x_0, z)^{j,k}\zeta_j \partial_{z_k} + 2 b^{j,k}(x_0) z_j \partial_{\zeta_k},
\]
where $a$ and $b$ are symmetric, positive definite matrices.
Linearizing $S$ about, say, $x_0$ and $z = \zeta = 0$ we get that
$dS(0,0)$ has all eigenvalues on the unit circle, in complex conjugate
pairs.
That is, $\Gamma$ is still a periodic elliptic orbit of the classical
flow of $p$.
Since $p$ is defined on a fixed scale, we can glue $p$ together with
an operator which is elliptic at infinity so that $p$ is of
real principal type so that we can apply Theorem \ref{T:quasi-ch} in
the appendix to construct linear quasimodes. Note that since we have
quasi-eigenvalue of order ${\mathcal O}(h)$, Theorem \ref{T:quasi-ch}
implies the quasimodes are concentrated on a scale $|z| \leq h^{1/2}$,
$| \zeta | \leq h^{1/2}$.
This is made precise in the following proposition.
\begin{prop}
\label{P:solution-to-0-eqn}
There exists $w_0\in L^2$, $\| w_0 \|_{L^2} = 1$, and ${E}_0 = {\mathcal O}(1)$
such that
\[
(2ihX_0 + h^2 \Delta_{\Gamma^{\perp}} - R_2(z) - hE_0)w_0 =
e_0(h).
\]
Here the error $e_0(h) = {\mathcal O}(h^\infty) \in L^2$ (or in any other seminorm), and
$w_0$ has
$h$-wavefront set sharply localized on $\Gamma$ in the sense that if
$\varphi \in \Psi^0_{1/2 - \delta, 1/2 - \delta}$ is $1$ near $\Gamma$, then for any $\delta>0$, $\varphi
w_0 = w_0 + {\mathcal O}(h^\infty)$, and if $\delta = 0$, $\|
\varphi w_0 \| \geq c_0 \| w_0 \|$ for some positive
$c_0$ depending on the support of $\varphi$.
Moreover, $w_0 \in H^\infty(M)$ and satisfies the estimate
\[
\| w_0 \|_{H^s(M)} = {\mathcal O}(h^{-s/2} ).
\]
\end{prop}
\begin{proof}
The construction follows from Theorem \ref{T:quasi-ch},
and is well-known in other settings, see for instance \cite{Ral-beam},
\cite{Bab-geo}, and \cite{CaPo-quasi}.
To get the sharp localization, apply Lemmas \ref{L:comm-1} and
\ref{L:comm-2} from the appendix to get the localization on $w_0$.
Once we know that $w_0$ is so localized, we can replace $w_0$ with
$\varphi w_0$, where $\varphi \in \Psi^0_{1/2 - \delta}$ is as in the
proposition. Then $\varphi w_0$ satisfies
\[
(2ihX_0 + h^2 \Delta_{\Gamma^{\perp}} - R_2(z) - hE_0) \varphi w_0 =
\tilde{e}_0(h),
\]
where
\[
\tilde{e}_0(h) = \varphi e_0(h) + [(2ihX_0 + h^2 \Delta_{\Gamma^{\perp}}
- R_2(z) ), \varphi] w_0.
\]
But now $\varphi e_0(h) = {\mathcal O}(h^\infty)$, while the
commutator is supported outside of an $h^{1/2-\delta}$ neighbourhood
of $\Gamma$, so by the localization of $w_0$ is ${\mathcal O}(h^\infty)$ and
localized in a slightly larger set on the scale $h^{1/2 - \delta}$.
\end{proof}
Now recalling $v_0 = T_{h,0}^{-1} w_0$, then $v_0$ satisfies
\begin{align*}
(2iX_0 + & \Delta_{\Gamma^{\perp}} - R_2(z) - E_0) v_0 \\
& = h^{-1} T_{h,0}^{-1} (2 i hX_0 + h^2 \Delta_{\Gamma^{\perp}} -
R_2(z) - E_0) T_{h,0} {v}_0 \\
& = h^{-1} T_{h,0}^{-1} (2 i hX_0 + h^2 \Delta_{\Gamma^{\perp}} -
R_2(z) - E_0) w_0\\
& = h^{-1} T_{h,0}^{-1} e_0(h).
\end{align*}
The error $h^{-1} T_{h,s}^{-1}e_0(h) = {\mathcal O}(h^\infty)$ in any seminorm
still, but the function $v_0$ is now
localized on a scale $h^{-\delta}$ in space. That is,
\[
\mathrm{WF}_{h, 0, 1 - \delta} v_0 \subset \{ | x | \leq \epsilon
h^{-\delta}, | \xi | \leq \epsilon h^{1 - \delta} \}.
\]
\subsection{The inhomogeneous equation}
We are now in a position to solve the lower order inhomogeneous
equations in \eqref{SystemEqs}. The quasimode $v_0$ has
been constructed as a ``Gaussian beam'' (see \cite{Ral-beam}); it is a harmonic oscillator
eigenfunction extended in the $x_0$ direction by the {\it quantum
monodromy operator} from \cite{SjZw-mono}, which
is defined in \eqref{E:mondef} below. From this construction, the boundedness of
the error term $Pv_0$ as stated in Lemma \ref{L:P-error-est}. In what
follows we construct $v_1$ in the case $0 \leq q \leq {\frac{1}{2}}$ and
$v_1,v_2$ in the case ${\frac{1}{2}} < q < 1$ and show that a similar bound holds
for $h^q Pv_1$ and $h^{\frac{1}{2}} P v_1 + h^q P v_2$ respectively.
We want to solve
\begin{equation}
\label{E:hq}
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - \tilde{E}_0) v_1
= E_1v_0 +G,
\end{equation}
where $\tilde{E}_0= E_0 + h^\eta E_1$ for some $\eta>0$, $E_0$ and $v_0$ have been fixed by the previous construction,
and $G = G(x_0, x)$ is a given function (periodic in $x_0$) with
\[
\mathrm{WF}_{h,0,1/2} G \subset \{ | x | \leq \epsilon h^{-\delta}, | \xi |
\leq \epsilon h^{1/2} \}.
\]
Note, the localization of $G$ follows from the nonlinearity as well as geometric multipliers in the
operator $L$, see Appendix \ref{A:harmosc} and \eqref{E:v-WF}.
Conjugating by $T_{h,0}$ as before, we
get the equation
\begin{equation}
\label{E:hq2}
(2ihX_0 + h^2\Delta_{\Gamma^{\perp}} - R_2(x) - h\tilde{E}_0) w_1
=G_2 ,
\end{equation}
where
\[
G_2 = h T_{h,0}(E_1 v_0 + G).
\]
We observe then that
\[
\mathrm{WF}_{h, 1/2 - \delta, 0} G_2 \subset \{ | x | \leq \epsilon h^{1/2 -
\delta}, | \xi | \leq \epsilon \}.
\]
Specifically, let $M(x_0)$ be the deformation family to
the quantum monodromy operator defined as the solution to:
\begin{eqnarray}
\label{E:mondef}
\left\{ \begin{array}{c}
(2ihX_0 + h^2\Delta_{\Gamma^{\perp}} - R_2(x) - h\tilde{E}_0) M(x_0) = 0, \\
M(x_0) = \,\mathrm{id}\,,
\end{array} \right.
\end{eqnarray}
which exists microlocally in a neighbourhood of $\Gamma \subset
T^*X$ (for further discussion see also
Appendix \ref{quasi} and references therein). By the Duhamel formula, we write
\[
w_1 = M(x_0) w_{1,0} + \frac{i}{h} \int_{0}^{x_0} M(x_0) M(y_0)^* {G}_2
(y_0, \cdot) d y_0.
\]
We have to choose $w_{1,0}$ and $E_1$ (implicit in $G_2$) in such a
fashion to make $w_1$ periodic in $x_0$; in other words, to solve the
equation (approximately) around $\Gamma$. Let $L$ be the primitive period of
$\Gamma$, so that $x_0 = 0$ corresponds to $x_0 = L$. Then we require
\[
w_1(L, \cdot) = w_1(0, \cdot),
\]
or
\[
w_{1,0} = M(L) w_{1,0} + \frac{i}{h} \int_0^L M(L) M(y_0)^*
{G}_2(y_0, \cdot) d y_0.
\]
In other words, we want to be able to invert the operator $(1 -
M(L))$. The problem is that $w_{0,0}: = w_0 (0, \cdot) = T_{h,0} v_0(0, \cdot)$ is in the kernel of
$(1 - M(L))$, so we need to choose $E_1$ in such a fashion to kill the
contribution of ${G}_2$ in the direction of $w_{0,0}$.
Recall that
\begin{align*}
G_2 & = hT_{h,0} (E_1 v_0 + G ) \\
& = h (E_1 w_0 + \tilde{G}),
\end{align*}
where
\[
\tilde{G} = T_{h,0}G.
\]
We want to solve microlocally
\begin{align}
(1 - M(L)) w_{1,0} & = \frac{i}{h} \int_0^L M(L) M(y_0)^*
(h (E_1 w_0 + \tilde{G})) d y_0 \notag \\
& = i \int_0^L M(L) M(y_0)^*
(E_1 M(y_0) w_{0,0} + \tilde{G}) d y_0 \notag \\
& = i L M(L) E_1 w_{0,0} +
i \int_0^L M(L) M(y_0)^*
\tilde{G} d y_0 .\label{E:RHS-orth}
\end{align}
Let
\[
E_1 = - \frac{1}{L} \left\langle \int_0^L M(y_0)^*
\tilde{G} d y_0 , w_{0,0} \right\rangle,
\]
so that \eqref{E:RHS-orth} is orthogonal to $w_{0,0}$.
If we denote
\[
L^2_{w_{0,0}^\perp} = \{ u \in L^2 : \left\langle u, w_{0,0} \right\rangle = 0 \},
\]
then
by the nonresonance assumption (since $E_0 + h^\eta E_1$ is a small
perturbation of $E_0$), and the fact that $M(L)$ is unitary on
$L^2$, $(I -M(L))^{-1}$ is a bounded operator (see \cite{Chr-QMNC})
\[
(1 - M(L))^{-1} : L^2_{w_{0,0}^\perp} \to L^2_{w_{0,0}^\perp}.
\]
Hence
\[
w_{1,0} = (1 - M(L))^{-1} \frac{i}{h} \int_0^L M(L) M(y_0)^*
{G}_2(y_0, \cdot) d y_0
\]
satisfies
\[
\| w_{1,0} \|_{L^2 } \leq C h^{-1} \left\| \int_0^L M(y_0)^*
{G}_2(y_0, \cdot) d y_0 \right\|_{L^2} \leq C L^{1/2} h^{-1} \| G_2
\|_{L^2(x_0) L^2}
\]
and
\[
w_{1,0} \in L^2_{w_{0,0}^\perp}.
\]
Furthermore, we have the estimate
\[
\| w_{1,0} \|_{\dot{H}^s} \leq C h^{-1} h^{-s/2} \left\| \int_0^L M(y_0)^*
{G}_2(y_0, \cdot) d y_0 \right\|_{L^2} \leq C L^{1/2} h^{-1-s/2} \| G_2
\|_{L^2(x_0) L^2},
\]
that is, the $\dot{H}^s$ norm is controlled, but not by the homogeneous
Sobolev norm.
We have proved the following Proposition, which follows simply from
tracing back the definitions.
\begin{prop}
\label{P:inhomog}
Let $v_0$ be as constructed in the previous section, and let $G \in H^s$ for $s\geq 0$ sufficiently large satisfy
\[
\mathrm{WF}_{h,0,1/2} G \subset \{ | x | \leq \epsilon h^{-\delta}, | \xi |
\leq \epsilon h^{1/2} \}.
\]
Then for any $\eta>0$, there exists $v_1 \in L^2$ and $E_1 = {\mathcal O}(\| G \|_{L^2})$ such
that
\[
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0 - h^\eta E_1) v_1
= E_1v_0 +G,
\]
and moreover
\begin{eqnarray*}
\| v_1 \|_{\dot{H}^s} \leq C (\| G \|_{H^s} + \| v_0 \|_{H^s} ).
\end{eqnarray*}
\end{prop}
\begin{rema}
\label{rem:x0bdd}
We note here that by construction of $v_1$ we have implicitly
microlocalized into a periodic tube in the $x_0$ variable. Using the fact that the Quantum
Monodromy Operator is a microlocally unitary operator (see \cite{SjZw-mono,Chr-QMNC}), the bound
\begin{eqnarray*}
\| X_0^j v_1 \|_{L^2} \lesssim \| v_1 \|_{L^2}
\end{eqnarray*}
follows easily for any $j$, which is required for proof of Lemma
\ref{L:P-error-est}.
\end{rema}
\subsection{Construction of quasimodes in the context of Theorem \ref{T:2}}
We now have all the tools to construct the quasimodes which will be used to prove Theorem \ref{T:2}. Let $d\geq 2$ and $0<p<4/(d-1)$. As previously, denote by $q=1-p(d-1)/4$. The main results of this part will be stated in Propositions \ref{PQ1} and \ref{PQ2}.
\subsubsection{The case $0 < q \leq 1/2$, i.e. $\frac2{d-1}\leq p< \frac4{d-1}$}
\label{S:q-leq-half}
In this subsection, we see how to apply Proposition \ref{P:inhomog} in
the case $0 < q \leq 1/2$. As described previously, in this case, the
nonlinearity is the next largest term, and we have only one
inhomogeneous equation so solve (see \eqref{SystemEqs}).
According to Propostion \ref{P:v-symbolic} and Corollary \ref{C:v-symbolic} from Appendix \ref{A:harmosc}, if
\[
G = \sigma | v_0 |^p v_0 - h^{1/2-q}L v_0
\]
is the nonlinear term on the right-hand side, then $G$ is sharply
localized in space but weakly localized in frequency. That is, if
$\chi \in {\mathcal C}^\infty_c( T^*X)$ is equal to $1$ in a neighbourhood of
$\Gamma$, then for any $0 \leq \delta < 1/2$ and any $0 \leq \gamma
\leq 1$,
\[
\chi(h^\delta x, h^{1-\gamma} D_x) G(x_0, x) = G(x_0, x) + E,
\]
where for any $0 \leq r \leq 3/2$,
\[
\| E \|_{\dot{H}^r} \leq C h^{(1 - \gamma)(3/2 + p -r)}.
\]
We are interested in the case where $\gamma = 1/2$, since in that case
$G$ is weakly concentrated in frequencies comparable to $h^{1/2}$, so
by cutting off, satisfies the assumptions of Proposition
\ref{P:inhomog}. That is, take $\gamma = 1/2$, and replace $G$ with
$\tilde{G} = \chi(h^\delta x, h^{1-\gamma} D_x) G$, and apply Proposition
\ref{P:inhomog} to get $v_1$ and $E_1$ satisfying
\[
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0 - h^q E_1) v_1
= E_1v_0 +\tilde{G},
\]
or in other words
\[
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0 - h^q E_1) v_1
= E_1v_0 + \sigma| v_0|^p v_0 + h^{1/2-q}L v_0+ \tilde{Q}_1,
\]
where
\[
\| \tilde{Q}_1 \|_{\dot{H}^r} \leq C h^{(1/2)(3/2 + p -r)}.
\]
Now letting $v = v_0 + h^q v_1$ and $E = E_0 + h^q E_1$, we have solved
\begin{align*}
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) -E) v
& =h^q \sigma| v_0|^p v_0 + h^{1/2}L v_0+ h^q \tilde{Q}_1 \\
& = h^q \sigma |v|^p v + h^{1/2} L v + \tilde{Q}_2,
\end{align*}
where
\[
\tilde{Q}_2 = h^q \tilde{Q}_1 - h^{1/2+q} Lv_1 +{\mathcal O}(h^{2q} | v |^{p+1} ).
\]
The remainder $\tilde{Q}_2$ satisfies
\[
\| \tilde{Q}_2 \|_{\dot{H}^s} \leq C h^{-s/2 + \min \{ 2q, 1/2 + q, q + 3/4 + p/2 \} } =
Ch^{-s/2 + 2q} ,
\]
since $q\leq 1/2$.
Recalling the definitions, $\varphi = e^{ix_0 /h} T_{h,0} v$ satisfies
\begin{itemize}
\item[(i)]
\[
\mathrm{WF}_{h, 1/2 - \delta, 0} \,\varphi \subset \{ | x | \leq \epsilon h^{1/2 -
\delta}, | \xi | \leq \epsilon \};
\]
\item[(ii)]
\[
\| \varphi \|_{L^2} \sim 1, \qquad \| D_{x_0}^\ell \varphi \|_{L^2} \sim h^{-\ell},
\]
and
\[
\| D_{x'}^\ell \varphi \|_{L^2} \leq C h^{-\ell/2} ;
\]
\item[(iii)]
\begin{align*}
(\Delta + \lambda)\varphi & = h^{-1} e^{i x_0/h} T_{h,0} h e^{-i x_0/h} T_{h,0}^{-1}
\Delta T_{h,0} T_{h,0}^{-1} \varphi \\
& = h^{-1} e^{i x_0/h} T_{h,0} (2 i X_0 + \Delta_{\Gamma^\perp} - R_2
-E
- h^{1/2} L) v \\
& = h^{-1} e^{i x_0/h} T_{h,0} ( \sigma h^q | v |^p v +
Q_2),
\end{align*}
or
\[
(\Delta + \lambda)\varphi = \sigma | \varphi |^p \varphi + h^{\alpha(p)}Q,
\]
where $Q = h^{-1} e^{i x_0/h} T_{h,0} \tilde{Q}_2$ satisfies
$\| Q \|_{\dot{H}^s} \leq C h^{-s}$
and where
\begin{equation}\label{error1}
\alpha(p) := -1+ 2q =1 - p \left( \frac{d-1}{2} \right).
\end{equation}
\end{itemize}
We now sum up what we have proven in a proposition. Consider the objects we have just defined : $v=v_{0}+h^{q}v_{1}$, $\varphi_{h}=e^{ix_{0}/h}T_{h,0}v$ and $\lambda(h)=h^{-2}-E_{0}h^{-1}-E_{1}h^{-1+q}$. Then we can state
\begin{proposition}\label{PQ1}
Let $2/(d-1)\leq p<4/(d-1)$ and $\alpha(p)$ be given by \eqref{error1}. Then the function $\varphi_{h}$ satisfies the equation
\begin{equation*}
\big(\Delta+\lambda(h)\big)\varphi_{h}=\sigma |\varphi_{h}|^{p}\varphi_{h}+h^{\alpha(p)}Q(h),
\end{equation*}
where ${Q}(h)$ is an error term which satisfies $\|{Q}(h)\|_{\dot{H}^{s}}\lesssim h^{-s}$, for all $s\geq 0$.
\end{proposition}
\subsubsection{The case $1/2 < q \leq 1$, i.e. $0\leq p< \frac2{d-1}$}
\label{S:q-g-half}
We again construct
$v_0$ as a Gaussian beam using the quantum monodromy operator. We
then set
\[
G_1 = E_1 v_0 + L v_0 ,
\]
which is smooth with compact wavefront set contained in the wavefront
set of $v_0$, so no phase space cutoff is necessary to apply the
inhomogeneous argument to get $v_1$ with wavefront set contained in
the wavefront set of $v_0$.
Now let
\[
G_2 = E_2 v_0 + h^{1-q} E_1 v_1 + h^{1/2} E_2 v_1+ \sigma | v_0 |^p
v_0 + h^{1-q} L v_1,
\]
and solve for $v_2$ as in the previous subsection. This is possible
since $v_1$ is orthogonal to $v_0$ by construction.
We then have
\[
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) - E_0 - h^q E_1) v_2
= E_2 v_0 + h^{1-q} E_1 v_1 + h^{1/2} E_2 v_1+ \sigma | v_0 |^p
v_0 + h^{1-q} L v_1+ \tilde{Q}_1,
\]
where
\[
\| \tilde{Q}_1 \|_{\dot{H}^r} \leq C h^{(1/2)(3/2 + p -r)}.
\]
Letting $v = v_0 + h^{1/2} v_1 + h^q v_2$ and $E = E_0 + h^{1/2} E_1 +
h^q E_2$, we have solved
\begin{align*}
(2iX_0 + \Delta_{\Gamma^{\perp}} - R_2(z) -E) v
& =h^q \sigma| v_0|^p v_0 + h^{1/2}L v_0 + hL v_1 + \tilde{Q}_1 \\
& = h^q \sigma |v|^p v + h^{1/2} L v + \tilde{Q}_2,
\end{align*}
where
\[
\tilde{Q}_2 = h^q \tilde{Q}_1 - h^{1/2 + q} Lv_2 +{\mathcal O}(h^{1/2 + q} | v |^{p+1} ).
\]
We now have the remainder estimate
\[
\| \tilde{Q}_2 \|_{\dot{H}^s} \leq C h^{-s/2+ \min \{ 1/2 + q, q + 3/4 + p/2 \} } = C
h^{-s/2 + 1/2 + q}.
\]
Recalling the definitions, $\varphi: = e^{ix_0 /h} T_{h,0} v$ satisfies
\begin{itemize}
\item[(i)]
\[
\mathrm{WF}_{h, 1/2 - \delta, 0} \,\varphi \subset \{ | x | \leq \epsilon h^{1/2 -
\delta}, | \xi | \leq \epsilon \};
\]
\item[(ii)]
\[
\| \varphi \|_{L^2} \sim 1, \qquad \|D_{x_0}^\ell \varphi \|_{L^2} \sim h^{-\ell},
\]
and
\[
\| D_{x'}^\ell \varphi \|_{L^2} \leq C h^{-\ell/2} ;
\]
\item[(iii)]
\begin{align*}
(\Delta + \lambda)\varphi & = h^{-1} e^{i x_0/h} T_{h_0} h e^{-i x_0/h} T_{h_0}^{-1}
\Delta T_{h_0} T_{h_0}^{-1} \varphi \\
& = h^{-1} e^{i x_0/h} T_{h_0} (2 i X_0 + \Delta_{\Gamma^\perp} - R_2
-E
- h^{1/2} L) v \\
& = h^{-1} e^{i x_0/h} T_{h_0} ( \sigma h^q | v |^p v +
\tilde{Q}_2),
\end{align*}
or
\[
(\Delta + \lambda)\varphi = \sigma | \varphi |^p \varphi + h^{\alpha(p)}Q,
\]
where $Q = h^{-1} e^{i x_0/h} T_{h_0} \tilde{Q}_2$ satisfies
$\| Q \|_{\dot{H}^s} \leq C h^{-s}$ and
where
\begin{equation}\label{error2}
\alpha(p):= -1/2 +q=\frac{1}{2} - p \left( \frac{d-1}{4} \right).
\end{equation}
\end{itemize}
Once again, by construction we have
\begin{eqnarray*}
\| X_0^j v_j \|_{L^2} \lesssim \| v_j \|_{L^2}
\end{eqnarray*}
for $j =0,1,2$.
Consider $v=v_{0}+h^{\frac12}v_{1}+h^{q}v_{2}$, $\varphi_{h}=e^{ix_{0}/h}T_{h,0}v$ and $\lambda(h)=h^{-2}-E_{0}h^{-1}-E_{1}h^{-\frac12}-E_{2}h^{-1+q}$ defined previously, then we have proven
\begin{proposition}\label{PQ2}
Let $0<p\leq 2/(d-1)$ and $\alpha(p)$ be given by \eqref{error2}. Then the function $\varphi_{h}$ satisfies the equation
\begin{equation*}
\big(\Delta+\lambda(h)\big)\varphi_{h}=\sigma |\varphi_{h}|^{p}\varphi_{h}+h^{\alpha(p)}{Q(h)},
\end{equation*}
where ${Q(h)}$ is an error term which satisfies $\|Q(h)\|_{\dot{H}^{s}}\lesssim h^{-s}$, for all $s\geq 0$.
\end{proposition}
\subsection{Higher order expansion and proof of Thereorem \ref{T:0}} \label{sec:nthorder}
\subsubsection{The two dimensional cubic equation}
We first deal with the simpler case $d=2$, $p=2$ and $s=0$. As in the previous section, we define $q=1-p(d-1)/4$, thus for this choice we have $q = \frac12$
allowing us to match powers of the asymptotic
parameters in a canonical way. The general algorithm for any smooth
nonlinearity arising when rescaling in the appropriate $H^s$ space will follow
similarly.
Using \eqref{EqToSolve} and the Taylor expansions of the geometric
components $g^{ij}$ and $\Gamma^i_{jk}$ for $i,j,k = 0,\dots,d$ in
\eqref{geomexps}, we look for an asymptotic series solution of the
form
\begin{equation*}
v = v_0 + h^{\frac12} v_1 + h^{1} v_2 + \dots + h^{m \frac12}
v_{m} + \dots + h^{N \frac12} v_N + \tilde{v} ,
\end{equation*}
for $N$ sufficiently large.
Then, we have the following equations:
\begin{eqnarray*}
h^0 : && (2 i X_0 + \Delta_{\Gamma^\perp} - R_2 (z) - E_0 ) v_0 = 0 ,
\\
h^{\frac12} : && (2 i X_0 + \Delta_{\Gamma^\perp} - R_2 (z) - E_0 ) v_1 =
E_1 v_0 + (i \delta_{jk} \tilde{\Gamma}^0_{jk1} - 2i \tilde{g}^{k0}
\partial_{z_k} + R_3 (z)) v_0 + \sigma |v_0|^2 v_0 , \\
h^{1} : && (2 i X_0 + \Delta_{\Gamma^\perp} - R_2 (z) - E_0 ) v_2 =
E_2 v_0 + E_1 v_1 + (i \delta_{jk} \tilde{\Gamma}^0_{jk1} - 2i \tilde{g}^{k0}
\partial_{z_k} + R_3 (z)) v_1 \\
&&+ (i \delta_{jk} \tilde{\Gamma}^0_{jk2} + R_4) v_0 + \sigma (2 |v_0|^2 v_1 + v_0^2 \bar{v}_1), \\
&\vdots& \\
h^{\frac{m}2} : && (2 i X_0 + \Delta_{\Gamma^\perp} - R_2 (z) - E_0 ) v_m =
\sum_{j=0}^{m-1} E_{m-j} v_j + \sigma (\sum_{j,k,l=0}^{m-1} (c^m_{jkl}
v_j v_k \bar{v}_l)) \\
&&+ \sum_{j=0}^{m-1} (f^{\partial_{z_k}}_{j,m} (z) \partial_{z_k} +
f^{X_0}_{j,m} X_0 + f^1_{j,m} (z) ) v_j , \\
&\vdots& , \\
h^{\frac{N}2} : && (2 i X_0 + \Delta_{\Gamma^\perp} - R_2 (z) - E_0
- \sum_{j=1}^N h^{\frac{j}{2}} E_j) v_N =
\sum_{j=0}^N E_{N-j} v_j + \sigma (\sum_{j,k,l=0}^{N-1} (c^N_{jkl} v_j
v_k \bar{v}_l)) \\
&&+ \sum_{j=0}^{N-1} (f^{\partial_{z_k}}_{j,N} (z) \partial_{z_k} +
f^{X_0}_{j,N} X_0+ f^1_{j,N} (z) ) v_j + P_N v,
\end{eqnarray*}
where
\begin{eqnarray*}
f^{\partial_{z_k}}_{j,m} &=& {\mathcal O}_N (|z|^{m-j}), \\
f^{X_0}_{j,m} &=& {\mathcal O}_N (|z|^{m-j}), \\
f^1_{j,m} &=& {\mathcal O}_N (|z|^{m-j})
\end{eqnarray*}
for $j,m = 0, \dots, N$ and
\begin{eqnarray*}
P_{N} & = & {\mathcal O}(h^{N/2-2} | z |^N) + X_0 X_0 + {\mathcal O}(h^{N/2} | z |^{N} )
h^{-1/2} \partial_{z_k} X_0 \\
&&+ {\mathcal O}(h^{N/2-3/2}|z|^N) \partial_{z_k} + {\mathcal O}(h^{N/2-1} |
z |^N) + {\mathcal O}(h^{N/2} | z |^N) X_0 + {\mathcal O}(h^{N/2} | z |^{N} )
h^{-1/2} \partial_{z_k}.
\end{eqnarray*}
Note that all constants have implicit dependence upon $N$ relating the
number of terms in the expansion at each order. The expansion is
valid provided first of all that
\begin{eqnarray*}
\sum_{j=1}^N h^{\frac{j}{2}} E_j < E_0
\end{eqnarray*}
in order to justify the solvability of the ${\mathcal O} (h^{N/2})$ equation.
\begin{rema}
We note here that
in this expansion, the sign of $\sigma$ can effect the sign and value
of $E_1$, which will impact the remaining asymptotic expansion and in
particular the order of quasimode expansion possible. It is
possible that the focussing/defocussing problem enters in to the
stability analysis of these quasimodes through this point.
\end{rema}
Applying Proposition \ref{P:inhomog} at each asymptotic order and
bounds similar to those in Lemma \ref{L:P-error-est} at order $h^{N/2}$ as in
Section \ref{S:q-leq-half}, we have by a simple calculation that $v$
is a quasimode for the nonlinear elliptic equation with remainder $Q_N$
such that
\begin{eqnarray*}
\| Q_N \|_{\dot{H}^s} \leq C_N h^{-s-1+N/2} .
\end{eqnarray*}
As a result, for sufficiently smooth nonlinearities, one is capable of
constructing higher order asymptotic expansions and hence a quasimode
of higher order accuracy.
\subsubsection{The general case}
Let $d\geq 2$, $p\in 2{\mathbb N}$ and $s\geq0$. We define here $q_{s}=1+p(s-\frac{d-1}4)$. Assume that $p(\frac{d-1}4-s)<1$, or equivalently that $q_{s}>0$. Firstly, write $w=h^{s}v$. Then $v$ has to satisfy \eqref{EqToSolve} but where the power in front of the nonlinearity is $h^{q_{s}}$.
Hence, we can look for $v$ and $E$ of the form
\begin{equation*}
v = \sum_{j,\ell=0}^{N}h^{j/2+\ell q_{s} }\,v_{j,\ell}+\tilde{v} \quad \text{and}\quad E=\sum_{j,\ell=0}^{N}h^{j/2+\ell q_{s} }\,E_{j,\ell}+\widetilde{E}.
\end{equation*}
Since $q_{s}>0$, the nonlinearity does not affect the equation giving $v_{0,0}$, and we have
\begin{equation*}
(2 i X_0 + \Delta_{\Gamma^\perp} - R_2 (z) - E_{0,0} ) v_{0,0} = 0,
\end{equation*}
which is the same equation as before. Then, using Taylor expansions, we write all the equations, similarly to the previous case, in powers of $h^{j/2+\ell q_{s}}$. Again, we can solve each equation and obtain bounds of the solutions and of the error terms. Moreover, it is clear that we can go as far as we want in the asymptotics, so that we can construct a $\mathcal{O}(h^{\infty})$ quasimode, and this proves Theorem \ref{T:0}.
\section{Error estimates}
\subsection{The regular case}
In this section, we assume that $p$ is an even integer.\\
Fix an integer $k>d/2$ (the fact that $k$ is an integer is not necessary). We then define the semiclassical norm
\begin{equation*}
\|f\|_{H^k_{h}}=\|\big(1-{h}^{2}\Delta\big)^{k/2}f\|_{L^2(M)}.
\end{equation*}
In the previous section we have shown the following : Given $\alpha \in \mathbb{R}$, there exist two functions $\varphi_{h}\in H^{k}(M)$ and ${Q}(h)\in H^{k}(M)$ and $\lambda(h) \in \mathbb{R}$ so that
\begin{equation*}
\big(\Delta+\lambda(h)\big) \varphi_{h} +\sigma |\varphi_{h}|^{p}\varphi_{h} +h^{\alpha} {Q}(h).
\end{equation*}
Moreover, microlocally the function $\varphi_{h}$ takes the form
\begin{equation}\label{dephi}
\varphi_{h}(\sigma,x',h)=h^{-\frac{d-1}4+s}e^{i\sigma/h}f(\sigma,h^{-1/2}x',h),
\end{equation}
and we have $\| {Q}(h)\|_{H^{k}_{h}}\leq C$.\\
We set $\displaystyle u_{app}(t,\cdot)=e^{-it \lambda(h)} \varphi_{h}$. Then if we denote by $\widetilde{Q}(h):=e^{-it\lambda}Q(h)$, the following equation is satisfied
\begin{equation}\label{equapp}
i\partial_{t} u_{\text{app}}-\Delta u_{\text{app}}=\sigma | u_{\text{app}}|^{p} u_{\text{app}}+h^{\alpha}\widetilde{Q}(h).
\end{equation}
\begin{prop}\label{properreur}
Let $s\geq 0$. Consider the function $\varphi_{h}$ given by
\eqref{dephi}. Let $u$ be the solution of
\begin{equation}\label{nlsapp}
\left\{
\begin{aligned}
&i \partial_t u-\Delta u=\sigma|u|^{p} u,\\
&u(0,\cdot)=\varphi_{h}.
\end{aligned}
\right.
\end{equation}
Assume that $\displaystyle \alpha>\frac{d+1}4+s+p(-\frac{d-1}4+s)$. Then there exists $C>0$ and $c_{0}>0$ independent of $h$ so that
$$\|u- u_{\text{app}}\|_{L^{\infty}([0,T_{h}];H^{s}(M))}\leq C h^{(d+1)/4},$$
for $0\leq T_{h}\leq c_{0} \,h^{p(\frac{d-1}4-s)}\ln(\frac1h)$.
\end{prop}
This result shows that $u_{app}$ is a good approximation of $u$, provided that the quasimode $\varphi_{h}$ has been computed at a sufficient order $\alpha$.
\begin{proof} Here we follow the main lines of \cite[Corollary 3.3]{Tho3}.
With the Leibniz rule and interpolation we check that for all
$f\in H^{k}(M)$ and $g\in W^{k,\infty}(M)$
\begin{equation}\label{produithh}
\|f\,g\|_{H^k_{h}}\lesssim
\|f\|_{H^k_{h}}\|g\|_{L^{\infty}(M)}+\|f\|_{L^2(M)}
\|\big(1-{h}^{2}\Delta\big)^{k/2}g\|_{L^{\infty}(M)}.
\end{equation}
Moreover, as $k>d/2$, for all $f_1,f_2\in {H^k(M)}$
\begin{equation}\label{produithh2}
\|f_1\,f_2\|_{H^k_{h}}\lesssim h^{-d/2}\|f_1\|_{H^k_{h}}\|f_2\|_{H^k_{h}}.
\end{equation}
Let $u$ be the solution of \eqref{nlsapp} and define
$w=u-u_{\text{app}}$. Then, by \eqref{equapp}, $w$ satisfies
\begin{equation}\label{nlsapp2}
\left\{
\begin{aligned}
&i \partial_t w-\Delta w = \sigma\big( |w+u_{\text{app}}|^{p}
(w+u_{\text{app}})- |u_{\text{app}}|^{p} u_{\text{app}} \big)-h^{\alpha}\widetilde{Q}(h)\\
&w(0,x)=0.
\end{aligned}
\right.
\end{equation}
We expand the r.h.s. of \eqref{nlsapp2}, apply the operator
$\big(1-{h}^{2}\Delta\big)^{k/2}$ to the equation, and take
the $L^2$- scalar product with
$\big(1-{h}^{2}\Delta\big)^{k/2}w$. Then we obtain
\begin{equation}\label{nlsapp5}
\frac{\text{d}}{\text{d}t}\|w\|_{H^k_h}\lesssim \sum_{j=1}^{p+1}
\|w^j\,u^{p+1-j}_{\text{app}}\|_{H^k_h}+h^{\alpha}.
\end{equation}
We now have to estimate the terms
$\|w^j\,u^{p+1-j}_{\text{app}}\|_{H^k_h},$
for $1\leq j\leq p+1$. From \eqref{produithh} we deduce
\begin{equation}\label{nlsapp3}
\|w^j\,u^{p+1-j}_{\text{app}}\|_{H^k_h}\lesssim
\|w^j\|_{H^k_h}\|u^{p+1-j}_{\text{app}}\|_{L^{\infty}(M)}+\|w^j\|_{L^2(M)}\|\big(1-{h}^{2}\Delta\big)^{k/2}u^{p+1-j}_{\text{app}}\|_{L^{\infty}(M)}.
\end{equation}
By \eqref{produithh2}, and as we have
\begin{equation}\label{nlsapp4}
\|u^{p+1-j}_{\text{app}}\|_{L^{\infty}(M)}\lesssim h^{(p+1-j)(-\frac{d-1}4+s)},\quad
\|\big(1-{h}^{2}\Delta\big)^{k/2}u^{p+1-j}_{\text{app}}\|_{L^{\infty}(M)}\lesssim h^{(p+1-j)(-\frac{d-1}4+s)},
\end{equation}
thus inequality \eqref{nlsapp3} yields
\begin{equation*}
\|w^j\,u^{p+1-j}_{\text{app}}\|_{H^k_h}\lesssim h^{-d(j-1)/2} h^{(p+1-j)(-\frac{d-1}4+s)}\|w\|^{j}_{H^k_h}.
\end{equation*}
Therefore, from \eqref{nlsapp5} we have
\begin{equation*}
\frac{\text{d}}{\text{d}t}\|w\|_{H^k_h}\lesssim h^{p(-\frac{d-1}4+s)}\|w\|_{H^k_h}
+ h^{-dp/2}\|w\|^{p+1}_{H^k_h}+h^{\alpha}.
\end{equation*}
Observe that $\|w(0)\|_{H^k_h}=0$. Now, for times $t$ so that
\begin{equation}\label{bootstrap}
h^{-dp/2}\|w\|^{p+1}_{H^k_h} \lesssim h^{p(-\frac{d-1}4+s)}\|w\|_{H^k_h},
\end{equation}
i.e. $\displaystyle \|w\|_{H^{k}_{h}}\leq C h^{(d+1)/4+s}$, we can remove the nonlinear term in \eqref{nlsapp3}, and by the
Gronwall Lemma,
\begin{equation}\label{bootstrap2}
\|w\|_{H^k_h}\leq C h^{\alpha -p(-\frac{d-1}4+s)}e^{Ch^{p(-\frac{d-1}4+s)}t}.
\end{equation}
If $c_{0}>0$ is small enough, and $t\leq c_{0} h^{p(\frac{d-1}4-s)}\ln \frac1h$, then
\begin{equation*}
C h^{\alpha -p(-\frac{d-1}4+s)}e^{Ch^{p(-\frac{d-1}4+s)}t}\leq C h^{(d+1)/4+s},
\end{equation*}
so that inequality \eqref{bootstrap} is satisfied. By the usual bootstrap argument, we infer that for all $$t\leq c_{0} h^{p(\frac{d-1}4-s)}\ln \frac1h$$ we have $$\displaystyle \|w\|_{H^{k}_{h}}\leq C h^{(d+1)/4+s}.$$
Finally, by interpolation we get $\displaystyle \|w\|_{H^{s}} \leq h^{-s}\| w\|_{H_{h}^{k}}$, hence the result.
\end{proof}~
We are now ready to complete the proof of Theorem \ref{T:1}. Consider $u_{h}$, the exact solution to \eqref{nls} with initial condition $\varphi_{h}$, then by the previous proposition and the description of $u_{app}$, we can write
\begin{align}
\| u_{h} \|_{L^\infty([0,T_h]; L^2( M \setminus U_{h^{1/2-\delta
}}))} & \leq \| u_{app} \|_{L^\infty([0,T]; L^2( M \setminus U_{h^{1/2-\delta}}))} + \| u_{h} - u_{app} \|_{L^\infty([0,T]; L^2( M \setminus U_{h^{1/2-\delta}}))} \nonumber\\
& = {\mathcal O}(h^{\infty}) + {\mathcal O}(h^{(d+1)/4})={\mathcal O}(h^{(d+1)/4}),\label{arg}
\end{align}
which was the claim.
\subsection{The non regular case and $d=2$}~
In this section we compute the error estimate in the case of a non smooth nonlinearity in dimension $d=2$. Moreover we restrict ourselves to the case $s=0$ in \eqref{dephi} (case of an $L^{2}$-normalized initial condition).
\begin{prop}\label{properreur2}
Let $\varphi_{h}$ be the function given by \eqref{dephi} with $s=0$. Let $u$ be solution of
\begin{equation*}
\left\{
\begin{aligned}
&i \partial_t u-\Delta u=\sigma|u|^{p} u,\\
&u(0,\cdot)=\varphi.
\end{aligned}
\right.
\end{equation*}
Let $\epsilon>0$. For $p\in (0,4)\backslash \{1\}$, we set $T_{h}=h^{p}$, and in the case $p=1$, $T_{h}=h^{1+\epsilon}$. Then there exists $C>0$ and $\nu>0$ independent of $h$ so that
$$\|u- u_{\text{app}}\|_{L^{\infty}([0,T_{h}];L^{2}(M))}\leq Ch^{\nu}.$$
\end{prop}
\begin{rema}
Note the difference between the results of Propositions \eqref{properreur} (when $s=0$ and $d=2$) and \eqref{properreur2}. In the first case, we have $T_{h}$ of order $h^{p/4}$, which is better than $T_{h}\sim h^{p}$ obtained in the second result. However, in this latter result, there is no restrictive condition on the size of the error term in the equation.
\end{rema}
\begin{proof} First, we follow the strategy of Burq-G\'erard-Tzvetkov \cite[Section 3.]{BGT-comp}
Let $0<p<4$, choose $r>\max(p,2)$ and take $\displaystyle 1-\frac1r<s<1$ (there will be an additional constraint on $s$ in the sequel). Then take $q$ so that $\frac1r+\frac{1}q=\frac{1}2$ and $s_{1}=s-\frac1r$. For $T>0$ define the space
\begin{equation*}
Y^{s}=\mathcal{C}\big([0,T];H^{s}(M)\big)\cap L^{r}\big([0,T];W^{s_{1},q}(M)\big),
\end{equation*}
which is endowed with the norm
\begin{equation*}
\|u\|_{Y^{s}}=\max_{0\leq t\leq T}\|u(t)\|_{H^{s}}+\|(1-\Delta)^{s_{1}/2}u\|_{L^{r}([0,T];L^{q})}.
\end{equation*}
By the Sobolev embeddings, we have $Y^{s}\subset L^{r}\big([0,T],L^{\infty}\big)$.
Now, define $w=u-u_{app}$. Then $w$ satisfies the equation
\begin{equation}\label{nlsapp*}
\left\{
\begin{aligned}
&i \partial_t w-\Delta w = \sigma\big( |w+u_{\text{app}}|^{p}
(w+u_{\text{app}})- |u_{\text{app}}|^{p} u_{\text{app}} \big)-h^{\alpha(p)}\widetilde{Q}(h),\\
&w(0,x)=0,
\end{aligned}
\right.
\end{equation}
with $\alpha(p)=1-p/2$ when $2\leq p\leq 4$ and $\alpha(p)=1/2-p/4$ when $0\leq p\leq 2$ (see \eqref{error1} and \eqref{error2}) and $\|\widetilde{Q}(h)\|_{H^{s}}\leq Ch^{-s}$.\\[5pt]
$\bullet$ Case $0<p<1$. In \cite[Estimate (2.25)]{CaFaHa}, Cazenave, Fang and Han prove that for all $0\leq s<1$
\begin{equation}\label{est.CFH}
\big\||w+u_{app}|^{p}(w+{u_{app}})-|u_{app}|^{p}u_{app}\big\|_{H^{s}}\leq
C\|u_{app}\|_{H^{s}}\|w\|^{p}_{L^{\infty}}+C\|w\|_{H^{s}}\big(\|u_{app}\|^{p}_{L^{\infty}}+\|w\|^{p}_{L^{\infty}}\big).
\end{equation}
Indeed, in \cite{CaFaHa}, the estimate is not stated exactly with these indices, but the proof still holds true. Moreover, in \cite{CaFaHa}, \eqref{est.CFH}
is proved for $x\in \mathbb{R}^{d}$, but the inequality can be adapted to the case of a compact manifold thanks to a partition of unity argument.\\[5pt]
Assume that $w$ satisfies the equation
\begin{equation*}
i \partial_t w-\Delta w =F,\quad w(0,x)=0,
\end{equation*}
then with the Strichartz estimates of \cite{BGT-comp}, the estimate $\|w\|_{Y^{s}}\leq C\|F\|_{L^{1}_{T}H^{s}}$ holds true.
Thus, with the notation $\displaystyle \gamma=1-\frac{p}r$, with \eqref{nlsapp*} and \eqref{est.CFH} we have
\begin{eqnarray}
\|w\|_{Y^{s}}&\leq& C\int_{0}^{T}\|u_{app}\|_{H^{s}}\|w\|^{p}_{L^{\infty}}+C\int_{0}^{T}\|w\|_{H^{s}}\big(\|u_{app}\|^{p}_{L^{\infty}}+\|w\|^{p}_{L^{\infty}}\big)+CTh^{\alpha(p)}\|\widetilde{Q}\|_{L^{\infty}_{T}H^{s}}\nonumber\\
&\leq &CT^{\gamma}\|u_{app}\|_{L_{T}^{\infty}H^{s}}\|w\|^{p}_{Y^{s}}+C\|w\|_{Y^{s}}\big(T\|u_{app}\|^{p}_{L_{T}^{\infty}L^{\infty}}+T^{\gamma} \|w\|^{p}_{Y^{s}}\big)+CTh^{\alpha(p)}\|\widetilde{Q}\|_{L^{\infty}_{T} H^{s}}\nonumber\\
&\leq &CT^{\gamma}h^{-s}\|w\|^{p}_{Y^{s}}+C\|w\|_{Y^{s}}\big(Th^{-p/4}+T^{\gamma} \|w\|^{p}_{Y^{s}}\big)+CTh^{-s+1/2-p/4}.\label{bo1}
\end{eqnarray}
Similarly, we obtain
\begin{equation}\label{bo01}
\|w\|_{L^{\infty}_{T}L^{2}}\leq CT^{\gamma}\|w\|^{p}_{Y^{s}}+C\|w\|_{L^{\infty}_{T}L^{2}}\big(T\|u_{app}\|^{p}_{L_{T}^{\infty}L^{\infty}}+T^{\gamma} \|w\|^{p}_{Y^{s}}\big)+CTh^{1/2-p/4}.
\end{equation}
Therefore, if we define the semiclassical norm $\|\;\;\|_{Y^{s}_{h}}$ by
\begin{equation}\label{semic}
\|u\|_{Y^{s}_{h}}=h^{-s}\|u\|_{L^{\infty}_{T}L^{2}}+\|u\|_{Y^{s}},
\end{equation}
thanks to \eqref{bo1} and \eqref{bo01} we infer
\begin{equation}\label{borne}
\|w\|_{Y^{s}_{h}}\leq CT^{\gamma}h^{-s}\|w\|^{p}_{Y_{h}^{s}}+C\|w\|_{Y^{s}_{h}}\big(Th^{-p/4}+T^{\gamma} \|w\|^{p}_{Y_{h}^{s}}\big)+CTh^{-s+1/2-p/4}.
\end{equation}
Next we use the inequality $\displaystyle ab\leq \frac1{p_{1}}\epsilon^{p_{1}}a^{p_{1}}+\frac1{p_{2}}\epsilon^{-p_{2}}b^{p_{2}}$ which holds for $a,b,\epsilon>0$ and $\displaystyle \frac1{p_{1}}+\frac1{p_{2}}=1$. With a suitable choice of $\epsilon$ and $p_{1}$ (here we use that $0<p\leq 1$) we get
\begin{equation}\label{bo2}
CT^{\gamma}h^{-s}\|w\|^{p}_{Y_{h}^{s}}\leq \frac12 \|w\|_{Y_{h}^{s}}+C(T^\gamma h^{-s})^{1/(1-p)}.
\end{equation}
Now, re-inject \eqref{bo2} into \eqref{borne} and obtain
\begin{equation}\label{bo3}
\|w\|_{Y_{h}^{s}}\leq C\|w\|_{Y_{h}^{s}}\big(Th^{-p/4}+T^{\gamma} \|w\|^{p}_{Y_{h}^{s}}\big)+C(T^\gamma h^{-s})^{1/(1-p)}+CTh^{-s+1/2-p/4}.
\end{equation}
We now perform a bootstrap argument :
Fix $\epsilon >0$ and set $T_{h}=h^{p}$. Fix $\displaystyle 1-\frac1r<s<1-\frac{p}r$. Then it is possible to pick $\nu>0$ small enough so that $\displaystyle \gamma=1-\frac{p}r>s+\frac{\nu(1-p)}{p}$. Assume that
\begin{equation}\label{bo4}
\|w\|_{Y_{h}^{s}}\leq h^{-s+\nu}.
\end{equation}
Then
\begin{equation*}
T_{h}h^{-p/4}+T_{h}^{\gamma} \|w\|^{p}_{Y_{h}^{s}}\leq h^{3p/4}+h^{p(\gamma-s+\nu)},
\end{equation*}
which tends to 0 with $h$, thanks to the assumption made on $\gamma$. Hence, for $h>0$ small enough, with \eqref{bo3} we get
\begin{equation*}
\|w\|_{Y_{h}^{s}}\leq C(T_{h}^\gamma h^{-s})^{1/(1-p)}+CT_{h}h^{-s+1/2-p/4} \leq Ch^{(p\gamma-s)/(1-p)}+Ch^{-s+3p/4+1/2}.
\end{equation*}
Finally, observe that $-s+3p/4+1/2>-s+\nu$, and the assumption $\displaystyle \gamma>s+\frac{\nu(1-p)}{p}$ is equivalent to $(p\gamma-s)/(1-p)>-s+\nu$. Hence for $h>0$ small enough, we recover $\displaystyle \|w\|_{Y_{h}^{s}}\leq \frac12 h^{-s+\nu}$, and by the usual bootstrap argument, the condition \eqref{bo4} holds for $T_{h}=h^{p}$. Now we can deduce the bound $\|u\|_{L^{\infty}_{T}L^{2}}\leq h^{s}\|w\|_{Y_{h}^{s}}\leq h^{\nu}$, which was the claim.
\\[8pt]
$\bullet$ Case $1<p< 4$. Here we have, by \cite[Estimate (2.25)]{CaFaHa}, for all $0\leq s<1$
\begin{multline*}
\big\||w+u_{app}|^{p}(w+{u_{app}})-|u_{app}|^{p}u_{app}\big\|_{H^{s}}\leq \\
C\|u_{app}\|_{H^{s}}( \|u_{app}\|^{p-1}_{L^{\infty}}+\|w\|^{p-1}_{L^{\infty}} )\|w\|_{L^{\infty}}+C\|w\|_{H^{s}}\big(\|u_{app}\|^{p}_{L^{\infty}}+\|w\|^{p}_{L^{\infty}}\big).
\end{multline*}
With the same arguments as for \eqref{bo1} we get, with $\widetilde{\gamma}=1-1/r$
\begin{eqnarray*}
\|w\|_{Y^{s}}&\leq& CT^{\widetilde{\gamma}}\|u_{app}\|_{L_{T}^{\infty}H^{s}}\|u_{app}\|^{p-1}_{L_{T}^{\infty}L^{\infty}}\|w\|_{Y^{s}}+CT^{\gamma}\|u_{app}\|_{L_{T}^{\infty}H^{s}} \|w\|^{p}_{Y^{s}}+\\
&&+C\big(T\|u_{app}\|^{p}_{L_{T}^{\infty}L^{\infty}}+T^{\gamma} \|w\|^{p}_{Y^{s}}\big)\|w\|_{Y^{s}}+CTh^{\alpha(p)}\|Q\|_{L^{\infty}_{T} H^{s}}\nonumber\\
&\leq &C\big( T^{\widetilde{\gamma}}h^{-s-(p-1)/4}+Th^{-p/4} \big)\|w\|_{Y^{s}}+CT^{\gamma}h^{-s}\|w\|^{p}_{Y^{s}} +CT^{\gamma}\|w\|^{p+1}_{Y^{s}} +CTh^{-s+\alpha(p)},
\end{eqnarray*}
with $\alpha(p)=1-p/2$ when $2\leq p\leq 4$ and $\alpha(p)=1/2-p/4$ when $1\leq p\leq 2$ (see \eqref{error1} and \eqref{error2}). Then, by the same manner, we get the following a priori estimate with the semiclassical norm $\|\;\;\|_{Y^{s}_{h}}$ (recall definition \eqref{semic} )
\begin{equation}\label{828}
\|w\|_{Y_{h}^{s}}\leq C\big( T^{\widetilde{\gamma}}h^{-s-(p-1)/4}+Th^{-p/4} \big)\|w\|_{Y_{h}^{s}}+CT^{\gamma}h^{-s}\|w\|^{p}_{Y_{h}^{s}} +CT^{\gamma}\|w\|^{p+1}_{Y_{h}^{s}} +CTh^{-s+\alpha(p)}.
\end{equation}
We now perform the bootstrap : Let $r>\max{(2,p)}$ (there will be an additional constraint on $r$). Fix $\displaystyle 1-\frac1r<s<1$ and set $T_{h}=h^{p}$. Then if $r$ is large enough (recall that $\widetilde{\gamma}=1-1/r$), the term $\displaystyle T_{h}^{\widetilde{\gamma}}h^{-s-(p-1)/4}+T_{h}h^{-p/4}$ tends to 0 with $h$, therefore if $h>0$ is small enough, from \eqref{828} we deduce that
\begin{equation}\label{829}
\|w\|_{Y_{h}^{s}}\leq CT_{h}^{\gamma}h^{-s}\|w\|^{p}_{Y_{h}^{s}} +CT_{h}^{\gamma}\|w\|^{p+1}_{Y_{h}^{s}} +CT_{h}h^{-s+\alpha(p)}.
\end{equation}
Choose $0<\nu<p+\alpha(p)$. As previously we assume that
\begin{equation}\label{pot}
\|w\|_{Y_{h}^{s}}\leq h^{-s+\nu}.
\end{equation}
Then with \eqref{829} we get
\begin{equation*}
\|w\|_{Y_{h}^{s}}\leq Ch^{\gamma p -s+p(-s+\nu)}+Ch^{\gamma p +(p+1)(-s+\nu)}+Ch^{p-s+\alpha(p)}.
\end{equation*}
Next when $r>0$ is large enough (and under the assumption $0<\nu<p+\alpha(p)$), we have $\gamma p -s+p(-s+\nu)>-s+\nu$, $\gamma p +(p+1)(-s+\nu)>-s+\nu$ and $p-s+\alpha(p)>-s+\nu$. To see this, observe that $\gamma\longrightarrow 1$ and $s\longrightarrow 1$ when $r\longrightarrow +\infty$. Therefore for $h>0$ small enough, we recover $\displaystyle \|w\|_{Y_{h}^{s}}\leq \frac12 h^{-s+\nu}$, hence the condition \eqref{pot} holds for $T_{h}=h^{p}$, and similarly to the previous part, we deduce that $\|u\|_{L^{\infty}_{T}L^{2}}\leq h^{s}\|w\|_{Y_{h}^{s}}\leq h^{\nu}$.
\\[8pt]
$\bullet$ Case $p=1$. By \eqref{828} we have
\begin{equation*}
\|w\|_{Y_{h}^{s}}\leq C\big( T^{1-\frac1r}h^{-s}+Th^{-1/4} \big)\|w\|_{Y_{h}^{s}}+CT^{1-\frac1r}\|w\|^{2}_{Y_{h}^{s}} +CTh^{-s+\frac14}.
\end{equation*}
here we set $T_{h}=h^{1+\epsilon}$ with $\epsilon>0$, and we perform the same argument as in the previous case.
\end{proof}
Thanks to this proposition and the same argument as \eqref{arg}, we can conclude the proof of Theorem \ref{T:2}.
|
\section{Introduction}
Since the advent of the microarray technology, high-throughput experiments generating vast amounts of data have been ubiquitous in genetics, for studying e.g. genome-wide patterns of gene expression and copy number alterations. The output of univariate analysis of such high-throughput experiments is often a {\em gene list}, consisting of genes related to some response of interest (e.g. the discrimination between groups or a quantitative trait). The gene list can be ordered or unordered (i.e. ranking the genes by their association to the response or just listing all genes whose association exceeds some threshold) and it can consist of all studied genes or only a subset. The challenge is then to interpret the obtained list in a biological context to understand the underlying processes and generate biologically valid hypotheses. An inherent problem compromising the interpretability of the observed gene lists is that they are often highly unstable, both with regards to small changes in the underlying data set and with regards to changes in the ranking method \citep{Fortunel_etal_03,Irizarry_etal_05,Michiels_etal_05,EinDor_etal_06,Fan_etal_06,Boulesteix_Slawski_09,Abraham_etal_10}. This could be due to redundancy in the cell machinery, i.e. the existence of many genes having similar functions in the cell and thereby being exchangeable in a given experimental list. In this case, the observed gene list depends on the selection of samples in the data set. This means that the functional overlap between two lists may be substantial even though the actual gene overlap is very small. Other possible causes of the apparent instability are noisy measurements and the generally small sample sizes in this type of experiments \citep{EinDor_etal_06,He_Yu_10}.
In this paper we propose a method for stabilization of observed gene rankings, using information extracted from the experimental data. We employ the concept of exchangeability of random variables to quantify the functional redundancy among the genes and we propose a general framework for incorporating exchangeability into the representation of gene lists.
In this framework, each list is represented as a vector in $\mathbb{R}^M$ where $M$ is the number of genes in some universal set, typically the genes measured by a microarray chip. Each entry of the list vector quantifies the contribution to the list from one of the genes.
This representation allows straightforward comparison of any two gene lists by means of any of the conventional measures of similarity or dissimilarity defined on $\mathbb{R}^M\times\mathbb{R}^M$. This is in contrast to previously proposed methods for list comparison, which are tailored to compare specific types of lists. We show that using the proposed method, we obtain gene rankings that are more robust than the original lists against sampling variations in the underlying data, without compromising the relevance to the response.
\section{Related work}
The stabilization of gene rankings has attracted considerable interest during the last decade. Some authors have addressed the ranking method directly and proposed methods providing more robust and accurate ranking results and differential expression detection for the ``large $p$, small $N$'' situation which is standard in biomedical applications (e.g. \citet{Tusher_etal_01,Perelman_etal_07}).
Another way to obtain more stable rankings is to combine the information from several different rankings (e.g. \citet{Rhodes_etal_02,Rhodes_etal_04,Breitling_etal_04,DeConde_etal_06,Hong_Breitling_08,Abeel_etal_10}).
An overview of the most well-known aggregation methods is given by \citet{Boulesteix_Slawski_09}. The most straightforward method is to compute some univariate statistic for each gene from the set of rankings and re-order the genes by their value of this statistic. The statistic can be e.g. the mean of the positions for the gene \citep[e.g.][]{Jurman_etal_08}, a rank product of the positions \citep{Breitling_etal_04} or the fraction of the rankings where the gene is among the top-$k$ genes for some $k$ \cite[e.g.][]{Pepe_etal_03,Jurman_etal_08}. There are also more complex aggregation methods for extracting an optimal top-$k$ list based on e.g. Markov chains \citep{DeConde_etal_06}.
Comparison of gene lists is an essential part of many algorithms, e.g. for enrichment analysis of gene sets \citep[e.g.][]{Draghici_etal_03,Subramanian_etal_05,EinDor_etal_06,Ackermann_Strimmer_09} and assessment of the stability of gene rankings \citep[e.g.][]{Jurman_etal_08,Jurman_etal_10,Abraham_etal_10}. In general, each of the existing methods can only be used to compare specific types of lists, e.g. two ordered lists with the same number of genes, or one ordered list and one short unordered list. Appendix I provides a brief overview of the most well-known list comparison methods and show that many of them can be cast in the here proposed framework if its components are chosen suitably.
\section{Methods}
\subsection{Exchangeability of random variables}
Consider a probability triple $\left(\Omega,\mathcal{F},P\right)$ and let $X_1,\ldots,X_m$ denote random variables on $\Omega$, taking values in some space $\mathbb{M}$. Given $X_1,\ldots,X_m$ we define the multivariate random variable $$X_{1}\times\ldots\times X_{m}:\Omega\rightarrow\underbrace{\mathbb{M}\times\ldots\times\mathbb{M}}_{m}$$by $X_{1}\times\ldots\times X_{m}(\omega)=(X_{1}(\omega),\ldots,X_{m}(\omega)).$ To each random variable $X_{1}\times\ldots\times X_{m}$ there is an associated measure $Pr_{X_{1}\times\ldots\times X_{m}}$ defined by $$Pr_{X_{1}\times\ldots\times X_{m}}(A)=P\left(\left\{\omega\in\Omega;\,\,X_{1}\times\ldots\times X_{m}(\omega)\in A\right\}\right)$$ for all measurable subsets $A\subseteq\mathbb{M}\times\ldots\times\mathbb{M}$.
Conventionally, a finite sequence $(X_1,\ldots,X_m)$ of random variables is called {\em exchangeable} if their joint distribution is invariant under permutation of $X_1,\ldots,X_m$, i.e. if $$Pr_{X_1\times\ldots\times X_m}=Pr_{X_{\pi(1)}\times\ldots\times X_{\pi(m)}}$$ for each $\pi\in S_m$ (the group of permutations of $\{1,\ldots,m\}$). This means that from a statistical point of view the order of the variables in the product is completely irrelevant.
From this definition, it is clear that any sequence of independent and identically distributed (i.i.d.) random variables is exchangeable, but the reverse implication is false. For overviews on exchangeability, see e.g. \citet{Kingman_78,Aldous_85}. The definition of exchangeability given above is rather strong, and we introduce a much weaker notion of exchangeability as follows:
\begin{defi}\label{def:exchangeable} The finite sequence of random variables $\left( X_1,\ldots, X_m\right)$ is {\em weakly exchangeable} if the null sets of $Pr_{ X_1\times\ldots\times X_m}$ are invariant under permutations, i.e. $$Pr_{X_{\pi(1)}\times\ldots\times X_{\pi(m)}}<<Pr_{X_{\tau(1)}\times\ldots\times X_{\tau(m)}}$$ for all $\pi,\tau\in S_m$. Here $\mu<<\nu$ denotes that the positive measure $\mu$ is absolutely continuous with respect to the positive measure $\nu$.\end{defi}
It is clear that a finite sequence of random variables $\left( X_1,\ldots, X_m\right)$ that is exchangeable is weakly exchangeable, but that the opposite implication is false in general.
\subsection{Measures of exchangeability}
In this section we will discuss some ways to quantify the degree of exchangeability for a sequence of random variables.
\begin{defi}Given a finite sequence of random variables $\left( X_1,\ldots, X_m\right)$, the {\em total exchangeability variation} is given by $$P_{ X_1\times\ldots\times X_{m}}^{Var}:=\frac{1}{|S_m|-1}\sum_{\pi\in S_m}\left\|Pr_{X_{\pi(1)}\times\ldots\times X_{\pi(m)}}-\frac{1}{|S_m|}\sum_{\tau\in S_m}Pr_{X_{\tau(1)}\times\ldots\times X_{\tau(m)}}\right\|.$$ Here, $\|\mu\|$ denotes the total variation of the (real-valued) measure $\mu$.\end{defi}
We note that $P_{ X_1\times\ldots\times X_ m}^{Var}=0$ iff the sequence $( X_1,\ldots, X_m)$ is exchangeable.
We now turn to a discrete probability space $\left(\Omega,\mathcal{F},P\right)$, where $\Omega$ is a finite set, $\mathcal{F}=2^\Omega$ is the $\sigma$-algebra consisting of all events and $P:\mathcal{F}\rightarrow[0,1]$ is a probability measure.
We let $X_1,\ldots,X_m$ be random variables on $\Omega$ taking values in $\mathbb{M}:=\{1,\ldots,M\}$. The {\em support} of the random variable $X_{1}\times\ldots\times X_{m}$ is defined by $$\operatorname{supp} X_{1}\times\ldots\times X_{m}:=\left\{\left(q_1,\ldots,q_m\right)\in\mathbb{M}\times\ldots\times\mathbb{M};\,Pr_{X_{1}\times\ldots\times X_{m}}\left(\left\{\left(q_1,\ldots,q_m\right)\right\}\right)>0\right\}.$$For finite sequences of discrete random variables, Definition~\ref{def:exchangeable} implies that $(X_1,\ldots,X_m)$ is weakly exchangeable iff $$\operatorname{supp}\left(X_1\times\ldots\times X_m\right)=\operatorname{supp}\left(X_{\pi(1)}\times\ldots\times X_{\pi(m)}\right)$$for all $\pi\in S_m$. Therefore, to quantify the degree of weak exchangeability for a sequence of discrete random variables we will compare the support of the joint distributions.
Let {$\rho:(\mathbb{M}\times\ldots\times\mathbb{M})\times(\mathbb{M}\times\ldots\times\mathbb{M})\rightarrow\mathbb{R}$} be a metric and define the distance between two sets $A,B\subseteq\mathbb{M}\times\ldots\times\mathbb{M}$ by $$\operatorname{dist}_\rho(A,B):=\min_{a\in A,b\in B}\rho(a,b).$$Furthermore, define the Hausdorff distance between the two sets by $$HD_\rho(A,B):=\max\left(\sup_{a\in A}\operatorname{dist}_\rho(\{a\},B),\sup_{b\in B}\operatorname{dist}_\rho(\{b\},A)\right).$$
\begin{defi}Given a finite sequence of discrete random variables $(X_1,\ldots,X_m)$, the {\em maximal exchangeability distance} is given by $$ED^{max}_{X_1\times\ldots\times X_m}=\frac{\sum_{\pi\in S_m}\sum_{\tau\in S_m}HD_\rho\left(\operatorname{supp} X_{\pi(1)}\times\ldots\times X_{\pi(m)},\operatorname{supp} X_{\tau(1)}\times\ldots\times X_{\tau(m)}\right)}{\rho\left(1_m,M_m\right)|S_m|(|S_m|-1)}$$and the {\em mean exchangeability distance} is given by \begin{align*}ED&^{mean}_{X_1\times\ldots\times X_m}\\&=\frac{\sum_{\pi\in S_m}\sum_{\tau\in S_m}\mathbb{E}_{X_{\pi(1)}\times\ldots\times X_{\pi(m)}}\left[\operatorname{dist}_\rho\left(X_{\pi(1)}\times\ldots\times X_{\pi(m)},\operatorname{supp} X_{\tau(1)}\times\ldots\times X_{\tau(m)}\right)\right]}{\rho\left(1_m,M_m\right)|S_m|(|S_m|-1)}.\end{align*}Here, $1_m=\left(1,\ldots,1\right)$ and $M_m=\left(M,\ldots,M\right)$.\end{defi}
Clearly, $ED^{max}_{X_1\times\ldots\times X_m}=ED^{max}_{X_{\pi(1)}\times\ldots\times X_{\pi(m)}}$ for any $\pi\in S_m$, and $ED^{max}_{X_1\times\ldots\times X_m}=0$ iff $\left(X_1,\ldots,X_m\right)$ is weakly exchangeable, and the same is true for $ED^{mean}_{X_1\times\ldots\times X_m}$.
In the rest of this paper we will mainly consider exchangeability of pairs of discrete random variables with values in $\mathbb{M}=\{1,\ldots,M\}$. In this special case, since $X_1\times X_2$ is the reflection of $X_2\times X_1$ with respect to the line $\{(x,y)\in\mathbb{R}^2;\,y=x\}$, we get \begin{align}P^{Var}_{X_1\times X_2}&=\|Pr_{X_1\times X_2}-Pr_{X_2\times X_1}\|\label{eq:totalvariationpair}\\ED^{max}_{X_1\times X_2}&=\frac{HD_\rho\left(\operatorname{supp} X_1\times X_2,\operatorname{supp} X_2\times X_1\right)}{\rho\left((1,1),(M,M)\right)}\label{eq:maxdistpair}\\ED^{mean}_{X_1\times X_2}&=\frac{\mathbb{E}_{X_1\times X_2}\left[\operatorname{dist}_\rho\left(X_1\times X_2,\operatorname{supp} X_2\times X_1\right)\right]}{\rho\left((1,1),(M,M)\right)}\label{eq:meandistpair}.\end{align}
\subsection{The exchangeability plot}
To illustrate the degree of weak exchangeability of a pair of discrete random variables visually we propose the {\em exchangeability plot}. The exchangeability plot for the random variables $ X_1$ and $ X_2$ is obtained by depicting both $\operatorname{supp} X_1\times X_2$ and $\operatorname{supp} X_2\times X_1$ in the same figure. A pair of random variables $(X_1,X_2)$ is weakly exchangeable iff the two sets overlap completely. Figure~\ref{fig:exchplots} shows two exchangeability plots. The pair of variables in the left panel is weakly exchangeable, while the pair of variables in the right panel is not. Given samples of $X_1\times X_2$ and $X_2\times X_1$ we can also define a {\em sample exchangeability plot}, depicting the observed supports. Further details are provided in Appendix A.
\subsection{The exchangeability of genes}
In this section we use the terminology developed in the previous sections to define the exchangeability of a set of genes with respect to a specific experiment. We will also define another measure of exchangeability which is specifically adapted to the study of ranked gene lists.
We assume that we are given a universal set of $M$ genes, denoted $\mathcal{G}=\{g_1,\ldots,g_M\}$. We use the word {\em experiment} to denote a pair consisting of a population (e.g. cancer patients and healthy control subjects) and a variable ranking method (e.g. a t-test contrasting the two groups in the population). The sample space $\Omega$ consists of all possible rankings of the $M$ genes, and the random variables $X_1,\ldots, X_M:\Omega\rightarrow\{1,\ldots,M\}$ represent the ranking positions of the genes in $\mathcal{G}$. A finite set of genes $\{g_{i_1},\ldots,g_{i_m}\}$ is said to be (weakly) exchangeable iff the corresponding sequence of random variables $(X_{i_1},\ldots,X_{i_m})$ is (weakly) exchangeable. Intuitively, a set of genes is exchangeable iff their positions in the variable ranking can be interchanged without changing the biological interpretation of the ranking.
Of course, we do not know the measures $Pr_{X_i}$ for the variables in practice, so these have to be estimated. In this paper we use subsampling to generate a collection of $B$ data sets for which we compute gene rankings. From the $B$ gene rankings, we construct a {\em position vector} $S_i$ for each gene $g_i$ by collecting all positions of the gene in the $B$ rankings. The elements of a position vector $S_i$ are then samples of the random variable $X_i$. Combining two position vectors $S_i$ and $S_j$ gives samples of the variables $X_i\times X_j$ and $X_j\times X_i$ which can be used to obtain estimates of $P^{Var}_{X_i\times X_j}$, $ED^{max}_{X_i\times X_j}$ and $ED^{mean}_{X_i\times X_j}$. This estimation is discussed further in Appendix A.
We now introduce another measure of exchangeability which is especially adapted to the study of ranked gene lists. The rationale behind this measure is that we may not want two genes to obtain a small exchangeability distance if they always appear in the same order in the rankings. For example, a gene which is always ranked first is not highly exchangeable with a gene that is always ranked second, since the first gene is clearly more strongly related to the response than the second. The new measure penalizes such situations in the computation of the exchangeability distance. For a pair of random variables $(X_1,X_2)$ we define a new set-valued random variable on $\Omega$ by
$$R\left(X_1\times X_2\right)\left(\omega\right):=\{(x,y)\in\mathbb{M}\times\mathbb{M};\,\,\operatorname{sign}\left(x-y\right)=\operatorname{sign}\left(X_1\left(\omega\right)-X_2\left(\omega\right)\right)\}.$$
\begin{defi}The {\em one-sided mean exchangeability distance} for a pair of discrete random variables $(X_1,X_2)$ is defined by $$oED^{mean}_{X_1\times X_2}:=\frac{\mathbb{E}_{X_1\times X_2}\left[\operatorname{dist}_\rho\left(X_1\times X_2,\operatorname{supp} X_2\times X_1\bigcap R(X_1\times X_2)\right)\right]}{\rho\left(\left(1,2\right),\left(M-1,M\right)\right)}$$if $\operatorname{supp} X_2\times X_1\cap R(X_1\times X_2)(\omega)\neq\emptyset$ for all $\omega\in\Omega$ with $Pr_{X_1\times X_2}(X_1\times X_2(\omega))>0$,
and $oED^{mean}_{X_1\times X_2}=1$ otherwise.\end{defi}
Details on the estimation of the one-sided mean exchangeability distance are provided in Appendix A. It is also possible to define a one-sided variant of the maximal exchangeability distance ($oED^{max}_{X_i\times X_j}$) in an analogous manner.
We note that due to the normalization factors introduced in the estimates of weak exchangeability, all measures introduced above attain only values in $[0,1]$. This allows us to define similarity measures ({\em exchangeability scores}) for pairs of genes as follows: $$\begin{array}{lllll}PS^{Var}_{X_i\times X_j}=1-P^{Var}_{X_i\times X_j},&&ES^{mean}_{X_i\times X_j}=1-ED^{mean}_{X_i\times X_j},&&ES^{max}_{X_i\times X_j}=1-ED^{max}_{X_i\times X_j},\\oES^{mean}_{X_i\times X_j}=1-oED^{mean}_{X_i\times X_j},&&oES^{max}_{X_i\times X_j}=1-oED^{max}_{X_i\times X_j}.\end{array}$$
Finally, we define normalized values of the exchangeability scores by comparing them to the corresponding values for pairs of random variables with some pre-specified distribution representing a null hypothesis of no association. In this paper, the main focus is on weak exchangeability of pairs of discrete random variables, in which case it is natural to compare to a random variable $Y_1\times Y_2$ uniformly distributed on a set $S\subseteq\mathbb{M}\times\mathbb{M}$ with cardinality equal to that of $\operatorname{supp} X_1\times X_2$. We show only the normalization for $oES^{mean}_{X_i\times X_j}$, the other scores can be normalized analogously.
\begin{defi}The {\em normalized one-sided mean exchangeability score} for a pair of discrete random variables $(X_1,X_2)$ is defined by \begin{equation}\label{form:normalizedexch}noES^{mean}_{X_1\times X_2}=\left(\frac{oES^{mean}_{X_1\times X_2}-oES^{mean}_{Y_1\times Y_2}}{1-oES^{mean}_{Y_1\times Y_2}}\right)_+\end{equation}where $Y_1\times Y_2$ is a random variable uniformly distributed on a set $S\subseteq\mathbb{M}\times\mathbb{M}$ with $|S|=|\operatorname{supp} X_1\times X_2|$, and $(a)_+=\max(a,0)$. \end{defi}
We note that the measures of exchangeability depend on the number of genes in the ranking ($M$). For two genes having the exchangeability plot shown in the left panel of Figure~\ref{fig:exchplots} we get ${noES}^{mean}_{X_1\times X_2}=1.0$ irrespective of the number of genes since in this case, the distance between any value of $X_1\times X_2$ and $\operatorname{supp} X_2\times X_1$ is zero. For the exchangeability plot in the right panel of Figure~\ref{fig:exchplots} we obtain ${noES}^{mean}_{X_1\times X_2}=0.17$ if $M=15$ and ${noES}^{mean}_{X_1\times X_2}=0.99$ if $M=1,000$. In Appendix H we compute exchangeability matrices for some synthetic example data sets.
\section{A general framework for list representation and comparison}\label{sec:framework}
In this section, we present a general framework for list representation and comparison. The lists are represented as vectors in $\mathbb{R}^M$, where the entry in position $i$ gives the contribution of gene $g_i$. The vector representation allows us to compare both ranked and unranked gene lists within the same framework, using one of the many similarity or dissimilarity measures available to compare vectors in $\mathbb{R}^M$. This is an advantage compared to existing methods for list comparison, which are specifically designed to compare certain types of lists. Our framework also provides a way to determine which genes are most important for explaining the similarity between two lists.
Assume for example that some measure based on the scalar product in $\mathbb{R}^M$ is used to measure the similarity between two vectors $x$ and $y$. Then the value of $x_iy_i$ is a measure of the influence of the $i$'th variable on the similarity between the two lists (see Appendix G).
Finally, ordering the genes by their weights in the vector gives a new ranking of the genes.
As above, we have a universal set of $M$ genes, $\mathcal{G}=\{g_1,\ldots,g_M\}$, where the genes are indexed in a fixed (but otherwise arbitrary) fashion. The universal set can be e.g. all genes on a microarray chip. An ordered (unordered) {\em gene list} is then an ordered (unordered) subset of the universal set. By defining a function $$f:(positions,exchangeabilities,reliability,...)\mapsto l_\ell\in\mathbb{R}^M$$ we use information about various characteristics of the given list to create a vector representation.
\subsection{General idea}
Let $\ell\subseteq\mathcal{G}$ denote a list. If $\ell$ is ordered and if gene $g_i$ is contained in $\ell$, we denote its position by $\pi_\ell(i)$. If $g_i\not\in\ell$, we define $\pi_\ell(i)=0$. For an unordered list, we let $\pi_\ell(i):=\chi_\ell(g_i)$, where $\chi_\ell$ is the characteristic function of the set $\ell$.
Given a list $\ell$ we define a corresponding {\em list matrix} $G_\ell$ as the product of three basic $M\times M$ matrices; $$G_\ell:=A_\ell V_\ell W_\ell.$$The three basic matrices are designed to represent different characteristics of $\ell$. We call $A_\ell$ the {\em position matrix}, $V_\ell$ the {\em exchangeability matrix} and $W_\ell$ the {\em global weight matrix}. From the list matrix we form a {\em list vector} $l_\ell:=((l_\ell)_1,\ldots,(l_\ell)_M),$ by letting $(l_\ell)_i:=h((G_\ell)_i)$ where $(G_\ell)_i$ denotes the $i$'th column of $G_\ell$ and $h:\mathbb{R}^M\rightarrow\mathbb{R}$ is a summarization function, e.g. a norm.
The list vector will be used as the vector representation of the list. Once all lists of interest are represented by vectors in $\mathbb{R}^M$, we can define the similarity between them e.g. as the cosine of the angle between the corresponding list vectors, i.e. $$s(\ell_1,\ell_2)=\frac{l_{\ell_1}\cdot l_{\ell_2}}{\|l_{\ell_1}\|_2\|l_{\ell_2}\|_2},$$where $\cdot$ denotes the inner product in $\mathbb{R}^M$, and we can obtain a dissimilarity coefficient as \begin{equation}\label{eq:distmeas}d(\ell_1,\ell_2)=1-s(\ell_1,\ell_2).\end{equation}
Choosing $A_\ell, V_\ell, W_\ell, h$ and the (dis)similarity coefficient on $\mathbb{R}^M$ suitably, most methods currently available for list comparison fit into this general framework. In Appendix I we show how this can be done for a collection of well-known methods.
\subsection{The position matrix $A_\ell$}
The position matrix $A_\ell$ is defined as a diagonal matrix that contains information about the type of list (ordered or unordered) and the positions of the genes within the list. We define the diagonal element $(A_\ell)_{ii}$ (the position value of gene $g_i$) via a monotone transformation of the ranking statistic of the gene.
This means that the diagonal elements corresponding to genes in the top of the list $\ell$ are high, while the genes further down in the list obtain lower values. All genes not in the list are given position value zero. We note that in some cases, other choices of position values may be better suited for unordered lists, where it may be desirable to give the genes different weights, e.g. depending on some external criterion, even though there is no specified ordering.
\subsection{The exchangeability matrix $V_\ell$}
The exchangeability matrix $V_\ell$ carries information about the exchangeability between the genes in $\mathcal{G}$ in the specific experiment giving the list $\ell$. In most examples in this paper, we define the entry $(V_\ell)_{ij}$ to be the estimated normalized one-sided mean exchangeability score of $g_i$ and $g_j$ (i.e. $\widehat{noES}^{mean}_{X_i\times X_j}$), so the diagonal elements are always 1.
If $V_\ell$ is diagonal, i.e. $V_\ell=I_M$, then the only non-zero elements in the list vector are those corresponding to genes that are actually contained in $\ell$ and consequently only the genes that are present in the list affect its vector representation. However, if $V_\ell$ is not diagonal, there is a possibility that the vector representation of the list is extended, i.e. that it contains non-zero entries for genes which are not themselves present in the list, but are exchangeable with some gene in the list. The (absolute) weight of a gene in the list can also be increased if it is highly exchangeable with a gene with a higher (absolute) position value, since the high exchangeability indicates that the genes could as well have switched positions without changing the interpretation of the list.
We note that the general framework for list representation supports any matrix of gene similarities in the place of $V_\ell$. For example, $V_\ell$ could be defined from some kind of expert biological knowledge, e.g. concerning which genes are related to the same biological function. This could be used for example when comparing lists from different experiments to each other.
Yet another option is to use the positive part of the correlation matrix in place of $V_\ell$, since a high correlation between the expression levels of two genes is often considered to indicate similar biological functions of the genes.
\subsection{The global weight matrix $W_\ell$}
The global weight matrix $W_\ell$ is a diagonal matrix that permits weighting the influence of the genes differently, depending on some informativeness or reliability estimate. For example, we may wish to downweight the influence of a gene that has a high probability to be present in an arbitrarily chosen list, since this gene is unspecific and may not give much relevant information about the similarity between a pair of lists.
\section{Applications}
\subsection{Data sets}
To illustrate the proposed methods, we use three microarray data sets, which were downloaded from \textit{http://www.broadinstitute.org/gsea/datasets.jsp}. These data sets have already been pre-processed by replacing the original probe set IDs with gene symbols and summarizing all probe sets mapping to the same gene by the largest value for each sample. The two lung cancer data sets were re-analyzed by \citet{Subramanian_etal_05}.
\begin{itemize}\item{\bf Boston lung cancer data} \citep{Bhattacharjee_etal_01}. This data set contains expression measurements of 5,217 genes in 62 lung cancer patients, classified according to outcome (good or poor) with 31 observations in each group.
\item{\bf Michigan lung cancer data} \citep{Beer_etal_02}. This data set contains expression measurements of the same 5,217 genes as in the Boston lung cancer data, for 86 lung cancer patients (24 with poor outcome and 62 with good outcome).
\item{\bf Diabetes data} \citep{Mootha_etal_03}. This data set contains expression measurements of 15,056 genes in 17 diabetic patients and 17 control subjects.
\end{itemize}
\subsection{Stabilization of ranked gene lists}\label{sec:stability}
The main objective for introducing exchangeabilities into the list representation is to increase the robustness of the resulting gene list. In this section we evaluate different aspects of the stability of the extended gene lists by comparing with non-extended lists, lists extended using correlations and lists generated by aggregation.
\subsubsection{Stability of top-$k$ gene lists}\label{sec:rankingstability}
In many cases, only the top-ranked genes from an experiment are studied further, which stresses the importance of obtaining a robust and informative set of top-ranked genes. To study the usefulness of exchangeability stabilization for this purpose we apply the following steps to the Boston lung cancer data.
\begin{enumerate}
\item We generate 10 modified data sets by bootstrapping samples from the original data set, taking the class labels into account.
\item\label{step:createrankings} For each modified data set, we rank the genes using five different methods:
\begin{enumerate}
\item Ranking the genes according to their signal-to-noise ratio (SNR) when comparing the two patient groups. Genes positively associated with good outcome are placed in the top. Here, $$SNR(i) = \frac{m_i(good)-m_i(poor)}{\sigma_i(good)+\sigma_i(poor)}$$where $m_i(poor)$ and $\sigma_i(poor)$ denote the mean value and standard deviation of gene $i$ in the patients with poor outcome, and $m_i(good)$ and $\sigma_i(good)$ are the corresponding values in the patients with good outcome. Ranking the genes by their SNR values gives the {\bf non-extended lists}. In general, lists generated in this or similar ways are those that are used for interpretation and biological conclusions.
\item\label{step:extended} Computing the extended list vector as described in Section~\ref{sec:framework} and ranking the variables according to their contribution to the list vector. The position matrix is derived from the SNR-based ranking of the genes, by letting $$(A_\ell)_{ii}=\left\{\begin{array}{cl}\frac{b^2}{(\pi_\ell(i)-1)^2+b^2}&\textrm{if }SNR(i)\geq 0\\-\frac{b^2}{(M-\pi_\ell(i))^2+b^2}&\textrm{if }SNR(i)<0,\end{array}\right.$$with $b^2=350$ and $M=5,217$. We comment on the selection of this function and the parameter values in Appendix B. The position vectors are computed by subsampling the original data set $B=20$ times (each time keeping 2/3 of the samples from each group) and ranking the variables by their SNR values. From the position vectors we then compute the normalized one-sided mean exchangeability scores $\widehat{noES}^{mean}_{X_i\times X_j}$ for all gene pairs to create the exchangeability matrix. We take the global weight matrix $W_\ell=I_M$. To create the list vector from the list matrix, we define
the $i$'th entry of the list vector as the element with the largest magnitude in the $i$'th column of the list matrix $G_\ell$. This means that a gene which is strongly exchangeable with a gene with a highly negative position value can be moved downwards in the list, so the two extreme ends are treated symmetrically. In the final ranking, genes with a highly positive contribution are placed in the top and genes with a highly negative contribution are placed in the bottom. This gives the {\bf extended lists}.
\item Subsampling the modified data set 100 times, and each time deducing a ranking from the SNR values as above. The final ranking is then obtained as an aggregate ranking, by computing the median position of each gene in the 100 subsample rankings. The gene with the lowest median position is placed in the top. This procedure gives the {\bf median aggregated lists}. We also create aggregated lists by computing the product of the ranks of each gene in the 100 subsample rankings and ordering the genes by increasing value of the rank product. This gives the {\bf rank product aggregated lists}.
\item Computing the extended list vector as described in step~\ref{step:extended}, but using the positive part of the correlation matrix of the original data set as the exchangeability matrix. The genes are ordered by decreasing contribution to the resulting list vector. This gives the {\bf correlation extended lists}.
\end{enumerate}
\item The correspondence between the lists from different bootstrap data sets are visualized through concordance plots. For each of the five ranking methods described in step~\ref{step:createrankings}, let $f_k$ denote the number of genes that are among the top-$k$ in the resulting lists from all bootstrapped data sets. The concordance plot depicts $f_k$ as a function of $k$ for $k\in\{1,\ldots,5217\}$. If the lists are highly reproducible with respect to sampling variation in the underlying data, we get $f_k\approx k$ for all $k$. We also construct concordance plots for the reversed lists, i.e. letting $f_k$ be the number of genes that are among the bottom-$k$ in all lists. As another measure of the stability of the gene rankings, we compute the mean overlap between the top-30 and bottom-30 genes from each pair of bootstrapped data sets for each of the five ranking methods defined in step~\ref{step:createrankings}. The resulting figures are given in Appendix E.
\end{enumerate}
The top row in Figure~\ref{fig:concordance} shows concordance plots for the gene lists obtained by the five ranking methods. It is clear that the exchangeability-extended lists are more stable than the lists obtained by the other methods with respect to sampling variations in the underlying data set. Notably, the correlation-extended lists are less stable than the exchangeability-extended lists, indicating that the correlations in this case do not capture the relevant characteristics of the data. The bottom row in Figure~\ref{fig:concordance} shows corresponding concordance plots for gene lists extracted from a data set where the sample labels have been randomly permuted. These figures show that the stability of the extended lists that was noted in the top row is clearly dependent on that the gene lists actually share some information. Hence, the stability is not due to spurious features unrelated to the discrimination between patients with good and poor outcome. For this data set, the correlation between the expression values for a pair of genes has little to do with the estimated exchangeability of the genes (see Appendix D).
\subsubsection{Stability of distances between ranked lists}
Next, we study the stability of the distance between ranked lists from the three different data sets.
First, the data sets are adjusted to contain the same genes, which leaves 5,149 genes. We then compute the exchangeability matrix for each data set (normalized one-sided mean exchangeability scores, position vectors obtained by subsampling $B=20$ times). For each data set, we construct 10 modified data sets by bootstrapping, taking the class labels into account, and compute extended and non-extended list vectors for each bootstrapped data set. The pairwise distances between all extended (or non-extended, respectively) list vectors are computed using (\ref{eq:distmeas}) and we study the variation of the distances from each comparison. For list vectors from different data sets, the aim is to obtain a robust value of the distance. For list vectors from the same data set the distance should also be close to zero.
Figure~\ref{fig:distancestabilityrankedlists} shows histograms of the computed distances for each comparison. It is clear that the lists from the same data set are much more similar after extension than before (the distances are much closer to zero). Comparing the two lung cancer data sets, the extended list vectors are more similar and the distance estimates are more stable than without extension. This suggests that the extended list vectors incorporate information which is shared between the two data sets and that is missed if we only study the top genes. For the comparisons between the lung cancer data sets and the diabetes data set, the non-extended list vectors are almost orthogonal (implying a dissimilarity around 1), which indicates that the top genes from the data sets are completely different. The dissimilarities are close to 1 also after extension, which suggests that there is much less shared functional information between a lung cancer data set and the diabetes data set than between the two lung cancer data sets. In Appendix F we give the corresponding histograms for rankings obtained from data sets were the class labels are permuted independently in each bootstrap round.
\subsection{Informativeness of ranked gene lists}
Although stability of gene rankings is an important and desirable property, it is not the only thing that is of interest. For example, if we define a ranking method which always assigns a gene the same pre-defined position, the ranking would be extremely stable but most likely useless. We therefore study the informativeness of the rankings obtained as described above by examining the ability of the top-ranked genes in each list to discriminate between the two patient groups in the Boston lung cancer data. We use ten-fold cross-validation to assess the performance of the classifiers. For each training/test set split we compute the five rankings as described in Section~\ref{sec:rankingstability} for the training set, and extract the top- and bottom-$k$ genes from each ranking. The expression levels for these genes are centered and standardized based on their mean value and standard deviation in the training set. The standardized expression levels of the selected genes are then used as features in a centroid classifier \citep{Scholkopf_Smola_02} which is used to classify the remaining (test) samples. The reported classification accuracy is the mean area under the receiver operating characteristic curve (AUC) across the 10 training/test set splits.
Table~\ref{tab:classificationability} shows the estimated classification accuracy for the top and bottom genes from the five rankings as well as the mean classification accuracy for top and bottom genes from 20 random rankings. Note that the top-ranked gene is always the same for the extended and non-extended rankings. The classification ability of the top-ranked genes in the extended list vectors is considerably higher than for the other methods, indicating that the increased stability observed in the previous section does not come at the expense of decreased biological significance. \begin{table}
\caption{Classification accuracy (mean AUC across 10 training/test set splits) for the union of the top- and bottom-$k$ lists obtained from different rankings of the genes in the Boston Lung Cancer data with respect to their association with the discrimination between patients with good and poor outcome. The best performing method for each $k$ is highlighted in bold. }
\centering
\begin{tabular}{lcccc}
&$k=1$&$k=10$&$k=30$&$k=100$\\
\hline
Extended&0.344&{\bf 0.774}&{\bf 0.837}&{\bf 0.832}\\
Non-extended&0.344&0.583&0.606&0.566\\
Median aggregated&0.410&0.565&0.617&0.566\\
Rank product&0.400&0.538&0.606&0.566\\
Correlation extended&0.344&0.594&0.572&0.594\\
Random&{\bf 0.464}&0.489&0.473&0.478\\
\end{tabular}
\label{tab:classificationability}
\end{table}
\section{Discussion}
Univariate analysis of multivariate genetic data sets usually results in a ranking of the variables according to some criterion. This ranking is then interpreted to gain biological knowledge and understanding. However, it has been noted that the variable rankings are often highly unstable with respect to small changes in the underlying data set or the method used to obtain the ranking and therefore, methods for stabilizing the variable ranking and allowing more robust comparison to other lists are much needed. In this paper we have presented a general framework for robust representation and comparison of variable lists. The framework encompasses both ordered and unordered lists, which can therefore be compared on similar terms. Having a robust measure of similarity between any pair of variable lists can furthermore enable visualization through e.g. multidimensional scaling to obtain a low-dimensional visual representation of large collections of lists. We have shown that the extended variable lists are more stable than the original variable lists from an experiment, and also more stable than lists obtained by aggregation of several lists from subsampled data sets. These results suggest that the exchangeability concept for random variables may be a suitable tool for quantifying the functional redundancy among genes and incorporating this information into the list representation.
The results from the proposed method can be used in different ways. Given the vector representation of the gene lists there are many natural choices of similarity and dissimilarity measures that can be applied to compare lists. By ranking the genes according to their contribution to the list vector we also obtain a new gene ranking which may be used to obtain more robust results with other methods, such as e.g. Gene Set Enrichment Analysis (GSEA) \citep{Subramanian_etal_05} to study the enrichment of gene sets among the genes most highly related to a response.
\bibliographystyle{natbib}
|
\section{Motivation and Summary}
Assume that a particle system $\eta_s(x),\,x\in\mathbb{Z}^d,\,s\ge 0$,
gives raise to a scaled field of type
\begin{equation}\label{field}
Y_s^\varepsilon\,=\,
\varepsilon^{\lambda_d}
\sum_x\frac{\eta_{s\varepsilon^{-\kappa}}(x)-a}{\chi}
\,\delta_{\varepsilon x-bs\varepsilon^{-\tilde{\kappa}}}
,\quad s\ge 0,
\end{equation}
where $\delta_{\varepsilon x-bs\varepsilon^{-\tilde{\kappa}}}$
denotes the Dirac measure
concentrated in the macroscopic point
$\varepsilon x-bs\varepsilon^{-\tilde{\kappa}}$
and one wishes to understand the limiting behaviour,
$\varepsilon\downarrow 0$,
of this field.
Usually it follows from the martingale problem
for the strong Markov process $\eta_s(x)$
that there is an approximate equation
for $Y_s^\varepsilon$, $\varepsilon$ small, which often reads like
\begin{equation}\label{approx equ}
{\rm d} Y_s^\varepsilon(G)\,\sim\,
Y_s^\varepsilon({\cal A}G)\,{\rm d} s\,+\,
V_\varepsilon^G(s,\xi_{s\varepsilon^{-\kappa}})\,{\rm d} s\,+\,
{\rm d} M_s^{G,\varepsilon}
\end{equation}
where $G$ is a smooth test function on $\mathbb{R}^d$ with compact support,
${\cal A}$ a partial differential operator,
$V_\varepsilon^G$ is for fixed $\varepsilon,G$ a bounded measurable function,
$\xi_{s\varepsilon^{-\kappa}}$ stands for
$(\eta_{s\varepsilon^{-\kappa}}(x)-a)/\chi$
and $M^{G,\varepsilon}$ denotes a martingale.
For further analysis of this approximate equation it becomes necessary
to express $V_\varepsilon^G(s,\xi_{s\varepsilon^{-\kappa}})$ in terms
of the field $Y_s^\varepsilon$, that is, one wants to replace
$$\int_0^t V_\varepsilon^G(s,\xi_{s\varepsilon^{-\kappa}})\,{\rm d} s
\quad\mbox{by}\quad
\int_0^t F(s,Y_s^\varepsilon,G)\,{\rm d} s$$
in some sense where the functional $F$ has to be found
(see \cite{KL1999} for a good review of this method). Typically the
difference
$$\int_0^t F(s,Y_s^\varepsilon,G)\,{\rm d} s
\,-\int_0^t V_\varepsilon^G(s,\xi_{s\varepsilon^{-\kappa}})\,{\rm d} s
\quad\mbox{simplifies to}\quad
\sum_{i=1}^m
\int_0^{t}
V_\varepsilon^{G,i}(s,\xi_{s\varepsilon^{-\kappa}})\,{\rm d} s$$
and the task is to estimate an appropriate norm of
$t\mapsto\int_0^{t}\hspace{-1pt}
V_\varepsilon(s,\eta_{s\varepsilon^{-\kappa}}){\rm d} s$
where $V_\varepsilon(s,\eta)$ stands for one of the functions
$V_\varepsilon^{G,i}(s,(\eta-a)/\chi),\,i=1,\dots,m$.
First, for an arbitrary but fixed $\beta>0$
and a finite time horizon $T$,
it follows from Lemma \ref{lem_inhomo}
in Section \ref{section resolv method} that
\begin{equation}\label{inhomo}
\int_0^T\hspace{-5pt}{\rm d} t\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V_\varepsilon(s,\eta_{s\varepsilon^{-\kappa}})\,{\rm d} s\,]^2
\,\le\,
\frac{2e^{\beta T}}{\beta}\,\varepsilon^{2\kappa}
(\tilde{V}_\varepsilon\,|\,
G_{\frac{\beta}{2}\varepsilon^{\kappa}}\tilde{V}_\varepsilon)_{L^2({\rm d} s\otimes{\rm d}\nu)}
\end{equation}
where $\tilde{V}_\varepsilon$ stands for the function
$(s,\eta)\mapsto e^{-\frac{\beta}{2}s\varepsilon^{\kappa}}
V_\varepsilon(s\varepsilon^{\kappa},\eta)$
and $(G_\alpha)_{\alpha>0}$ denotes the strongly continuous contraction resolvent
associated with the process $(s,\eta_s)_{s\ge 0}$
on the Hilbert space $H=L^2({\rm d} s\otimes{\rm d}\nu)$
assuming that there exists an invariant state $\nu$
of the system $(\eta_s)_{s\ge 0}$.
Notice that (\ref{inhomo}) is valid in the context of general right
processes.
The observation is now that in many cases one has the inequality
\begin{equation}\label{not new}
(u\,|\,G_{\alpha}u)
\,\le\,
\left(\rule{0pt}{10pt}\right.\!
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_{s})^{-1}\,
\!\left.\rule{0pt}{10pt}\right)
\quad u\in H
\end{equation}
for all $\alpha>0$ by abstract theory on resolvents
where $L_{s}$ stands for the symmetric part of the generator of the
resolvent $(G_\alpha)_{\alpha>0}$ in $H$.
Second, choosing $\alpha=\frac{\beta}{2}\varepsilon^{\kappa}$ and
$u=\tilde{V}_\varepsilon$ in (\ref{not new})
and applying (\ref{inhomo}) yields
\begin{equation}\label{inhomo separated}
\int_0^T\hspace{-5pt}{\rm d} t\,
{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V_\varepsilon(s,\eta_{s\varepsilon^{-\kappa}})\,{\rm d} s\,]^2
\,\le\,
\frac{2e^{\beta T}}{\beta}\,
\varepsilon^\kappa\hspace{-3pt}\int_0^T\hspace{-5pt}
\left(\rule{0pt}{10pt}\right.\!
V_\varepsilon(s,\cdot)
\;\rule[-4pt]{1pt}{14pt}\;
(\frac{\beta\varepsilon^\kappa}{2}-L^{\rm sym})^{-1}V_\varepsilon(s,\cdot)
\!\left.\rule{0pt}{10pt}\right)_{L^2(\nu)}{\rm d} s
\end{equation}
where $L^{\rm sym}$ denotes the symmetric part of the generator of the
process $(\eta_s)_{s\ge 0}$ in $L^2(\nu)$.
The details of how to replace $L_s$ by $L^{\rm sym}$
are explained by Lemma \ref{decouple}
in Section \ref{section resolv method}.
Remark that applying
Kipnis-Varadhan's inequality\footnote{See \cite[Lemma 4.3]{CLO2001}
for the version used here.} which is widely used in the context of
particle systems would yield
$$\int_0^T\hspace{-5pt}{\rm d} t\,
{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V_\varepsilon(s,\eta_{s\varepsilon^{-\kappa}})\,{\rm d} s\,]^2
\,\le\,
14T\,\varepsilon^\kappa\hspace{-3pt}\int_0^T\hspace{-5pt}
\left(\rule{0pt}{10pt}\right.\!
V_\varepsilon(s,\cdot)
\;\rule[-4pt]{1pt}{14pt}\;
(-L^{\rm sym})^{-1}V_\varepsilon(s,\cdot)
\!\left.\rule{0pt}{10pt}\right)_{L^2(\nu)}{\rm d} s$$
instead of (\ref{inhomo separated}).
But, in the important case where
$(\eta_s)_{s\ge 0}$ is a simple one-dimensional exclusion process and
$\nu$ is the symmetric Bernoulli product measure on $\{0,1\}^\mathbb{Z}$,
the right-hand side of the last inequality
is infinite for $V_\#(\eta)=(\eta(0)-1/2)(\eta(1)-1/2)$
which is the quadratic part of the normalised current and the most
basic quadratic fluctuation.
So Kipnis-Varadhan's inequality cannot be used
to estimate time integrals of
$\varepsilon$-scaled quadratic fluctuations
$V_\varepsilon$ build from $V_\#$.
However, the right-hand side of the inequality (\ref{inhomo separated})
is always finite for bounded measurable functions $V_\varepsilon$.
Hence, when substituting $V_\varepsilon^{G,i}$ for
$V_\varepsilon,\,i=1,\dots,m$, this inequality
gives a tool for how to show
\begin{equation}\label{tool}
\lim_{\varepsilon\to 0}\;
\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\nu\left(
\int_0^t F(s,Y_s^\varepsilon,G)\,{\rm d} s
\,-\,
\int_0^t V_\varepsilon^G(s,\xi_{s\varepsilon^{-\kappa}})\,{\rm d} s
\right)^{\hspace{-2pt}2}\,=\,0.
\end{equation}
Remark that replacing
$\int_0^t V_\varepsilon^G(s,\xi_{s\varepsilon^{-\kappa}})\,{\rm d} s$
by
$\int_0^t F(s,Y_s^\varepsilon,G)\,{\rm d} s$
in the sense of the above limit
is rather \underline{weak} since the replacement does \underline{not}
hold for every $t\in[0,T]$ but only for an average over $t\in[0,T]$.
Nevertheless, as recently shown in \cite{A2012},
this weak form of a replacement
is still sufficient for deriving a meaningful
equation which could be used as the limit of (\ref{approx equ}).
Section \ref{section resolv method} presents the resolvent method
which is based on
(\ref{inhomo}),(\ref{not new}),(\ref{inhomo separated}).
A detailed proof of the inequality (\ref{not new})
is given in a general setting (see Corollary \ref{my inequ}).
The result as such cannot be new. However, the author could not
find a reference. The proof is based on a variational formula
(see Lemma \ref{variational}) which was also used in the proof of
\cite[Lemma 2.1]{LQSY2004} but without explicit proof and only in
the framework of exclusion processes.
The detailed proof is added for completeness and for
having a good account on the precise conditions needed.
In Section \ref{section application} the resolvent method is applied
in the case of $\sqrt{\varepsilon}$-asymmetric one-dimensional simple
exclusion in equilibrium. The field (\ref{field}) of interest is the diffusively
scaled density fluctuation field. In the corresponding equation
(\ref{approx equ}), ${\cal A}$ is the one-dimensional Laplacian and
the bounded functions $V_\varepsilon^G$ are
$\varepsilon$-scaled quadratic fluctuations
build from $V_\#$ introduced above.
The key result is Lemma \ref{key result} which leads to
a replacement of the time integrals of these quadratic fluctuations in
the sense of (\ref{tool}), see Corollary \ref{weak replacement}.
Remark that the density fluctuations in
$\sqrt{\varepsilon}$-asymmetric exclusion are related to non-trivial
distributions like the Tracy-Widom distribution hence they are a good
`medium' for testing techniques. Proving Lemma \ref{key result} using a
resolvent-type method can be considered to be such a test.\\
\noindent
{\it Acknowledgement}.
The author thanks Wilhelm Stannat for helpful discussions.
\section{A Resolvent Method}\label{section resolv method}
Let $X$ be a Hausdorff topological space and assume that the
Borel-$\sigma$-algebra on $X$ is equal to the $\sigma$-algebra
generated by the set of all continuous functions on $X$.
Denote by
$(\Omega,{\cal F},{\bf P}_{\!\!\eta},\eta\in X,(\eta_s)_{s\ge 0})$
a right process with state space $X$, infinite life time and
corresponding filtration ${\cal F}_s,\,s\ge 0$, satisfying the usual
conditions (see \cite{S1988} for a good account on general right
processes). Assume that there exists a measure $\nu$ on $X$ which is
an invariant state of this right process and denote by ${\bf E}_\nu$ the
expectation operator given by the probability measure
$\int_X{\bf P}_{\!\!\eta}\,\nu({\rm d}\eta)$.
Obviously, the pair $(s,\eta_s)_{s\ge 0}$ gives another right process
$(\Omega,{\cal F},{\bf P}_{\!\!s,\eta},
(s,\eta)\in[0,\infty)\times X,(s,\eta_s)_{s\ge 0})$
corresponding to the same filtration ${\cal F}_s,\,s\ge 0$, such that
$${\bf P}_{\!\!s,\eta}\left(\rule{0pt}{10pt}\right.
\{\eta_s=\eta\}
\left.\rule{0pt}{10pt}\right)\,=\,1
\quad\mbox{for all}\quad(s,\eta)\in[0,\infty)\times X.$$
Define the transition semigroup and the resolvent of $(s,\eta_s)_{s\ge 0}$
by
$$p_r V(s,\eta)\,=\,{\bf E}_{s,\eta}V(s+r,\eta_{s+r}),\quad r\ge 0,$$
and
$$R_\alpha V(s,\eta)\,=\,
\int_0^\infty{\rm d} r\,e^{-\alpha r}\,p_r V(s,\eta),\quad\alpha>0,$$
respectively, where $V$ is an arbitary bounded measurable function on
$[0,\infty)\times X$.
Denote by $\ell$ the Lebesgue measure on $[0,\infty)$ and notice that
$\ell\otimes\nu$ is an excessive measure on $[0,\infty)\times X$
with respect to
$(\Omega,{\cal F},{\bf P}_{\!\!s,\eta},
(s,\eta)\in[0,\infty)\times X,(s,\eta_s)_{s\ge 0})$
because
$$\int p_r V\,{\rm d}(\ell\otimes\nu)\,\le\,
\int V\,{\rm d}(\ell\otimes\nu),\quad r\ge 0,$$
for all non-negative measurable functions
$V$ on $[0,\infty)\times X$.
As a consequence, see Section IV.2 in \cite{MR1992} for the details,
there exists a strongly continuous contraction resolvent
$(G_\alpha)_{\alpha>0}$ on $L^2(\ell\otimes\nu)$ such that
$G_\alpha V$ is an $(\ell\otimes\nu)$-version of $R_\alpha V$ for all $\alpha>0$
and all bounded functions $V$ in $L^2(\ell\otimes\nu)$.
\begin{lemma}\label{lem_inhomo}
Fix $\beta>0$, let $T>0$ be a finite time horizon and consider the
process $(\eta_{cs})_{s\ge 0}$ time-scaled by a factor $c>0$. Then
$$\int_0^T\hspace{-5pt}{\rm d} t\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V(s,\eta_{cs})\,{\rm d} s\,]^2
\,\le\,
\frac{2e^{\beta T}}{\beta c^2}\,
(\tilde{V}\,|\,
G_{\!\frac{\beta}{2c}}\tilde{V})_{L^2(\ell\otimes\nu)}$$
for all bounded functions $V$ in $L^2(\ell\otimes\nu)$
where $\tilde{V}(s,\eta)=e^{-\frac{\beta}{2}s/c}\,V(s/c\,,\eta)$.
\end{lemma}
{\it Proof}. Choose a bounded function $V$ in $L^2(\ell\otimes\nu)$
and set $\hat{V}(s,\eta)=V(s/c\,,\eta)$. Then:
$$\int_0^T\hspace{-5pt}{\rm d} t\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V(s,\eta_{cs})\,{\rm d} s\,]^2
\,\le\,
e^{\beta T}
\int_0^\infty\hspace{-5pt}{\rm d} t\,e^{-\beta t}\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
\hat{V}(cs,\eta_{cs})\,{\rm d} s\,]^2$$
$$=\,\frac{2e^{\beta T}}{c^2}
\int_0^\infty\hspace{-5pt}{\rm d} t\,e^{-\beta t}\,{\bf E}_\nu
\int_0^{ct}\hspace{-1pt}{\rm d} s\,\hat{V}(s,\eta_{s})
\int_s^{ct}\hspace{-1pt}{\rm d} r\,\hat{V}(r,\eta_{r})$$
$$=\,\frac{2e^{\beta T}}{c^2}
\int_0^\infty\hspace{-5pt}{\rm d} t\,e^{-\beta t}\hspace{-5pt}
\int_0^{ct}\hspace{-5pt}{\rm d} s\int_s^{ct}\hspace{-5pt}{\rm d} r\,
{\bf E}_\nu\hat{V}(s,\eta_{s})\,p_{r-s}\hat{V}(s,\eta_{s})$$
$$=\,\frac{2e^{\beta T}}{c^2}
\int_0^\infty\hspace{-5pt}{\rm d} t\,e^{-\beta t}\hspace{-5pt}
\int_0^{ct}\hspace{-5pt}{\rm d} s\int_0^{ct-s}\hspace{-10pt}{\rm d} r
\int{\rm d}\nu\,\hat{V}(s,\cdot)\,p_{r}\hat{V}(s,\cdot)$$
\begin{equation}\label{no t}
=\,\frac{2e^{\beta T}}{\beta c^2}
\int_0^{\infty}\hspace{-5pt}{\rm d} s\int_0^{\infty}\hspace{-5pt}{\rm d} r
\,e^{-\beta(s+r)/c}\hspace{-5pt}
\int{\rm d}\nu\,\hat{V}(s,\cdot)\,p_{r}\hat{V}(s,\cdot)
\end{equation}
$$=\,\frac{2e^{\beta T}}{\beta c^2}
\int_0^{\infty}\hspace{-5pt}{\rm d} s\int_0^{\infty}\hspace{-5pt}{\rm d} r
\,e^{-\beta(s+r)/c}\,
{\bf E}_\nu\,\hat{V}(s,\eta_{s})\hat{V}(s+r,\eta_{s+r})$$
$$=\,\frac{2e^{\beta T}}{\beta c^2}
\int_0^{\infty}\hspace{-5pt}{\rm d} s\int_0^{\infty}\hspace{-5pt}{\rm d} r
\,e^{-\frac{\beta}{2c}r}\,
{\bf E}_\nu\,\tilde{V}(s,\eta_{s})\tilde{V}(s+r,\eta_{s+r})$$
$$=\,\frac{2e^{\beta T}}{\beta c^2}
\int_0^{\infty}\hspace{-5pt}{\rm d} s
\int{\rm d}\nu\,\,\tilde{V}(s,\cdot)\hspace{-3pt}
\int_0^{\infty}\hspace{-5pt}{\rm d} r\,e^{-\frac{\beta}{2c}r}\,
p_{r}\tilde{V}(s,\cdot)
\,=\,
\frac{2e^{\beta T}}{\beta c^2}\,
(\tilde{V}\,|\,
G_{\!\frac{\beta}{2c}}\tilde{V})_{L^2(\ell\otimes\nu)}
\eqno\rule{6pt}{6pt}$$
\begin{rem}\rm\label{no t remark}
\begin{itemize}\item[(i)]
In the case were $V$ is time independent
it easily follows from (\ref{no t}) that
$$\int_0^T\hspace{-5pt}{\rm d} t\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V(\eta_{cs})\,{\rm d} s\,]^2
\,\le\,
\frac{2e^{\beta T}}{\beta^2 c}\,
({V}\,|\,
G_{\!\beta/c}{V})_{L^2(\nu)}$$
because $V=\hat{V}$ and
$\int_0^{\infty}\hspace{-2pt}{\rm d} s\,e^{-\beta s/c}
=c/\beta$.
\item[(ii)]
Notice that one cannot apply (i) to the process $(s,\eta_s)_{s\ge 0}$
because $\ell\otimes\nu$ is not an invariant measure for this process.
\end{itemize}
\end{rem}
The remaining part of this section deals with the problem of
estimating the right-hand side of the inequality
in Lemma \ref{lem_inhomo} by something which
is more likely to be computable in an explicit way.
Let $(G_\alpha)_{\alpha>0}$ be a strongly continuous contraction
resolvent on a real Hilbert space $H$ with inner product
$(\cdot\,|\,\cdot)$. If $G_\alpha^\ast$ denotes the adjoint of $G_\alpha$
then $(G_\alpha^\ast)_{\alpha>0}$ is also a strongly continuous
contraction resolvent on $H$.
Denote by $(L,D(L))$ and $(L^\ast,D(L^\ast))$ the generators of
$(G_\alpha)_{\alpha>0}$ and $(G_\alpha^\ast)_{\alpha>0}$, respectively.
\begin{rem}\rm
If the co-generator $(L^\ast,D(L^\ast))$ is not the adjoint of $(L,D(L))$
then $L^\ast$ coincides with the adjoint of $(L,D(L))$ on
$D(L^\ast)$, at least.
\end{rem}
The following assumption will play a
crucial role in the proof of Lemma \ref{variational} below.
$$\mbox{There exists $D_0\subseteq H$ which is a core for both
$(L,D(L))$ and $(L^\ast,D(L^\ast))$.}\eqno(A1)$$
Remark that there are unbounded operators $(L,D(L))$
on Hilbert spaces whose adjoints $(L^\ast,D(L^\ast))$ satisfy
$D(L)\cap D(L^\ast)=\{0\}$ in the worst case
hence there is something to check for this assumption to hold.
Assuming $(A1)$, the symmetric and antisymmetric parts of $L$ given by
$$L_s\,=\,\frac{L+L^\ast}{2}
\quad\mbox{and}\quad
L_a\,=\,\frac{L-L^\ast}{2}
\quad\mbox{respectively}$$
are defined on $D_0$. Of course $D_0$ is dense in $H$ since it is a core.
As $(L,D(L))$ and $(L^\ast,D(L^\ast))$ are both negative definite,
$(L_s,D_0)$ is a symmetric, negative definite, densely defined operator.
Hence it can be extended (Friedrich's extension for example)
to a self-adjoint negative definite operator $(L_s,D(L_s))$ on $H$.
Every such extension generates a strongly continuous
contraction resolvent $((\alpha-L_s)^{-1})_{\alpha\ge 0}$
of self-adjoint positive definite operators on $H$.
\begin{lemma}\label{variational}
Assume $(A1)$ and
fix both $\alpha>0$ as well as an arbitrary self-adjoint extension
of $(L_s,D_0)$. Then
$$(u\,|\,G_\alpha u)\,=\,\sup_{v\in D_0}
\left\{\rule{0pt}{12pt}\right.
2\,(u\,|\,v)-
\left(\rule{0pt}{10pt}\right.\!
(\alpha-L)v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)v
\!\left.\rule{0pt}{10pt}\right)
\left.\rule{0pt}{12pt}\right\}$$
for all $u\in H$.
\end{lemma}
{\it Proof}. Fix $u\in H$. As $D_0$ is a core for both, $(L,D(L))$ and
$(L^\ast,D(L^\ast))$, one can choose sequences $(v_n)_{n=1}^\infty$
and $(v_n^\ast)_{n=1}^\infty$ in $D_0$ such that
$$v_n\to\frac{G_\alpha}{2}u,\quad
(\alpha-L)v_n=:\frac{u_n}{2}\to\frac{u}{2},\quad
v_n^\ast\to\frac{G_\alpha^\ast}{2}u,\quad
(\alpha-L^\ast)v_n^\ast=:\frac{u_n^\ast}{2}\to\frac{u}{2},\quad$$
in $H$ when $n\to\infty$. Then
$$\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)(v_n+v_n^\ast)
\left.\rule{0pt}{10pt}\right)
\longrightarrow
(u\,|\,G_\alpha u),\quad n\to\infty.$$
Indeed
\begin{eqnarray*}
&&
\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)(v_n+v_n^\ast)
\left.\rule{0pt}{10pt}\right)\\
&=&
\left(\rule{0pt}{10pt}\right.
\frac{\alpha-L}{2}\,(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}u_n
\left.\rule{0pt}{10pt}\right)
+\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}\,\frac{\alpha-L}{2}\,2v_n^\ast
\left.\rule{0pt}{10pt}\right)
\end{eqnarray*}
where
\begin{eqnarray*}
&&
\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}\,\frac{\alpha-L}{2}\,2v_n^\ast
\left.\rule{0pt}{10pt}\right)\\
&=&
\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
2v_n^\ast
\left.\rule{0pt}{10pt}\right)
-\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}\,\frac{\alpha-L^\ast}{2}\,2v_n^\ast
\left.\rule{0pt}{10pt}\right)\\
&=&
\left(\rule{0pt}{10pt}\right.
(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
u_n^\ast
\left.\rule{0pt}{10pt}\right)
-\left(\rule{0pt}{10pt}\right.
\frac{\alpha-L}{2}\,(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}u_n^\ast
\left.\rule{0pt}{10pt}\right)
\end{eqnarray*}
hence
\begin{eqnarray*}
&&
\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)(v_n+v_n^\ast)
\left.\rule{0pt}{10pt}\right)\\
&=&
\left(\rule{0pt}{10pt}\right.
(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
u_n^\ast
\left.\rule{0pt}{10pt}\right)
+\left(\rule{0pt}{10pt}\right.
\frac{\alpha-L}{2}\,(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(u_n-u_n^\ast)
\left.\rule{0pt}{10pt}\right).
\end{eqnarray*}
Of course
$$\left(\rule{0pt}{10pt}\right.
(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
u_n^\ast
\left.\rule{0pt}{10pt}\right)
\,=\,
\left(\rule{0pt}{10pt}\right.
u_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(v_n+v_n^\ast)
\left.\rule{0pt}{10pt}\right)
\longrightarrow(u\,|\,G_\alpha u),\quad n\to\infty,$$
whereas
\begin{eqnarray*}
\left(\rule{0pt}{10pt}\right.
\frac{\alpha-L}{2}\,(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(u_n-u_n^\ast)
\left.\rule{0pt}{10pt}\right)
&=&
\left(\rule{0pt}{10pt}\right.
\frac{\alpha-L}{2}\,v_n
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(u_n-u_n^\ast)
\left.\rule{0pt}{10pt}\right)\\
&-&
\left(\rule{0pt}{10pt}\right.
\frac{\alpha-L^\ast}{2}\,v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(u_n-u_n^\ast)
\left.\rule{0pt}{10pt}\right)\\
&+&
\left(\rule{0pt}{10pt}\right.
(\alpha-L_s)v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(u_n-u_n^\ast)
\left.\rule{0pt}{10pt}\right)
\end{eqnarray*}
converges to zero when $n\to\infty$ which is obvious for the first two
terms on the right-hand side and follows from
$$\left(\rule{0pt}{10pt}\right.
(\alpha-L_s)v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(u_n-u_n^\ast)
\left.\rule{0pt}{10pt}\right)
\,=\,
\left(\rule{0pt}{10pt}\right.
v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(u_n-u_n^\ast)
\left.\rule{0pt}{10pt}\right)$$
for the last term.
Altogether one obtains that
\begin{eqnarray*}
&&(u\,|\,G_\alpha u)\\
&=&
\lim_{n\to\infty}
\left\{\rule{0pt}{12pt}\right.
\hspace{-2pt}
2\left(\rule{0pt}{10pt}\right.
u
\;\rule[-4pt]{1pt}{14pt}\;
(v_n+v_n^\ast)
\left.\rule{0pt}{10pt}\right)
-\left(\rule{0pt}{10pt}\right.
(\alpha-L)(v_n+v_n^\ast)
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)(v_n+v_n^\ast)
\left.\rule{0pt}{10pt}\right)
\hspace{-2pt}
\left.\rule{0pt}{12pt}\right\}\\
&\le&
\sup_{v\in D_0}
\left\{\rule{0pt}{12pt}\right.
2\,(u\,|\,v)-
\left(\rule{0pt}{10pt}\right.\!
(\alpha-L)v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)v
\!\left.\rule{0pt}{10pt}\right)
\left.\rule{0pt}{12pt}\right\}
\end{eqnarray*}
and it remains to show that
\begin{equation}\label{other direction}
\sup_{v\in D_0}
\left\{\rule{0pt}{12pt}\right.
2\,(u\,|\,v)-
\left(\rule{0pt}{10pt}\right.\!
(\alpha-L)v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)v
\!\left.\rule{0pt}{10pt}\right)
\left.\rule{0pt}{12pt}\right\}
\,\le\,
(u\,|\,G_\alpha u).
\end{equation}
Now choose $(v_n^\ast)_{n=1}^\infty\subseteq D_0$ such that
$$v_n^\ast\to G_\alpha^\ast u
\quad\mbox{and}\quad
(\alpha-L^\ast)v_n^\ast\to u
\quad\mbox{in $H$ when $n\to\infty$}$$
and remark that
$(\alpha-L_s)^{1/2}$ is well-defined on $D_0\subseteq D(L_s)$.
Then for every $v\in D_0$
\begin{eqnarray*}
(u\,|\,v)&=&
\left(\rule{0pt}{10pt}\right.
u
\;\rule[-4pt]{1pt}{14pt}\;
G_\alpha(\alpha-L)v
\left.\rule{0pt}{10pt}\right)
\,=\,
\left(\rule{0pt}{10pt}\right.
G_\alpha^\ast u
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L)v
\left.\rule{0pt}{10pt}\right)\\
&=&
\lim_{n\to\infty}
\left(\rule{0pt}{10pt}\right.
v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{1/2}(\alpha-L_s)^{-1/2}(\alpha-L)v
\left.\rule{0pt}{10pt}\right)\\
&=&
\lim_{n\to\infty}
\left(\rule{0pt}{10pt}\right.
(\alpha-L_s)^{1/2}v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1/2}(\alpha-L)v
\left.\rule{0pt}{10pt}\right)\\
&\le&
\lim_{n\to\infty}
\sqrt{
\left(\rule{0pt}{10pt}\right.
v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)v_n^\ast
\left.\rule{0pt}{10pt}\right)
\cdot
\left(\rule{0pt}{10pt}\right.
(\alpha-L_s)v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)v
\left.\rule{0pt}{10pt}\right)
}\\
&\le&
\lim_{n\to\infty}
\left[\rule{0pt}{11pt}\right.
\left(\rule{0pt}{10pt}\right.
v_n^\ast
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)v_n^\ast
\left.\rule{0pt}{10pt}\right)
+
\left(\rule{0pt}{10pt}\right.
(\alpha-L_s)v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)v
\left.\rule{0pt}{10pt}\right)
\left.\rule{0pt}{11pt}\right]/2\\
&=&
\left[\rule{0pt}{11pt}\right.
\left(\rule{0pt}{10pt}\right.
(u
\;\rule[-4pt]{1pt}{14pt}\;
G_\alpha u
\left.\rule{0pt}{10pt}\right)
+
\left(\rule{0pt}{10pt}\right.
(\alpha-L_s)v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)v
\left.\rule{0pt}{10pt}\right)
\left.\rule{0pt}{11pt}\right]/2
\end{eqnarray*}
since
$$(v_n^\ast\,|\,(\alpha-L_s)v_n^\ast)
=(v_n^\ast\,|\,(\alpha-L^\ast)v_n^\ast)
\longrightarrow
(G_\alpha^\ast u\,|\,u)
\,=\,
(u\,|\,G_\alpha u),\quad n\to\infty,$$
and (\ref{other direction}) follows.\hfill$\rule{6pt}{6pt}$
\begin{rem}\rm
For the proof of (\ref{other direction}), $D_0$ only needs
to be a core for $(L^\ast,D(L^\ast))$ but not for $(L,D(L))$.
Of course $D_0\subseteq D(L)$ must still be assumed.
\end{rem}
\begin{cor}\label{my inequ}
If there exists $D_0\subseteq H$
which is a core for $(L,D(L)),\,(L^\ast,D(L^\ast))$ and $(L_s,D(L_s))$ then
$$(u\,|\,G_\alpha u)
\,\le\,
\left(\rule{0pt}{10pt}\right.\!
u
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}\,u
\!\left.\rule{0pt}{10pt}\right),\quad u\in H,$$
for every $\alpha>0$.
\end{cor}
{\it Proof}.
If $D_0$ is a core for $(L,D(L)),\,(L^\ast,D(L^\ast))$ and
$(L_s,D(L_s))$ then Lemma \ref{variational} implies
$$(u\,|\,G_\alpha u)\,\le\,\sup_{v\in D_0}
\left\{\rule{0pt}{12pt}\right.
2\,(u\,|\,v)-
\left(\rule{0pt}{10pt}\right.\!
v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)v
\!\left.\rule{0pt}{10pt}\right)
\left.\rule{0pt}{12pt}\right\}
\,=\,
\left(\rule{0pt}{10pt}\right.\!
u
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}\,u
\!\left.\rule{0pt}{10pt}\right).$$
Indeed
\begin{eqnarray*}
\left(\rule{0pt}{10pt}\right.\!
(\alpha-L)v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}(\alpha-L)v
\!\left.\rule{0pt}{10pt}\right)
&=&
\left(\rule{0pt}{10pt}\right.\!
v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)v
\!\left.\rule{0pt}{10pt}\right)
+
\left(\rule{0pt}{10pt}\right.\!
L_a v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)^{-1}L_a v
\!\left.\rule{0pt}{10pt}\right)\\
&\ge&
\left(\rule{0pt}{10pt}\right.\!
v
\;\rule[-4pt]{1pt}{14pt}\;
(\alpha-L_s)v
\!\left.\rule{0pt}{10pt}\right)
\end{eqnarray*}
for $v\in D_0$ since $(\alpha-L_s)^{-1}$ is positive definite
and finally one applies Lemma \ref{variational} in the special case
$L=L_s$.\hfill$\rule{6pt}{6pt}$
\begin{rem}\rm
Not every perturbation $(L,D_0)$ of a symmetric operator
$(S,D_0)$ has $S$ as its symmetric part on $D_0$.
An easy example demonstrating this fact
will be given in the next section, see Remark \ref{not sym part}.
So one has to be careful when checking the assumptions of
Corollary \ref{my inequ}.
\end{rem}
Finally the impact of Corollary \ref{my inequ} on the situation
described in Lemma \ref{lem_inhomo} is discussed.
So $H=L^2(\ell\otimes\nu)$, $(G_\alpha)_{\alpha>0}$
is the strongly continuous contraction resolvent
associated with the right process $(s,\eta_s)_{s\ge 0}$ on $H$
and $(L,D(L))$ is the generator of $(G_\alpha)_{\alpha>0}$.
Denote by $(L^\eta,D(L^\eta))$ the generator of the strongly continuous
contraction resolvent associated with the right process
$(\eta_s)_{s\ge 0}$ on $L^2(\nu)$.
Corollary \ref{my inequ} suggests to develop the inequality
given in Lemma \ref{lem_inhomo} by using
$$(\tilde{V}\,|\,
G_{\!\frac{\beta}{2c}}\tilde{V})_{L^2(\ell\otimes\nu)}
\,\le\,
(\tilde{V}\,|\,
(\frac{\beta}{2c}-L_s)^{-1}\tilde{V})_{L^2(\ell\otimes\nu)}$$
and one wants to simplify
$(\beta/(2c)-L_s)^{-1}\tilde{V}$
in this specific situation.
Formally $L=\frac{\partial}{\partial s}+L^\eta$ and
$L^\ast=-\delta_0(s)-\frac{\partial}{\partial s}+(L^\eta)^\ast$
where $\delta_0$ denotes Dirac's delta function.
So, under certain conditions on $\tilde{V}$, one should have an
equality of the type
$
[(\alpha-L_s)^{-1}\tilde{V}](s,\eta)
\,=\,
[(\alpha-L^{\rm sym})^{-1}\tilde{V}(s,\cdot)](\eta)
\,\stackrel{\mbox{\tiny def}}{=}\,g(s,\eta)
$
where $L^{\rm sym}$ denotes
the symmetric part of $L^\eta$ in $L^2(\nu)$.
The following lemma, first, lists sufficient conditions
to ensure this and, second, presents the final
bound on the left-hand side of the inequality in Lemma \ref{lem_inhomo}.
\begin{lemma}\label{decouple}
\begin{itemize}\item[(i)]
Fix $\alpha>0$.
If $g,L^\eta g,(L^\eta)^\ast g,\frac{\partial}{\partial s}g
\in L^2(\ell\otimes\nu)$,
$g(s,\cdot)\in D(L^\eta)\cap D((L^\eta)^\ast)$, $s\ge 0$,
and
$g(\cdot,\eta)\in C_0^1([0,\infty)),\,\eta\in X$,
then
$(\alpha-L_s)^{-1}\tilde{V}\,=\,g$.
\item[(ii)] Fix $\beta>0,\,T>0,\,c>0$ and assume that there exists
$D_0\subseteq L^2(\nu)$ which is a core for
$(L^\eta,D(L^\eta))$, $((L^\eta)^\ast,D((L^\eta)^\ast))$
and $(L^{\rm sym},D(L^{\rm sym}))$. Then:
$$\int_0^T\hspace{-5pt}{\rm d} t\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V(s,\eta_{cs})\,{\rm d} s\,]^2
\,\le\,
\frac{2e^{\beta T}}{\beta c}\,
\hspace{-3pt}\int_0^T\hspace{-5pt}
\left(\rule{0pt}{10pt}\right.\!
V(s,\cdot)
\;\rule[-4pt]{1pt}{14pt}\;
(\frac{\beta}{2c}-L^{\rm sym})^{-1}V(s,\cdot)
\!\left.\rule{0pt}{10pt}\right)_{L^2(\nu)}{\rm d} s$$
for all bounded measurable functions $V$
on $[0,\infty)\times X$.
\end{itemize}
\end{lemma}
{\it Proof}. The conditions given in (i) are obvious conditions
for $L_s\,g(s,\eta)=[L^{\rm sym}g(s,\cdot)](\eta)$
to be true which indeed proves the claim.
Part (ii) only needs to be proven for
$V(s,\eta){\bf 1}_{[0,T]}(s)$ instead of $V(s,\eta)$.
The method is to approximate
$V{\bf 1}_{[0,T]}$ by `good' functions $V_n$
such that the functions $g_n$ corresponding
to $\tilde{V}_n$ satisfy the conditions of part (i).
Going through the proof of Corollary \ref{my inequ}
but using the conditions of part (i) reveals that
one {\em can} conclude that
\begin{equation}\label{one can}
(\tilde{V}_n\,|\,
G_{\!\frac{\beta}{2c}}\tilde{V}_n)_{L^2(\ell\otimes\nu)}
\,\le\,
(\tilde{V}_n\,|\,
(\frac{\beta}{2c}-L_s)^{-1}\tilde{V}_n)_{L^2(\ell\otimes\nu)}
\end{equation}
in this specific case
where, by part (i), the right-hand side is equal to
\begin{eqnarray*}
&&\int_0^\infty\hspace{-1pt}
(\tilde{V}_n(s,\cdot)\,|\,
(\frac{\beta}{2c}-L^{\rm sym})^{-1}\tilde{V}_n(s,\cdot))_{L^2(\nu)}\,{\rm d} s\\
&=&\rule{0pt}{13pt}
c\int_0^\infty\hspace{-1pt}e^{-\beta s}\,
({V}_n(s,\cdot)\,|\,
(\frac{\beta}{2c}-L^{\rm sym})^{-1}{V}_n(s,\cdot))_{L^2(\nu)}\,{\rm d} s
\end{eqnarray*}
so that
\begin{equation}\label{stand approx}
\int_0^T\hspace{-5pt}{\rm d} t\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V_n(s,\eta_{cs})\,{\rm d} s\,]^2
\,\le\,
\frac{2e^{\beta T}}{\beta c}\,
\hspace{-3pt}\int_0^\infty\hspace{-5pt}
\left(\rule{0pt}{10pt}\right.\!
V_n(s,\cdot)
\;\rule[-4pt]{1pt}{14pt}\;
(\frac{\beta}{2c}-L^{\rm sym})^{-1}V_n(s,\cdot)
\!\left.\rule{0pt}{10pt}\right)_{L^2(\nu)}{\rm d} s
\end{equation}
for all $n$ by Lemma \ref{lem_inhomo}. Taking limits when $n$ goes to
infinity in the above inequality finally proves part (ii).
Notice that the upper limit of the ${\rm d} s$-integration
can indeed be changed to $T$
because $V_n$ approximates $V{\bf 1}_{[0,T]}$.
\hfill$\rule{6pt}{6pt}$
\begin{rem}\rm\label{smooth}
\begin{itemize}\item[(i)]
Taking limits in
(\ref{stand approx}) makes clear that the approximation of
$V{\bf 1}_{[0,T]}$ by `good' functions $V_n$ is a standard
approximation of a function in $L^2(\ell\otimes\nu)$ by
in some sense `smooth' functions and the details are therefore omitted.
\item[(ii)]
Remember that the left-hand side of (\ref{one can}) can be transformed into
$$\int_0^{\infty}\hspace{-5pt}{\rm d} s\int_0^{\infty}\hspace{-5pt}{\rm d} r
\,e^{-\beta(s+r)/c}\,
{\bf E}_\nu\,{V}_n(s/c,\eta_{s}){V}_n((s+r)/c,\eta_{s+r})$$
as in the proof of Lemma \ref{lem_inhomo}.
However, it is not possible to transform the right-hand side of
(\ref{one can}) into a similar expression because the resolvent
$((\alpha-L_s)^{-1})_{\alpha>0}$ is not associated with a process of
the form $(s,\eta^{\rm sym}_s)_{s\ge 0}$.
\item[(iii)]
In the case were $V$ is time independent one can directly apply
Corollary \ref{my inequ} to the right-hand side of the inequality in
Remark \ref{no t remark}(i) which gives
$$\int_0^T\hspace{-5pt}{\rm d} t\,{\bf E}_\nu\,[\int_0^{t}\hspace{-1pt}
V(\eta_{cs})\,{\rm d} s\,]^2
\,\le\,
\frac{2e^{\beta T}}{\beta^2 c}\,
({V}\,|\,
(\beta/c-L^{\rm sym})^{-1}{V})_{L^2(\nu)}$$
for all bounded $V$ in $L^2(\nu)$ hence for all
$V\in L^2(\nu)$ by approximation.
Notice that $\nu$ is an invariant measure for
the resolvent $((\alpha-L^{\rm sym})^{-1})_{\alpha>0}$
under the conditions made.
\end{itemize}
\end{rem}
\section{Application to 1-dimensional
Simple Exclusion}\label{section application}
Fix $p,q\ge 0$ such that $p+q=1$ and
let $(\Omega,{\cal F},{\bf P}_{\!\!\eta},\eta\in\{0,1\}^\mathbb{Z},
(\eta_t)_{t\ge 0})$ denote the strong Markov Feller process
whose generator $L$
acts on local functions $f:\{0,1\}^\mathbb{Z}\to\mathbb{R}$ as
{\small
\begin{equation}\label{operator}
Lf(\eta)\,=
\sum_{x\in\mathbb{Z}}\left(
2p\,\eta(x)(1-\eta(x+1))[f(\eta^{x,x+1})-f(\eta)]
+\;2q\,\eta(x)(1-\eta(x-1))[f(\eta^{x,x-1})-f(\eta)]
\right)
\end{equation}
}
\noindent
where the operation
$$\eta^{x,y}(z)\,=\,\left\{\begin{array}{ccc}
\eta(z)&:&z\not=x,y\\
\eta(x)&:&z=y\\
\eta(y)&:&z=x
\end{array} \right.$$
exchanges the ``spins'' at $x$ and $y$.
This process is called simple exclusion process, see \cite{L1999}
for a good account on the existing theory.
Denote by $\nu_{1/2}$ the Bernoulli product
measure on $\{0,1\}^{\mathbb{Z}}$ satisfying
$\nu_{1/2}(\eta(x)=1)=1/2$ for all $x\in\mathbb{Z}$ which is one of the
invariant ergodic states of the simple exclusion process. If
$${\bf P}\,=\,\int{\bf P}_{\!\!\eta}\,{\rm d}\nu_{1/2}(\eta)
\quad\mbox{as well as}\quad
\xi_t(x)\,=\,\frac{\eta_t(x)-{\bf E}\eta_t(x)}
{\sqrt{\mbox{\bf Var}(\eta_t(x))}}$$
where ${\bf E}$ and $\mbox{\bf Var}$ stand for expectation and
variance with respect to {\bf P}, respectiveley,
then the process $(\xi_t)_{t\ge 0}$ is a stationary process on
$(\Omega,{\cal F},{\bf P})$ which takes values in $\{-1,1\}^{\mathbb{Z}}$
and the push forward of $\nu_{1/2}$ with respect to the map
$$\eta\mapsto\xi
\quad\mbox{given by}\quad
\xi(x)\,=\,\frac{\eta(x)-1/2}{\sqrt{1/4}},\;x\in\mathbb{Z},$$
is the invariant distribution of $\xi_t,\,t\ge 0$.
For $\Lambda\subseteq\mathbb{Z}$ finite, set
$\xi_\Lambda=\prod_{x\in\Lambda}\xi(x)$ if $\Lambda$ is not empty and
$\xi_\emptyset=1$ otherwise. Then,
$\{\xi_\Lambda:\mbox{$\Lambda\subseteq\mathbb{Z}$ finite}\}$ forms an
orthonormal basis of $L^2(\nu_{1/2})$.
Hence the linear hull $\mbox{\boldmath$Lin$}\{\xi_\Lambda\}$
of $\{\xi_\Lambda:\mbox{$\Lambda\subseteq\mathbb{Z}$ finite}\}$
is dense in $L^2(\nu_{1/2})$.
Remark that the operator $(L,\mbox{\boldmath$Lin$}\{\xi_\Lambda\})$
is closable on $L^2(\nu_{1/2})$
and that its closure $(L,D(L))$
generates a Markovian strongly continuous contraction semigroup
$(T_t)_{t\ge 0}$ on $L^2(\nu_{1/2})$ which is associated with the
transition semigroup of the strong Markov process
$(\Omega,{\cal F},{\bf P}_{\!\!\eta},\eta\in\{0,1\}^\mathbb{Z},
(\eta_t)_{t\ge 0})$.
Furthermore, it follows from (\ref{operator}) that
$$L\xi_\Lambda\,=\,
\gamma{\cal A}_+\xi_\Lambda-\gamma{\cal A}_+^\ast\xi_\Lambda
+{\cal S}\xi_\Lambda
\quad\mbox{where}\quad
\gamma=p-q$$
and
$$
{\cal A}_+\xi_\Lambda\,=\,
\sum_{x\in\Lambda}\left[
{\bf 1}_{\Lambda^c}(x+1)\xi_{\Lambda\cup\{x+1\}}
-{\bf 1}_{\Lambda^c}(x-1)\xi_{\Lambda\cup\{x-1\}}
\right].
$$
The adjoint operator of $({\cal A}_+,\mbox{\boldmath$Lin$}\{\xi_\Lambda\})$ with respect
to the inner product on $L^2(\nu_{1/2})$ is denoted by ${\cal A}_+^\ast$.
Its domain includes $\mbox{\boldmath$Lin$}\{\xi_\Lambda\}$ and, on this subdomain, it is
given by
$$
{\cal A}_+^\ast\xi_\Lambda\,=\,
\sum_{x\in\Lambda}\left[
{\bf 1}_{\Lambda^c}(x+1)-{\bf 1}_{\Lambda^c}(x-1)
\right]\xi_{\Lambda\setminus\{x\}}\,.
$$
The operator $({\cal S},\mbox{\boldmath$Lin$}\{\xi_\Lambda\})$ is a symmetric operator
on $L^2(\nu_{1/2})$ satisfying
$$\hspace{-7.7cm}
{\cal S}\xi_\Lambda\,=\,
{\cal S}_0\xi_\Lambda-2|\Lambda|\xi_\Lambda$$
where
\begin{eqnarray*}
{\cal S}_0\xi_\Lambda&=&
\sum_{x\in\Lambda}\left[
{\bf 1}_{\Lambda^c}(x+1)\xi_{\Lambda\setminus\{x\}\cup\{x+1\}}
+{\bf 1}_{\Lambda}(x+1)\xi_{\Lambda}\right.\\
&&\hspace{3cm}\left.
+{\bf 1}_{\Lambda^c}(x-1)\xi_{\Lambda\setminus\{x\}\cup\{x-1\}}
+{\bf 1}_{\Lambda}(x-1)\xi_{\Lambda}
\right]
\end{eqnarray*}
and $|\Lambda|$ denotes the cardinality of $\Lambda$.
As a consequence, the adjoint $(L^\ast,D(L^\ast))$ of $(L,D(L))$
satisfies
$$L^\ast\,=\,\gamma{\cal A}_+^\ast-\gamma{\cal A}_++{\cal S}
\quad\mbox{on}\quad
\mbox{\boldmath$Lin$}\{\xi_\Lambda\}$$
and $(L^\ast,D(L^\ast))$ is the closure of
$(L^\ast,\mbox{\boldmath$Lin$}\{\xi_\Lambda\})$. Thus if
$$G_\alpha\,=\,\int_0^\infty e^{-\alpha t}T_t\,{\rm d} t,
\quad\alpha>0,$$
is the Markovian strongly continuous contraction resolvent generated
by $(L,D(L))$ then $(L^\ast,D(L^\ast))$ generates $G_\alpha^\ast$
hence the condition $(A1)$ in Section \ref{section resolv method} is
satisfied. Furthermore
$$L_s\,=\,\frac{L+L^\ast}{2}\,=\,S
\quad\mbox{on}\quad
\mbox{\boldmath$Lin$}\{\xi_\Lambda\}$$
and $(S,\mbox{\boldmath$Lin$}\{\xi_\Lambda\})$ is of course essentially self-adjoint,
its closure being the generator of the symmetric simple exclusion process.
Altogether the domain $\mbox{\boldmath$Lin$}\{\xi_\Lambda\}$ satisfies
all the conditions imposed on the domain $D_0$ in Corollary \ref{my inequ}.
\begin{rem}\rm\label{not sym part}
The operator $(\gamma{\cal A}_++{\cal S},\mbox{\boldmath$Lin$}\{\xi_\Lambda\})$
is an easy example of an operator which, when seen as a perturbation of
$({\cal S},\mbox{\boldmath$Lin$}\{\xi_\Lambda\})$
does not have $S$ as its symmetric part on $\mbox{\boldmath$Lin$}\{\xi_\Lambda\}$.
\end{rem}
The example field of type (\ref{field}) discussed in this paper
is the density fluctuation field
$$Y_s^\varepsilon\,=\,\sqrt{\varepsilon}
\sum_{x\in\mathbb{Z}}\xi_{s\varepsilon^{-2}}(x)
\delta_{\varepsilon x},\quad s\ge 0,$$
in the case of $\sqrt{\varepsilon}$-asymmetric one-dimensional simple
exclusion where $p$ and $q$ hence $\gamma$ depend on $\varepsilon$
such that
$$\gamma\,=\,\gamma_\varepsilon\,=\,\tilde{\gamma}\cdot\sqrt{\varepsilon}.$$
As a consequence $L,T_t,G_\alpha$ and ${\bf E}$ introduced above
are denoted by
$L_\varepsilon,T_t^\varepsilon,G_\alpha^\varepsilon$ and
${\bf E}_\varepsilon$ in what follows.
If $\varepsilon$ is small then the field
has a related approximate equation of type (\ref{approx equ})
with ${\cal A}$ being the one-dimensional Laplacian and with
$V_\varepsilon^G(s,\cdot)$
being of the special form $\tilde{\gamma}V_\varepsilon^G$ where
$$V_\varepsilon^G(\xi)\,=\,
-\sum_{x\in\mathbb{Z}}G^\prime(\varepsilon x)\xi(x)\xi(x+1).$$
Functions of this type are called quadratic fluctuations in this
paper, {\it quadratic} since its summands are of type $\xi_\Lambda$
with $|\Lambda|=2$ and {\it fluctuations} since $\xi$ is
$(\eta-1/2)/\sqrt{1/4}$. Remark that $V_\varepsilon^G$
does not depend on $s$ because of the special choice of $\nu_{1/2}$
to be the invariant state of the system. However, the choice of a
different invariant state would only make the notation more
complicated but would not affect the arguments used below.
Furthermore, $V_\varepsilon^G$ remains bounded if $G$ is a test
function in the Schwartz space ${\mathscr S}(\mathbb{R})$ and this space of
test functions is chosen in what follows.
The functional $F$ used to replace
$$\int_0^t V_\varepsilon^G(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\quad\mbox{by}\quad
\int_0^t F(Y_s^\varepsilon,G)\,{\rm d} s$$
depends on a further parameter $N$, that is $F=F_N$ in this example.
It can be defined as follows.
Fix an \underline{even}
non-negative test function $d$ satisfying
$$\mbox{supp}\,d=[-1,1]\quad\mbox{and}\quad
\int_\mathbb{R} d(u)\,{\rm d} u=1$$
and denote by $d_N$ the function $x\mapsto Nd(Nx),\,N\ge 1$.
Remark that the convolution $Y^{\varepsilon}_s\star d_N$ is a
$C^\infty$-function on $\mathbb{R}$ satisfying
$$\int_\mathbb{R} H(u)\,{\rm d} Y^{\varepsilon}_s(u)
\,=\,\lim_{N\uparrow\infty}
\int_\mathbb{R} H(u)(Y^{\varepsilon}_s\star d_N)(u)\,{\rm d} u$$
for all $s\ge 0$ and all continuous functions $H$ on $\mathbb{R}$
with sufficiently fast decaying tails.
Now set
$$F_N({\cal Y},G)\,=\,
-\int_\mathbb{R} G^\prime(u)({\cal Y}\star d_N)^2(u)\,{\rm d} u
,\quad{\cal Y}\in{\mathscr S}^\prime(\mathbb{R}).$$
Fixing a finite time horizon $T>0$,
one wants to estimate
$$\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\varepsilon\left(
\int_0^t F_N(Y_s^\varepsilon,G)\,{\rm d} s
\,-\,
\int_0^t V_\varepsilon^G(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}$$
for small $\varepsilon$ and large $N$.
So fix $\varepsilon,N$ and observe that
$$\int_0^t F_N(Y_s^\varepsilon,G)\,{\rm d} s
\,-\,
\int_0^t V_\varepsilon^G(\xi_{s\varepsilon^{-2}})\,{\rm d} s$$
$$=-\int_0^t\!\int_\mathbb{R}
G^\prime(u)(Y^{\varepsilon}_s\star d_N)^2(u)\,{\rm d} u{\rm d} s
\,
\int_0^
\sum_{x\in\mathbb{Z}}G^\prime(\varepsilon x)
\xi_{s\varepsilon^{-2}}(x)\xi_{s\varepsilon^{-2}}(x+1)
\,{\rm d} s$$
hence, setting $H=-G^\prime$, one can split
$$\int_0^t F_N(Y_s^\varepsilon,G)\,{\rm d} s
\,-\,
\int_0^t V_\varepsilon^G(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\quad\mbox{into}\quad
\sum_{i=1}^4
\int_0^
V_{\varepsilon,N}^{H,i}(\xi_{s\varepsilon^{-2}})\,{\rm d} s$$
using further quadratic fluctuations
$V_{\varepsilon,N}^{H,1},V_{\varepsilon,N}^{H,2},V_{\varepsilon,N}^{H,3},
V_{\varepsilon,N}^{H,4}$
given by
\begin{eqnarray*}
V_{\varepsilon,N}^{H,1}(\xi)&=&
\sum_{x\in\mathbb{Z}}
\int_\mathbb{R}
[H(u)-H(\varepsilon x)]
d_N(u-\varepsilon x)
\sum_{\tilde{x}\in\mathbb{Z}}
\varepsilon
d_N(u-\varepsilon\tilde{x})\,{\rm d} u\;
\xi(x)\xi(\tilde{x}),\\
V_{\varepsilon,N}^{H,2}(\xi)&=&
\varepsilon
\sum_{x\in\mathbb{Z}}H(\varepsilon x)
\int_\mathbb{R} d_N^2(u-\varepsilon x)\,{\rm d} u\;
\xi(x)[\xi(x)-\xi(x+1)],\\
V_{\varepsilon,N}^{H,3}(\xi)&=&
\varepsilon
\sum_{x\not=\tilde{x}}H(\varepsilon x)
\int_\mathbb{R} d_N(u-\varepsilon x)d_N(u-\varepsilon\tilde{x})\,{\rm d} u\;
\xi(x)[\xi(\tilde{x})-\xi(x+1)],\\
V_{\varepsilon,N}^{H,4}(\xi)&=&
\sum_{x\in\mathbb{Z}}H(\varepsilon x)
\int_\mathbb{R} d_N(u-\varepsilon x)
\hspace{-3pt}\left[
\sum_{\tilde{x}\in\mathbb{Z}}\varepsilon d_N(u-\varepsilon\tilde{x})-1
\right]\!{\rm d} u\;
\xi(x)\xi(x+1).
\end{eqnarray*}
Hence, if $(\cdot\,|\,\cdot)$ denotes the inner product on
$L^2(\nu_{1/2})$ then
\begin{eqnarray}
&&{\bf E}_\varepsilon\left(
\int_0^t F_N(Y_s^\varepsilon,G)\,{\rm d} s
\,-\,
\int_0^{t
V_\varepsilon^G(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}\nonumber\\
&\le&\rule{0pt}{30pt}
4\sum_{i=1}^4
{\bf E}_\varepsilon\left
\int_0^
V_{\varepsilon,N}^{H,i}(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}\label{four summands}\\
&\le&\rule{0pt}{30pt}
\label{split difference}
8\sum_{i=1}^3\,
\varepsilon^4\hspace{-4pt}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}{\rm d} s
\int_0^s{\rm d} r\,
(V_{\varepsilon,N}^{H,i}\,|\,T_r^\varepsilon V_{\varepsilon,N}^{H,i})
\,+\,
4\,{\bf E}_\varepsilon\left(\varepsilon^2\hspace{-4pt}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}
V_{\varepsilon,N}^{H,4}(\xi_s)\,{\rm d} s
\right)^{\hspace{-2pt}2}
\end{eqnarray}
for all $t\ge 0$ by the Markov property.
In what follows, $\|H\|_p$
denotes the norm of a test function $H$ in $L^p(\mathbb{R}),\,1\le p\le\infty$.
Then, as
$|\sum_{\tilde{x}\in\mathbb{Z}}\varepsilon d_N(u-\varepsilon\tilde{x})-1|
\le\varepsilon^2 N^2\|d^{\prime\prime}\|_\infty$ is easily realised,
one obtains that
\begin{equation}\label{estiV4}
{\bf E}_\varepsilon\left(\varepsilon^2\hspace{-4pt}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}
V_{\varepsilon,N}^{H,4}(\xi_s)\,{\rm d} s
\right)^{\hspace{-2pt}2}
\le\,
\varepsilon^2 N^4\|d^{\prime\prime}\|_\infty^2\|H\|_1^2\cdot t^2,
\quad t\ge 0.
\end{equation}
The other three integrals in (\ref{split difference})
are much harder to control. But the following lemma gives bounds for
these integrals if $T_r^\varepsilon$ is substituted by the semigroup
$T_r^{\rm sym}$ associated with the symmetric simple exclusion
process.
\begin{lemma}\label{esti_sym}
Let $H\in{\mathscr S}(\mathbb{R})$ be a test function such that
$\int_\mathbb{R} H(u)\,{\rm d} u=0$. Then
{\small
$$\begin{array}{crcl}
(i)\hspace{.2cm}&\displaystyle
\varepsilon^4\hspace{-4pt}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}{\rm d} s
\int_0^s{\rm d} r\,
(V_{\varepsilon,N}^{H,1}\,|\,T_r^{\rm sym}V_{\varepsilon,N}^{H,1})
&\le&\displaystyle
c_0\,t^2
\left(\rule{0pt}{11pt}
\|d\|_2^4\,\frac{\|(1+u^2)H^{\prime\prime}\|_\infty^2}{N^2}
+\|d\|_\infty^2\,\frac{\|(1+u^2)H^\prime\|_\infty^2}{N}
\right)\\
\rule{0pt}{30pt}
(ii)\hspace{.2cm}&\displaystyle
\varepsilon^4\hspace{-4pt}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}{\rm d} s
\int_0^s{\rm d} r\,
(V_{\varepsilon,N}^{H,2}\,|\,T_r^{\rm sym}V_{\varepsilon,N}^{H,2})
&\le&\displaystyle
t^2\,\|d\|_2^4\,
\left(\rule{0pt}{11pt}
\varepsilon^2 N^2\,\|H^\prime\|_\infty^2
+\varepsilon N^2\,\|(1+u^2)H\|_\infty^2
\right)\\
\rule{0pt}{30pt}
(iii)\hspace{.2cm}&\displaystyle
\varepsilon^4\hspace{-4pt}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}{\rm d} s
\int_0^s{\rm d} r\,
(V_{\varepsilon,N}^{H,3}\,|\,T_r^{\rm sym}V_{\varepsilon,N}^{H,3})
&\le&\displaystyle
c_0\,t^2\,\|d\|_\infty^2\,\frac{\|(1+u^2)H\|_\infty^2}{N^{1/3}}
\end{array}$$
}
\noindent
for all $t\ge 0,\,N=1,2,\dots,\,\varepsilon>0$
where $c_0$ is a constant which neither depends
on the chosen test function $H$ nor on the mollifier $d$.
\end{lemma}
{\it Proof}. For showing $(i)$ one splits $V_{\varepsilon,N}^{H,1}$
into two sums
\begin{eqnarray}\label{splitV1}
&&\varepsilon
\sum_{x\in\mathbb{Z}}
\int_\mathbb{R}
[H(u)-H(\varepsilon x)]
d_N^2(u-\varepsilon x)\,{\rm d} u\nonumber\\
&+&
\sum_{x\not=\tilde{x}}
\int_\mathbb{R}
[H(u)-H(\varepsilon x)]
d_N(u-\varepsilon x)\,
\varepsilon d_N(u-\varepsilon\tilde{x})\,{\rm d} u\;
\xi(x)\xi(\tilde{x}).
\end{eqnarray}
Applying the Taylor expansion
$$H(u)-H(\varepsilon x)\,=\,(u-\varepsilon x)H^\prime(\varepsilon x)
\,+\,(u-\varepsilon x)^2 H^{\prime\prime}(\theta_{\varepsilon x}^u)/2
\quad\mbox{with}\quad
\theta_{\varepsilon x}^u\in[\varepsilon x-u,\varepsilon x+u]$$
to the first sum yields
$
\left|
\varepsilon
\sum_{x\in\mathbb{Z}}
\int_\mathbb{R}
[H(u)-H(\varepsilon x)]
d_N^2(u-\varepsilon x)\,{\rm d} u
\right|
\,\le\,
\varepsilon\hspace{-4pt}
\sum_{x\in\mathbb{Z}}\int_\mathbb{R}
\frac{(u-\varepsilon x)^2}{2}
|H^{\prime\prime}(\theta_{\varepsilon x}^u)|
d_N^2(u-\varepsilon x)\,{\rm d} u\\
$
where $\int_\mathbb{R}(u-\varepsilon x)d_N^2(u-\varepsilon x){\rm d} u$
vanishes because the mollifier $d$ is even.
Now observe that
$$|H^{\prime\prime}(\theta_{\varepsilon x}^u)|
\,\le\left\{\begin{array}{rcl}
\sup_{\tilde{u}}|(1+\tilde{u}^2)H^{\prime\prime}(\tilde{u})|
\cdot[1+(\varepsilon x-\frac{1}{N})^2]^{-1}&:&x>1/\varepsilon\\
\rule{0pt}{20pt}
\sup_{\tilde{u}}|(1+\tilde{u}^2)H^{\prime\prime}(\tilde{u})|
\cdot[1+(\varepsilon x+\frac{1}{N})^2]^{-1}&:&x<-1/\varepsilon
\end{array}\right.$$
on the set $\{u\in\mathbb{R}:d_N(u-\varepsilon x)\not=0\}$
since ${\rm supp}\,d_N=[-\frac{1}{N},\frac{1}{N}]$.
Furthermore
$$(u-\varepsilon x)^2\,\le\,1/N^2
\quad\mbox{on}\quad\{u\in\mathbb{R}:d_N(u-\varepsilon x)\not=0\}$$
and
$$\int_\mathbb{R} d_N^2(u-\varepsilon x)\,{\rm d} u\,=\,N\|d\|_2^2
\quad\mbox{for all $x\in\mathbb{Z}$.}$$
Therefore, the sum immediately above (\ref{splitV1})
can be estimated by
{\small
$$\frac{\|d\|_2^2}{2N}\left(\rule{0pt}{13pt}\right.
\|(1+{u}^2)H^{\prime\prime}\|_\infty
\overbrace{
\sum_{x<-\frac{1}{\varepsilon}}\varepsilon
[1+(\varepsilon x+\frac{1}{N})^2]^{-1}}^{\le \pi/2}
+\;2\|H^{\prime\prime}\|_\infty\,+\;
\|(1+{u}^2)H^{\prime\prime}\|_\infty
\overbrace{
\sum_{x>\frac{1}{\varepsilon}}\varepsilon
[1+(\varepsilon x-\frac{1}{N})^2]^{-1}}^{\le \pi/2}
\left.\rule{0pt}{13pt}\right)$$
}
\begin{equation}\label{with pi}
\le\,
\frac{\|d\|_2^2}{2N}\,(\pi+2)\|(1+{u}^2)H^{\prime\prime}\|_\infty
\end{equation}
which explains the first summand on the right-hand side of $(i)$.
The second summand is a bound of the integral on the left-hand side
of $(i)$ but with $V_{\varepsilon,N}^{H,1}$ replaced by the
fluctuation field given by (\ref{splitV1}). This bound
is obtained by copying the proof of Lemma 1 in \cite{A2007} for
$x_0=1$ using the equality
$$H(u)-H(\varepsilon x)\,=\,(u-\varepsilon x)
H^\prime(\tilde{\theta}_{\varepsilon x}^u)$$
where
$\tilde{\theta}_{\varepsilon x}^u\in[\varepsilon x-u,\varepsilon x+u]$.
The only difference to the proof of Lemma 1 in \cite{A2007} is that,
similar to how (\ref{with pi}) was derived, the sum
$\sum_{x\in\mathbb{Z}}\varepsilon|H^\prime(\tilde{\theta}_{\varepsilon x}^u)|$
is estimated by
$(\pi+2)\sup_{\tilde{u}}|(1+\tilde{u}^2)H^{\prime}(\tilde{u})|$
and not by $c_H\|H^\prime\|_\infty$.
The left-hand side of (ii) can be estimated
the same way the sum $S_1(t,\varepsilon,N)$ was estimated
in the proof of Theorem 1 in \cite{A2007}.
Following this proof would give
$$\varepsilon^4\hspace{-4pt}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}{\rm d} s
\int_0^s{\rm d} r\,
(V_{\varepsilon,N}^{H,2}\,|\,T_r^{\rm sym}V_{\varepsilon,N}^{H,2})
\,\le\,
t^2\,\|d\|_2^4\,
\left(\rule{0pt}{11pt}
N^2
(\sum_{x\in\mathbb{Z}}\varepsilon H(\varepsilon x))^2
+\varepsilon N^2\,c_H\|H\|_\infty^2
\right)$$
if $H$ were a test function with compact support.
But this implies $(ii)$ because,
by our assumption $\int_\mathbb{R} H(u)\,{\rm d} u=0$,
it holds that
$|\sum_{x\in\mathbb{Z}}\varepsilon H(\varepsilon x)|
\le\varepsilon\|H^\prime\|_\infty$ by our assumption
$\int_\mathbb{R} H(u)\,{\rm d} u=0$. Again, as in the proof of part $(i)$
above, the bound $c_H\|H\|_\infty$ is replaced by
$\|(1+u^2)H\|_\infty$.
Finally, the inequality $(iii)$ can be established
by copying the proof of Theorem 1 in \cite{A2007}
with respect to $S_{6-9}(t,\varepsilon,N)$ for $x_0=1$ and $\alpha=2/3$
manipulating the constant $c_H$ accordingly.\hfill$\rule{6pt}{6pt}$\\
The next lemma is the key result of this section. It translates the
estimates given by Lemma \ref{esti_sym} into estimates of the summands
in (\ref{four summands}) on page \pageref{four summands}
when integrating them against ${\rm d} t$ over $t\in[0,T]$.
\begin{lemma}\label{key result}
Let $H\in{\mathscr S}(\mathbb{R})$ be a test function such that
$\int_\mathbb{R} H(u)\,{\rm d} u=0$ and fix a finite time horizon $T>0$. Then
$$\begin{array}{crcl}
(i)\hspace{.2cm}&\displaystyle
\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\varepsilon\left
\int_0^
V_{\varepsilon,N}^{H,1}(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}
&\le&\displaystyle
e^T C_d
\left(\rule{0pt}{11pt}
\frac{\|(1+u^2)H^{\prime\prime}\|_\infty^2}{N^2}
+\frac{\|(1+u^2)H^\prime\|_\infty^2}{N}
\right)\\
\rule{0pt}{30pt}
(ii)\hspace{.2cm}&\displaystyle
\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\varepsilon\left
\int_0^
V_{\varepsilon,N}^{H,2}(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}
&\le&\displaystyle
e^T C_d
\left(\rule{0pt}{11pt}
\varepsilon^2 N^2\,\|H^\prime\|_\infty^2
+\varepsilon N^2\,\|(1+u^2)H\|_\infty^2
\right)\\
\rule{0pt}{30pt}
(iii)\hspace{.2cm}&\displaystyle
\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\varepsilon\left
\int_0^
V_{\varepsilon,N}^{H,3}(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}
&\le&\displaystyle
e^T C_d
\,\frac{\|(1+u^2)H\|_\infty^2}{N^{1/3}}\\
\rule{0pt}{30pt}
(iv)\hspace{.2cm}&\displaystyle
\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\varepsilon\left
\int_0^
V_{\varepsilon,N}^{H,4}(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}
&\le&\displaystyle
T^3 C_d
\,\varepsilon^2 N^4\|H\|_1^2
\end{array}$$
for all $N=1,2,\dots,\,\varepsilon>0$ where $C_d$ is a constant
which only depends on the choice of the mollifier $d$.
\end{lemma}
{\it Proof}.
Choose $\beta=1$ and consider $i=1,2,3$ first.
Then, applying the inequality in Remark \ref{smooth}(iii)
with respect to $c=\varepsilon^{-2}$, one obtains that
$$\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\varepsilon\left
\int_0^
V_{\varepsilon,N}^{H,i}(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}
\,\le\,
2e^{T}\varepsilon^2\,
(V_{\varepsilon,N}^{H,i}\,|\,
(\varepsilon^2-L^{\rm sym})^{-1}V_{\varepsilon,N}^{H,i})$$
where
$$(V_{\varepsilon,N}^{H,i}\,|\,
(\varepsilon^2-L^{\rm sym})^{-1}V_{\varepsilon,N}^{H,i})
\,=\,
\varepsilon^2
\int_0^\infty\hspace{-3pt}{\rm d} t\,e^{-t}
\int_0^{t\varepsilon^{-2}}\hspace{-12pt}{\rm d} s
\int_0^s{\rm d} r\,
(V_{\varepsilon,N}^{H,i}\,|\,T_r^{\rm sym}V_{\varepsilon,N}^{H,i})$$
by a calculation similar to the proof of Lemma \ref{lem_inhomo}
in the time independent case.
Hence (i)-(iii) follows from Lemma \ref{esti_sym} since
$\int_0^\infty\hspace{-1pt}t^2e^{-t}\,{\rm d} t$ is finite.
Finally, (iv) follows directly from (\ref{estiV4}) by integration against
${\rm d} t$ over $t\in[0,T]$.\hfill$\rule{6pt}{6pt}$\\
Applying Lemma \ref{key result} in the case where $H$ is taken to be
$-G^\prime$ immediately gives the corollary below. Notice that
$\int_\mathbb{R} H(u)\,{\rm d} u=0$ is of course satisfied for $H=-G^\prime$.
\begin{cor}\label{weak replacement}
Fix an arbitrary but finite time horizon $T>0$.
Then, for every smooth test function $G\in{\mathscr S}(\mathbb{R})$, it holds that
$$\lim_{\rule{0pt}{8pt}N\uparrow\infty}\;
\limsup_{\varepsilon\downarrow 0}
\int_0^T\hspace{-3pt}{\rm d} t\,
{\bf E}_\varepsilon\left(
\int_0^t F_N(Y_s^\varepsilon,G)\,{\rm d} s
\,-\,
\int_0^t V_\varepsilon^G(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\right)^{\hspace{-2pt}2}
=\,0$$
where
$$F_N({\cal Y},G)\,=\,
-\int_\mathbb{R} G^\prime(u)({\cal Y}\star d_N)^2(u)\,{\rm d} u
\quad\mbox{and}\quad
V_\varepsilon^G(\xi)\,=\,
-\sum_{x\in\mathbb{Z}}G^\prime(\varepsilon x)\xi(x)\xi(x+1)$$
for ${\cal Y}\in{\mathscr S}^\prime(\mathbb{R})$ and $\xi\in\{-1,1\}^\mathbb{Z}$, respectively.
\end{cor}
\begin{rem}\rm
A replacement of
$$\int_0^t V_\varepsilon^G(\xi_{s\varepsilon^{-2}})\,{\rm d} s
\quad\mbox{by}\quad
\int_0^t F_N(Y_s^\varepsilon,G)\,{\rm d} s$$
for every $t\in[0,T]$ and not only for an average over $t\in[0,T]$ was shown
in \cite{JG2010} for a slightly different functional $F_N$. However it
has been shown in \cite{A2012} that many of the conclusions drawn
from the stronger
replacement result\footnote{This replacement result was called
`Second-order Boltzmann-Gibbs principle'.}
in \cite{JG2010} can actually be
obtained by only applying the weaker replacement result
of Corollary \ref{weak replacement}.
\end{rem}
|
\section{Introduction}
In hole doped high-$T_c$ cuprates, the hole density $p$ in CuO$_2$ planes has been determined by various methods: indirect chemical methods like solid solutions \cite{Shafer,Torrance,Kishino,Tokura}, bond valence sums determined from structural bond lengths \cite{Brown,TallonBV,Cava}, or the Fermi surface topology \cite{Kordyuk}. Moreover, the thermoelectric power is a universal function of $p$ \cite{Obertelli,Tallon}, and the phase diagram for hole doped cuprates is well described by $T_c$=$T_{c,max}[1-82.6(p-0.16)^2]$ \cite{Presland}, which is applicable for the estimation of $p$ when suitable structural data are not available.
These methods are, however, inapplicable when we deal with multilayered compounds because they are composed of more than two inequivalent CuO$_2$ planes in a unit cell (see Fig.~\ref{fig:n3NMR}(a) as an example). In multilayered cuprates, doped hole carriers are inequivalent between outer CuO$_2$ planes (OPs) and inner CuO$_2$ planes (IPs) due to the imbalance of the Madelung potential on each CuO$_2$ plane; the above-mentioned methods would evaluate not {\it each hole density} inherent to CuO$_2$ planes but a {\it total hole density}.
Precise determination of hole density on respective CuO$_2$ planes is indispensable for gaining deep insight into the phenomena observed in multilayered cuprates \cite{Tokunaga,Kotegawa2001,Mukuda2008,ShimizuJPSJ,ShimizuFIN}.
It is known that the spin part of the Knight shift $K_s$(300 K) at room temperature increases with $p$ in hole doped cuprates \cite{Kotegawa2001,Walstedt,Ohsugi,Ishida,FujiwaraJPSJ,MagishiTl1212,Storey}.
The $K_s$(300 K) for OP and IP are separately determined by Cu-NMR. Therefore, each value of $p$ for OP and for IP can be estimated if the relationship between $K_s$(300 K) and $p$ is established.
Meanwhile, a linear equation $p$=0.502$K_s$(300 K)+0.0462 has been reported \cite{Kotegawa2001,TokunagaJLTP}, where $p$ was derived from the NQR frequencies of Cu and O in CuO$_2$ planes \cite{Zheng}.
The values of $p$ estimated by the linear equation, however, seem relatively large. For example, an optimal doping level has been evaluated to be $p$ $\sim$ 0.22-0.24 in five-layered Hg-based compounds \cite{Mukuda2008}, the value of which is larger than a widely-accepted optimal doping level in hole-doped cuprates, $p$ $\sim$ 0.16. This inconsistency is probably, in part, due to the calculation that connects NQR frequencies to $p$ \cite{Haase}.
In addition to that, it is nontrivial whether the $p$-dependence of $K_s$(300 K) is described with a single equation which holds among different materials. $K_s$ is proportional to spin susceptibility $\chi_s$ and hyperfine coupling constants $A_{hf}$ as $K_s$=$A_{hf} \chi_s$, where the supertransferred magnetic field $B$ in $A_{hf}$=$A$+4$B$ depends on the materials \cite{MR}.
In this paper, we report that the planar CuO$_2$ hole density $p$ in high-$T_c$ cuprates is consistently evaluated from the spin part of the $^{63}$Cu-Knight shift $K_s$(300 K) at room temperature.
We show the $p$ dependences of $K_s$(300K) and of $B$ in bi-layered Ba$_2$CaCu$_2$O$_4$(F,O)$_2$ (0212F), together with other single- and bi-layered materials with tetragonal symmetry.
NMR data on three-layered Ba$_2$Ca$_2$Cu$_3$O$_6$(F,O)$_2$ (0223F), HgBa$_2$Ca$_2$Cu$_3$O$_{8+\delta}$ (Hg1223) \cite{MagishiHg1223}, and HgBa$_2$Ca$_4$Cu$_5$O$_{12+\delta}$ (Hg1245) \cite{Kotegawa2004} suggest that the $p$-dependencies of $K_s$(300 K) and of $B$ hold in multilayered compounds as well.
These results show that the present method, based on the Knight shift, enables us to separately estimate $p$ for each CuO$_2$ plane in multilayered high-$T_c$ cuprates.
\begin{table*}[htbp]
\caption[]{List of samples measured in this study, bi-layered Ba$_2$CaCu$_2$O$_4$(F,O)$_2$ (0212F) and three-layered Ba$_2$Ca$_2$Cu$_3$O$_6$(F,O)$_2$ (0223F). $T_c$ was determined by the onset of SC diamagnetism using a dc SQUID magnetometer. The values of $p$ for 0212F are evaluated using $T_c$=$T_{c,max}[1-82.6(p-0.16)^2]$ \cite{Presland}. For three-layered 0223F, two values of $K_{s,ab}$(300 K) and of $B$ correspond to OP/IP.}
\label{samples}
\begin{center}
{\tabcolsep = 3.5mm
\renewcommand\arraystretch{1.2}
\begin{tabular}{cccccc}
\hline\hline
Sample &$T_c$(K)& $p$ & $K_{s,ab}$(300 K) ($\%$) & $B$(kOe) \\
\hline
0212F($\sharp$1) & 102 & 0.174 & 0.38 & 80 \\
0212F($\sharp$2) & 105 & 0.149 & 0.33 & 74 \\
0212F($\sharp$3) & 73 & 0.114 & 0.25 & 67 \\
0212F($\sharp$4) & 40 & 0.085 & 0.20 & 67 \\
\hline
0223F & 120 & & 0.349/0.275 & 72/67 \\
\hline\hline
\end{tabular}}
\end{center}
\label{t:ggg}
\end{table*}
\section{Experimental details}
Polycrystalline powder samples of Ba$_2$CaCu$_2$O$_4$(F,O)$_2$ (0212F) and Ba$_2$Ca$_2$Cu$_3$O$_6$(F,O)$_2$ (0223F), which are listed in Table \ref{t:ggg}, were prepared by high-pressure synthesis, as described elsewhere \cite{Shirage,Iyo1,Iyo2}. The crystal structures are shown in Fig.~\ref{fig:n2NMR}(a) and in Fig.~\ref{fig:n3NMR}(a).
Powder X-ray diffraction analysis shows that these compounds comprise almost a single phase, and that the $a$-axis length continually decreases with an increase in the nominal fraction of O$^{2-}$ \cite{Shirage}.
$T_c$ was uniquely determined by the onset of SC diamagnetism using a dc SQUID magnetometer.
Four 0212F samples exhibit a systematic change in $T_c$, as the nominal fraction of oxygen O$^{2-}$ decreases at the apical fluorine F$^{1-}$ sites, i.e., the hole doping level decreases. Note that the actual fraction of F$^{1-}$ and O$^{2-}$ is difficult to determine \cite{Shirage,ShimizuPRB,ShimizuPRL}. In Table \ref{t:ggg}, the hole density $p$ is evaluated by using $T_c$=$T_{c,max}[1-82.6(p-0.16)^2]$ \cite{Presland}.
For NMR measurements, the powder samples were aligned along the $c$-axis in an external field $H$ of $\sim$ 16 T and fixed using stycast 1266 epoxy. The NMR experiments were performed by a conventional spin-echo method in the temperature ($T$) range of 4.2 $-$ 300 K with $H$ perpendicular or parallel to the $c$-axis. The width of the first exciting $\pi$/2-pulse was 6 $\mu$s. The $H$ was calibrated by using the $^{27}$Al Free Induction Decay signal.
\section{Results}
\subsection{Bi-layered Ba$_2$CaCu$_2$O$_4$(F,O)$_2$}
\subsubsection{$^{63}$Cu-NMR with $H\parallel ab$}
\begin{figure}[htpb]
\begin{center}
\includegraphics[width=1.0\linewidth]{Fig1.eps}
\end{center}
\caption{\footnotesize (color online) (a) Crystal structure of bi-layered Ba$_2$CaCu$_2$O$_4$(F,O)$_2$ (0212F). The heterovalent substitution of O$^{2-}$ for F$^{1-}$ increases the hole density. (b) $^{63}$Cu-NMR spectra for 0212F($\sharp$1) at $T$=60 K with $H$ perpendicular to the $c$-axis ($H\parallel ab$). The dashed line points to $K$=0. (c) $^{63}$Cu-NMR spectra for 0212F($\sharp$1) at $T$=60 K with $H$ parallel to the $c$-axis ($H\parallel c$). The peak marked with an asterisk ($\ast$) is from unoriented grains with $\theta$ $\sim$ 90$^{\circ}$, where $\theta$ is the angle between $H$ and the $c$-axis.
If $\theta$ in the unoriented grains were under the completely random distribution, the peak ($\ast$) could coincide with the spectral peak in (b). (d) Cu-NQR spectrum at $T$=1.5 K. The spectrum is composed of $^{63}$Cu and $^{65}$Cu.}
\label{fig:n2NMR}
\end{figure}
Figure~\ref{fig:n2NMR}(b) shows a typical $^{63}$Cu-NMR spectrum of the central transition (1/2 $\Leftrightarrow$ $-$1/2) for 0212F($\sharp$1), which has the largest nominal O$^{2-}$ composition among the bi-layered samples used in this study. The field-swept NMR spectrum was measured at $H$ perpendicular to the $c$-axis ($H\parallel ab$). Here, the NMR frequency $\omega_0$ was fixed at 174.2 MHz.
According to the second order perturbation theory for the nuclear Hamiltonian \cite{Abragam,TakigawaNQRshift}, total NMR shifts consist of the Knight shift $K_{ab}$ with $H\parallel ab$ and the second order quadrupole shift as
\begin{equation}
\frac{\omega_0 - \gamma_N H_{res}}{\gamma_N H_{res}} = K_{ab}+\frac{3\nu_Q^2}{16(1+K_{ab})}\frac{1}{(\gamma_N H_{res})^2}~,
\label{eq:shift}
\end{equation}
where $\gamma_N$ is a nuclear gyromagnetic ratio, $H_{res}$ an NMR resonance field, and $\nu_Q$ the nuclear quadrupole frequency. In order to subtract $K_{ab}$ from the total NMR shift, we estimated the $\nu_Q$ by nuclear quadrupole resonance (NQR) measurements at $H$=0 T and $T$=1.5 K. The NQR spectrum for 0212F($\sharp$1) is shown in Fig.~\ref{fig:n2NMR}(d), which has the $^{63}$Cu and $^{65}$Cu components. The $^{63}\nu_Q$ of $^{63}$Cu is estimated to be 16.5 MHz by the spectral fitting shown in the figure. For other bi-layered compounds, the values of $^{63}\nu_Q$ are 15.7, 13.7, and 12.5 MHz for 0212F($\sharp$2), ($\sharp$3), and ($\sharp$4), respectively.
Figure \ref{fig:n2KS}(a) shows the spin part of the Knight shift, $K_{s,ab}(T)$, as a function of temperature.
The Knight shift $K$ in high-$T_c$ cuprates comprises a $T$-dependent spin part $K_s(T)$ and a $T$-independent orbital part $K_{orb}$ as follows:
\begin{equation}
K_\alpha = K_{s,\alpha}(T)+K_{orb,\alpha} ~~~(\alpha =c,ab),
\label{eq:K}
\end{equation}
where $\alpha$ is the direction of $H$.
We estimate $K_{orb,ab}$ $\simeq$ 0.23 $\%$, assuming $K_{s,ab}$ $\simeq$ 0 at $T$=0 limit. The value of $K_{orb,ab}$ is consistent with those in other hole-doped high-$T_c$ cuprates \cite{Walstedt,IshidaBi2212,FujiwaraJPSJ,MagishiTl1212}.
Upon cooling down to $T_c$, $K_{s,ab}(T)$ is roughly constant for 0212F($\sharp$1), while $K_{s,ab}(T)$ decreases for the other four samples in association with the opening of pseudogaps \cite{Yasuoka,REbook}. It is well known that the pseudogaps in $K_{s,ab}(T)$ emerge in underdoped regions, not in overdoped regions \cite{IshidaBi2212,FujiwaraJPSJ}. A steep decrease below $T_c$ for all samples indicates the reduction in spin susceptibility $\chi_s(T)$ proportional to $K_{s,ab}(T)$ due to the formation of spin-singlet Cooper pairing.
We note here that in hole-doped cuprates $K_{s,ab}(T)$ at room temperature, $K_{s,ab}$(300 K), increases with hole density $p$ \cite{Kotegawa2001,Walstedt,Ohsugi,Ishida,FujiwaraJPSJ,MagishiTl1212,Storey}. The values of $K_{s,ab}$(300 K) in the 0212F samples are listed in Table.~\ref{t:ggg}. The relationship between $p$ and $K_{s,ab}$(300 K) is discussed later.
\begin{figure}[tpb]
\begin{center}
\includegraphics[width=0.8\linewidth]{Fig2.eps}
\end{center}
\caption{\footnotesize (color online) (a) Spin part of the $^{63}$Cu Knight shift $K_{s,ab}(T)$ with $H\parallel ab$ as a function of temperature. The data plots are assigned as labeled in the figure. (b) Spin part of the $^{63}$Cu Knight shift $K_{s,c}(T)$ with $H\parallel c$. $K_{s,c}(T)$ shows the similar $T$-dependence with $K_{s,ab}(T)$.}
\label{fig:n2KS}
\end{figure}
\subsubsection{$^{63}$Cu-NMR with $H\parallel c$}
Figure \ref{fig:n2NMR}(c) shows a typical Cu-NMR spectrum of the central transition (1/2 $\Leftrightarrow$ $-$1/2) for 0212F($\sharp$1), where the field-swept NMR spectra were measured at $H$ parallel to the $c$-axis ($H\parallel c$) and at $\omega_0$=174.2 MHz. The small peak denoted by the asterisk ($\ast$) in the higher-field region arises from unoriented grains with $\theta$ $\sim$ 90$^{\circ}$, where $\theta$ is the angle between $H$ and the $c$-axis.
Figure \ref{fig:n2KS}(b) shows the $T$-dependence of $K_{s,c}(T)$. We estimate $K_{orb,c}$ $\simeq$ 1.22 $\%$, assuming $K_{s,c}$ $\simeq$ 0 at $T$=0 limit. Note that when $H\parallel c$-axis, the second order quadrupole shift, corresponding to the second term in Eq.~(\ref{eq:shift}), is zero \cite{Abragam,TakigawaNQRshift}.
The $K_{s,c}(T)$ in Fig.~\ref{fig:n2KS}(b) shows a similar $T$-dependence with $K_{s,ab}(T)$ in Fig.~\ref{fig:n2KS}(a), as reported for other compounds such as Tl- \cite{MagishiTl1212,Kitaoka}, Bi- \cite{Ishida}, and Hg-based compounds \cite{Julien,MagishiHg1223}. It is, however, different from L214 \cite{Ohsugi}, YBa$_2$Cu$_3$O$_{6+x}$ (Y123) \cite{Walstedt,Barrett,Takigawa} and YBa$_2$Cu$_4$O$_8$ (Y124) \cite{Zimmermann}; $K_{s,c}(T)$ {\it increases} below $T_c$ upon cooling. This inconsistency comes from the difference of hyperfine magnetic fields discussed below.
\subsubsection{Hyperfine magnetic field in CuO$_2$ plane}
According to the Mila-Rice Hamiltonian \cite{MR}, the spin Knight shift of Cu in the CuO$_2$ plane is expressed as
\begin{equation}
K_{s,\alpha}(T) = (A_{\alpha}+4B)\chi_s(T) ~~~(\alpha =c,ab),
\label{eq:Ks}
\end{equation}
where $A_{\alpha}$ and $B$ are the on-site and the supertransferred hyperfine fields of Cu, respectively. Here, the $A_{\alpha}$ consists of the contributions induced by on-site Cu $3d_{x^2-y^2}$ spins $-$ anisotropic dipole, spin-orbit, and isotropic core polarization; and the $B$ originates from the isotropic $4s$ spin polarization produced by neighboring four Cu spins through the Cu($3d_{x^2-y^2}$)-O($2p\sigma$)-Cu($4s$) hybridization.
Since the spin susceptibility $\chi_s(T)$ is assumed to be isotropic, the anisotropy $\Delta$ of $K_{s,\alpha}(T)$ is given by
\begin{equation}
\Delta \equiv \frac{K_{s,c}(T)}{K_{s,ab}(T)}=\frac{A_{c}+4B}{A_{ab}+4B}.
\label{eq:B}
\end{equation}
From Fig.~\ref{fig:n2KS}, $\Delta$ is evaluated as $\sim$ 0.42, 0.38, 0.33, and 0.33 for 0212F($\sharp$1), 0212F($\sharp$2), 0212F($\sharp$3), and 0212F($\sharp$4), respectively. The on-site hyperfine fields, $A_{ab}$ $\sim$ 37 kOe/$\mu_B$ and $A_c$ $\sim$ $-$170 kOe/$\mu_B$ \cite{Monien,Millis,Imai}, are assumed as material-independent in hole-doped high-$T_c$ cuprates, which allows us to estimate $B$ for 0212F samples as listed in Table \ref{t:ggg}.
These $B$ values are larger than $B\sim$ 40 kOe/$\mu_B$, which is a typical value for L214 \cite{Ohsugi}, Y123 \cite{Walstedt,Barrett,Takigawa}, and Y124 \cite{Zimmermann} compounds.
\subsection{Three-layered Ba$_2$Ca$_2$Cu$_3$O$_6$(F,O)$_2$ }
\begin{figure}[htpb]
\begin{center}
\includegraphics[width=1.0\linewidth]{Fig3.eps}
\end{center}
\caption{\footnotesize (color online) (a) Crystal structure of three-layered Ba$_2$Ca$_2$Cu$_3$O$_6$(F,O)$_2$ (0223F). In a unit cell, there are two kinds of CuO$_2$ planes: outer CuO$_2$ plane (OP) and inner CuO$_2$ plane (IP). (b,c) Cu-NMR spectrum of 0223F at $T$=200 K with $H\parallel ab$ and $H\parallel c$. The spectra have two components arising from OP and IP. (d,e) $T$-dependences of $K_{s,ab}(T)$ and $K_{s,c}(T)$ for 0223F.}
\label{fig:n3NMR}
\end{figure}
Figures \ref{fig:n3NMR}(b) and \ref{fig:n3NMR}(c) show the Cu-NMR spectra of the central transition (1/2 $\Leftrightarrow$ $-$1/2) for three-layered Ba$_2$Ca$_2$Cu$_3$O$_6$(F,O)$_2$ (0223F) at $T$=200 K. The field-swept NMR spectra in Figs.~\ref{fig:n3NMR}(a) and \ref{fig:n3NMR}(b) were measured at $H\parallel ab$ and $H\parallel c$, respectively. Here, the NMR frequency $\omega_0$ was fixed at 174.2 MHz.
As shown in the crystal structure of 0223F in Fig.~\ref{fig:n3NMR}(a), multilayered compounds, which have more than three CuO$_2$ planes in a unit cell, are composed of two kinds of CuO$_2$ planes: outer CuO$_2$ plane (OP) and inner CuO$_2$ plane (IP).
The NMR spectra in Figs.~\ref{fig:n3NMR}(b) and \ref{fig:n3NMR}(c) were assigned to OP and IP as denoted in the figure according to the literature \cite{JulienPRL}.
The NMR spectral width for OP is much broader than that for IP; OP is closer to the Ba-F layer (see Fig.~\ref{fig:n3NMR}(a)), which is the source of the disorder due to the atomic substitution at apical-F sites.
Figures \ref{fig:n3NMR}(d) and \ref{fig:n3NMR}(e) show the $T$-dependences of $K_{s,ab}(T)$ and $K_{s,c}(T)$ for 0223F, respectively. Above $T_c$=120 K, both $K_{s,ab}$ and $K_{s,c}$ decrease upon cooling $T$ due to the opening of pseudogaps, which suggests that both OP and IP are underdoped.
As mentioned before in connection with Fig.~\ref{fig:n2KS}(a), $K_{s,ab}$(300 K) increases with $p$.
Therefore, Fig.~\ref{fig:n3NMR}(d) suggests that $p$(OP) is larger than $p$(IP) in 0223F. The estimation of hole density in multilayered cuprates is discussed later. The anisotropy $\Delta$ is evaluated from $K_{s,ab}(T)$ and $K_{s,c}(T)$ (see Eq.~(\ref{eq:B})), as conducted in bi-layered 0212F samples. $\Delta$(OP) and $\Delta$(IP) are evaluated as $\sim$ 0.36 and $\sim$ 0.32, which provide $B$(OP) $\sim$ 72 kOe/$\mu_B$ and $B$(IP) $\sim$ 67 kOe/$\mu_B$.
\section{Discussions}
\subsection{$p$ dependence of $K_{s,ab}$(300 K), $B$, and $\chi_s$(300 K) in plane Cu site}
\begin{figure}[htpb]
\begin{center}
\includegraphics[width=0.85\linewidth]{Fig4.eps}
\end{center}
\caption{\footnotesize (color online) (a) $K_{s,ab}$(300 K) and (b) $B$ are plotted as a function of $p$ for 0212F ($\Box$), Bi2212 ($+$), Tl2201 ($\triangle$), and Tl1212 ($\bigcirc$). The values of $p$ are estimated from $T_c$=$T_{c,max}[1-82.6(p-0.16)^2]$ \cite{Presland}. The plots denoted by crosses ($\times$) in (b) are for multilayered compounds, which are listed in Table \ref{t:ggg2}. (c) $p$-dependence of $\chi_s$(300 K) evaluated from $K_{s,ab}$=($A_{ab}+4B$)$\chi_s$ (Eq.~(\ref{eq:Ks})). The solid lines in the figures are guides to the eye.}
\label{fig:Bterm}
\end{figure}
Figure \ref{fig:Bterm}(a) shows $K_{s,ab}$(300 K) for 0212F, Bi$_2$Sr$_2$CaCu$_2$O$_8$ (Bi2212) \cite{IshidaBi2212}, Tl$_2$Ba$_2$CuO$_{6+\delta}$ (Tl2201) \cite{KitaokaTl2201}, and TlSr$_2$CaCu$_2$O$_{7-\delta}$ (Tl1212) \cite{MagishiTl1212} plotted against the $p$ evaluated by using $T_c$=$T_{c,max}[1-82.6(p-0.16)^2]$ \cite{Presland}.
We note that those compounds have one or two CuO$_2$ planes with tetragonal symmetry, which are homologous series of the apical-F, Tl-, and Hg-based multilayered compounds.
As shown in Fig.~\ref{fig:Bterm}(a), $K_{s,ab}$(300 K) monotonically increases with $p$ from underdoped to overdoped regions. $K_{s,ab}$(300 K) seems material-independent, suggesting that $K_s$(300 K) is a good indication of $p$ in hole-doped cuprates. If the validity of the $K_{s,ab}$(300 K)-$p$ relationship is presented, we can apply it to the estimation of $p$ in multilayered compounds.
According to Eq.~(\ref{eq:Ks}), the $p$ dependence of $K_{s,ab}$(300 K) is derived from those of $B$ and $\chi_s$(300 K). Figure \ref{fig:Bterm}(b) shows the $p$ dependence of $B$ for the same materials shown in Fig.~\ref{fig:Bterm}(a) \cite{IshidaBi2212,KitaokaTl2201,MagishiTl1212}. As presented in Fig.~\ref{fig:Bterm}(b), $B$ increases with $p$, showing a steep increase at $p$ $\sim$ 0.18-0.20.
It is remarkable that the $B$ term exhibits weak $p$-dependence in the underdoped region, while it shows a steep increase with $p$ $>$ 0.16 in the overdoped region.
The $B$ term arises from the Cu($3d_{x^2-y^2}$)-O($2p\sigma$)-Cu($4s$) covalent bonds with the four nearest neighbor Cu sites; therefore, it is expected that the hybridization
between Cu($3d_{x^2-y^2}$) and O($2p\sigma$) orbits becomes larger as $p$ increases in an overdoped regime, and that as a result, $T_c$ starts to decrease.
In Fig.~\ref{fig:Bterm}(b), the $p$-dependent $B$ term seems {\it material-independent}; however, the $B$ terms for L214, Y123, and Y124, $\sim$ 40 kOe/$\mu_B$, are relatively small compared with the values shown in Fig.~\ref{fig:Bterm}(b).
This inconsistency is seen in the variation of $^{63}\nu_Q$ in Fig.~\ref{fig:nuQ} as well. The $^{63}\nu_Q$ increases with $p$ for all materials, while, for a given $p$, the absolute values for L214 and Y123 are about 2 to 3 times larger than those for others. The values of $^{63}\nu_Q$ depend on the hole number $n_d$ in Cu($3d_{x^2-y^2}$) orbit and $n_p$ in O($2p\sigma$) orbit. Therefore, it is expected that $n_d$ and $n_p$ in L214 and Y123 are distributed in a different manner from those in the cuprates presented here, even though
$p$ ($p$=$n_d$+2$n_p-$1) is the same between the former and the latter.
Actually, it has been reported that $n_d$ is large in L214 and Y123, which is the reason for the large $\nu_Q$ \cite{Zheng,Haase}.
In this context, it is considered that the hybridization between Cu($3d_{x^2-y^2}$) and O($2p\sigma$) orbits in L214,Y123, and Y124 is smaller than those in the cuprates presented here, resulting in the $B$ for the former being remarkably smaller than for the latter.
This is probably related to the crystal structures; L214, Y123, and Y124 have orthorhombic crystal structures in superconducting region, while 0212F, Bi2212, Tl2201, and Tl1212 have tetragonal ones. We conclude that the $p$-dependence of $B$ in Fig.~\ref{fig:Bterm}(b) holds in CuO$_2$ planes with tetragonal symmetry.
\begin{figure}[htpb]
\begin{center}
\includegraphics[width=0.9\linewidth]{Fig5.eps}
\end{center}
\caption{\footnotesize (color online) $^{63}$Cu-NQR frequency $^{63}\nu_Q$ for hole-doped cuprates, 0212F, Bi2212 \cite{IshidaBi2212}, Tl2201 \cite{Fujiwara}, Tl1212 \cite{MagishiTl1212}, L214 \cite{Ohsugi}, and Y123 \cite{Pennington,Yoshinari}. Here, $p$ is evaluated from Sr-content $x$ for L214 and from $T_{c,max}$ \cite{Presland} for Y123. The solid lines in the figure are guides to the eye.}
\label{fig:nuQ}
\end{figure}
Figure \ref{fig:Bterm}(c) shows the $p$ dependence of $\chi_s$(300 K), evaluated using $K_{s,ab}$(300 K)=$|A_{ab}+4B|\chi_s$(300 K) (see Eq.~(\ref{eq:Ks})).
The $\chi_s$(300 K) increases with $p$ and is nearly constant above $p$ $\sim$ 0.20-0.25, which is consistent with previous reports \cite{Loram}.
In general, the spin susceptibility $\chi_s$ is related to the density of states at the Fermi level. Therefore, the reduction of $\chi_s$(300 K) in an underdoped regime would be attributed to the emergence of pseudogaps as discussed in previous literature \cite{Loram}.
When we take into account the $p$-dependencies of $B$ and $\chi_s$(300 K), the $p$-dependence of $K_{s,ab}$(300 K) -- a monotonically increasing function -- is explained.
In over-doped regimes, a strong hybridization between Cu($3d_{x^2-y^2}$) and O($2p\sigma$) orbits increases the $B$, making $K_{s,ab}$(300 K) larger as $p$ increases. In under-doped regimes, the opening of the pseudogap decreases $\chi_s$(300 K), making $K_{s,ab}$(300 K) smaller as $p$ decreases.
As a result, we conclude that the $p$-dependence of $K_{s,ab}$(300 K) holds in CuO$_2$ planes with tetragonal symmetry.
This relationship between $K_{s,ab}$(300 K) and $p$ gives us an opportunity to determine $p$ separately for OP and IP in multilayered cuprates, if it is confirmed that the relationship holds in multilayered cuprates.
\subsection{Hole density estimation in multilayered compounds}
\begin{table*}[htbp]
\caption[]{List of $K_{s,ab}$(300 K), $p'$, and $B$ for multilayered compounds: three-layered 0223F, Hg1223 \cite{MagishiHg1223}, and five-layered Hg1245 \cite{Kotegawa2004}. Here, $p'$ is the hole density that is tentatively evaluated by using experimental values of $K_{s,ab}$(300 K) and the solid line in Fig.~\ref{fig:Bterm}(a).}
\label{samples}
\begin{center}
{\tabcolsep = 3.5mm
\renewcommand\arraystretch{1.2}
\begin{tabular}{ccccc}
\hline\hline
Sample, $T_c$ &layer & $K_{s,ab}$(300 K) ($\%$) & $p'$ & $B$(kOe/$\mu_B$) \\
\hline
0223F, 120 K & OP & 0.35 & 0.156 & 72 \\
& IP & 0.28 & 0.115 & 67 \\
Hg1223, 133 K & OP & 0.41 & 0.185 & 90 \\
& IP & 0.32 & 0.139 & 80 \\
Hg1245, 108 K & OP & 0.34 & 0.151 & 74 \\
& IP & 0.21 & 0.075 & 61 \\
\hline\hline
\end{tabular}}
\end{center}
\label{t:ggg2}
\end{table*}
In multilayered cuprates, doped hole carriers reside on OP and IP with different doping levels, and the CuO$_2$ layers show different physical properties due to the charge distribution even in the same sample. Therefore, it is required to separately estimate $p$ for OP and IP in order to study the electronic states of multilayered cuprates.
It is invalid to apply the estimation methods used in single- and bi-layered compounds, for example, the relationship $T_c$=$T_{c,max}$[1-82.6($p$-0.16)$^2$] \cite{Presland}, the thermoelectric power \cite{Obertelli,Tallon}, and the bond valence sums \cite{Brown,Cava,TallonBV}. Those methods are applicable to evaluate total hole density, but not to evaluate each hole density at CuO$_2$ planes.
On the other hand, Cu-NMR measures the respective values of $K_{s,ab}$(300 K) for OP and IP. If the $K_{s,ab}$(300 K)-$p$ relationship in Fig.~\ref{fig:Bterm}(a) holds even in multilayered cuprates, it allows us to separately estimate $p$ for OP and IP from experimental values of $K_{s,ab}$(300 K).
Table \ref{t:ggg2} lists $K_{s,ab}$(300 K), $p'$, and $B$ for multilayered cuprates: three-layered 0223F, HgBa$_2$Ca$_2$Cu$_3$O$_{8+\delta}$ (Hg1223) \cite{MagishiHg1223}, and five-layered HgBa$_2$Ca$_2$Cu$_3$O$_{12+\delta}$ (Hg1245) \cite{Kotegawa2004}. Here, $p'$ is the hole density tentatively estimated by using the $K_{s,ab}$(300 K)-$p$ relationship in Fig.~\ref{fig:Bterm}(a). We plot $B$ of those compounds as crosses ($\times$) in Fig.~\ref{fig:Bterm}(b). The data plots fit into the other data, suggesting that the $K_{s,ab}$(300 K)-$p$ relationship is also valid in multilayered cuprates.
We may consider the case that the $K_{s,ab}$(300 K)-$p$ relationship is modified in multilayered cuprates. According to Eq.~(\ref{eq:Ks}), the possible source of the modification is the $p$ dependence of $B$ or of $\chi_s$ or both.
Experimental $B$ values in multi-layered compounds, however, fall on the same universal curve as shown in Fig.~\ref{fig:Bterm}(b); some unexpected modifications on the side of $\chi_s$ can indeed be considered as unlikely.
Therefore, we conclude that the $K_{s,ab}$(300 K)-$p$ relationship in Fig.~\ref{fig:Bterm}(a) is valid for CuO$_2$ planes in multilayered compounds, regardless of OP and IP.
Moreover, when we apply the $K_{s,ab}$(300 K)-$p$ relationship to five-layered compounds,
the fact that the maximum $T_c$ takes place at $p$ $\sim$ 0.16 \cite{Mukuda2008} supports the validity of the application to multilayered compounds.
Several attempts have been used to determine $p$ of hole-doped cuprates so far -- the relationship between $p$ and $K_{s,ab}$(300 K) in this work is a promising approach to estimate $p$, which is {\it applicable to multilayered compounds}. To our knowledge, it is the only method to separately determine $p$ for OP and IP in multilayered compounds with a reliable accuracy.
\section{Conclusion}
We have shown that the planar CuO$_2$ hole densities $p$ in high-$T_c$ cuprates are consistently determined with the Cu-NMR Knight shift. It has been demonstrated that the spin part of the Knight shift $K_{s,ab}$(300 K) at room temperature is material-independent in CuO$_2$ planes with tetragonal symmetry, and that $K_{s,ab}$(300 K) monotonically increases with $p$ from underdoped to overdoped regions. These observations suggests that $K_{s,ab}$(300 K) is a reliable method for determining planar CuO$_2$ hole densities.
The experimental values of $K_{s,ab}$(300 K) and of hyperfine magnetic fields for three-layered Ba$_2$Ca$_2$Cu$_3$O$_6$(F,O)$_2$ and other multilayered compounds support the application of the $p$-$K_s$(300 K) relationship to multilayered compounds. We remark that, to our knowledge, the relationship is the only method to separately determine $p$ for OP and IP in multilayered compounds with a reliable accuracy.
\section*{Acknowledgement}
The authors are grateful to M. Mori for his helpful discussions. This work was supported by Grant-in-Aid for Specially promoted Research (20001004) and by the Global COE Program (Core Research and Engineering of Advanced Materials-Interdisciplinary Education Center for Materials Science) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan.
|
\section{Introduction}
Understanding anomalous transport in magnetically confined plasmas is an outstanding issue in controlled fusion research. A satisfactorily understanding of the non-local features as well as the non-Gaussian probability distribution functions (PDFs) found in experimental measurements of particle and heat fluxes is still lacking. In particular, experimental observations of the edge turbulence in the fusion devices \cite{Zweben} show that in the Scrape of Layer (SOL) the plasma fluctuations are characterized by non-Gaussian PDFs. It has been recognized that the nature of the cross-field transport through the SOL is dominated by turbulence with a significant ballistic or non-local component where a diffusive description is improper \cite{Naulin}. Moreover, the scaling of the confinement time $\tau \propto L^{\alpha}$ with $\alpha < 2$ \cite{kaye} is typical in low-confinement mode discharges, instead of the diffusion induced result $\tau \propto L^2$, where $L$ is the system size. There is a considerable amount of experimental evidence \cite{cardozo1995, gentle1995, callen1997, mantica1999, vanmilligen2002, BalescuBook} and recent numerical gyrokinetic \cite{pradalier, sanchez2008} and fluid \cite{negrete2005} simulations that plasma turbulence in tokamaks is highly non-local.
In addition, intermittent turbulence is characterized by patchy spatial structure that is bursty in time. The PDFs of these intermittent events shows unimodal structure with "elevated" tails that deviates from a Gaussian prediction. The understanding of these events are at best limited \cite{Zweben, bramwell2009, carreras1996, carreras1999, anderson1, anderson2}. Moreover, the high possibility of confinement degradation by intermittency strongly calls for a predictive theory.
A prominent candidate for explaining the suggestive non-local features of plasma turbulence is the inclusion of a fractional velocity derivative in the Fokker-Planck (FP) equation leading to an inherently non-local description as well as giving rise to non-Gaussian PDFs of e.g. densities and heat flux. The non-locality is introduced through the integral description of the fractional derivative \cite{zaslavsky, sanchez, negrete} and the non-Maxwellian distribution function drives the observed PDFs of densities and heat flux far from Gaussian.
The aim of this study is to elucidate the effects of a non-Maxwellian distribution function induced by the fractional velocity derivative in the Fokker-Planck equation. Some previous papers on plasma transport have used models including a fractional derivative where the fractional derivative is introduced on phenomenological premises \cite{sanchez, negrete}. In the present work we introduce the Levy statistics into the Langevin equation thus yielding a fractional FP description. This approach is similar to that of Ref. \cite{sanchez2006} resulting in a phenomenological description of the non-local effects in plasma turbulence. Using fractional generalizations of the Liouville equation, kinetic descriptions have been developed previously \cite{zaslovsky,tarasov}. It has been shown that the chaotic dynamics can be described by using the FP equation with coordinate fractional derivatives as a possible tool for the description of anomalous diffusion \cite{zaslovsky2}. Much work has been devoted on investigation of the Langevin equation with Levy white noise, see References \cite{West, Jeperson, Fogedby, Vlad}, or related fractional FP equation \cite{West}. Furthermore, fractional derivatives have been introduced into the FP framework in a similar manner as the present work \cite{chechkin2000, chechkin2002} but a study including drift waves is still called for. To this end we quantify the effects of the fractional derivative in the FP equation in terms of a modified dispersion relation for density gradient driven linear plasma drift waves where we have considered a case with constant external magnetic field and a shear-less slab geometry. In order to calculate an equilibrium PDF we use a model based on the motion of a charged Levy particle in a constant external magnetic field obeying non-Gaussian, Levy statistics. This assumption is the natural generalization of the classical example of the motion of a charged Brownian particle with the usual Gaussian statistics \cite{chandrasekhar}. The fractional derivative is represented with the Fourier transform containing a fractional exponent. We find a relation for the deviation from Maxwellian distribution described by $\epsilon$ through the quasi-neutrality condition and the characteristics of the plasma drift wave are fundamentally changed, i.e. the values of the growth-rate $\gamma$ and real frequency $\omega$ are significantly altered. A deviation from the Maxwellian distribution function alters the dispersion relation for the density gradient drift waves such that the growth rates are substantially increased and thereby may cause enhanced levels of transport.
The paper is organized as follows. In Sec. 2 the mathematical framework of the fractional FP equation (FFPE) is introduced. In Sec. 3 a dispersion relation for the density gradient driven drift modes using the FFPE are derived. In Sec 4 the deviations from a Maxwellian distribution function are investigated and the dispersion relation is solved in Sec. 5. We conclude the paper with a results and discussion in Sec. 6.
\section{Fractional Fokker-Planck Equation}
Following the theory of Brownian motion we write an equation of motion for a colloidal particle in a background medium as a Langevin equation of the following form \cite{chandrasekhar}
\begin{eqnarray}\label{eq:1.1}
\frac{d\mathbf{v}}{dt}=-\nu\mathbf{v}+A(t)
\end{eqnarray}
Here, we assumed that the influence of the background medium can be split into a dynamical friction, $-\nu\mathbf{v}$, and a fluctuating part, $A(t)$ which is a Gaussian white noise. The Gaussian white noise assumption is usually imposed in order to obtain a Maxwellian velocity distribution describing the equilibrium of the Brownian particle. This connection is due to the relation between the Gaussian central limit theorem and classical Boltzmann-Gibbs statistics \cite{Khintchine}. However, the Gaussian central limit theorem is not unique and a generalization of the Gaussian central limit theorem to the case of summation of independent identically distributed random variables described by long tailed distributions is performed by L\'{e}vy \cite{levy}, and Khintchine \cite{Khintchine}. In this case the L\'{e}vy distributions replace the Gaussian in a generalized central limit theorem.
The simplest case of generalized Brownian motion considered by West and Seshadri \cite{seshadri} is to assume for fluctuation part, $A(t)$, in Equation (\ref{eq:1.1}) to be a white L\'{e}vy noise. Following the approach used by Barkai \cite{barkai} we find the Fractional Fokker-Planck Equation (FFPE) with fractional velocity derivatives for shear-less slab geometry in the presence of a constant external force as
\begin{eqnarray}\label{eq:1.2}
\frac{\partial F_{s}}{\partial t}+\mathbf{v}\frac{\partial F_{s}}{\partial \mathbf{r}}+\frac{\mathbf{F}}{m_{s}}\frac{\partial F_{s}}{\partial \mathbf{v}}=\nu\frac{\partial }{\partial \mathbf{v}}(\mathbf{v}F_{s})+D\frac{\partial^{\alpha} F_{s}}{\partial |\mathbf{v}|^{\alpha}},
\end{eqnarray}
where $s(=e,i)$ represents the particle species and $0\le\alpha\le 2$. Here, the term $\frac{\partial^{\alpha} F_{s}}{\partial |\mathbf{v}|^{\alpha}}$ is the fractional Riesz derivative. The fractional differentiation may be represented through singular integrals or by its Fourier transform as we will see later in Equation (\ref{eq:1.4}). Note that the connection to the integral representation indicates that the model is inherently non-local in velocity space. The diffusion coefficient, $D$, is related to the damping term $\nu$, according to a generalized Einstein relation \cite{barkai}
\begin{eqnarray}\label{eq:1.3}
D=\frac{2^{\alpha-1}T_{\alpha}\nu}{\Gamma(1+\alpha)m_{s}^{\alpha-1}}.
\end{eqnarray}
Here, $T_{\alpha}$ is a generalized temperature, and taking force $\mathbf{F}$ to represent the Lorentz force (due to a constant magnetic field and a zero-averaged electric field) acting on the particles of species $s$ with mass $m_{s}$ and $\Gamma(1+\alpha)$ is the Euler gamma function. We find the solution by using the Fourier representation of equation (\ref{eq:1.2}) above as
\begin{eqnarray}\label{eq:1.4}
\frac{\partial \mathcal{F}_{s}}{\partial t}+(-\mathbf{k}+\Omega_{s}(\mathbf{k}^{v}\times \hat{b})+\nu\mathbf{k}^{v})\frac{\partial\mathcal{F}_{s}}{\partial \mathbf{k}^{v}}=-D|\mathbf{k}^{v}|^{\alpha} \mathcal{F}_{s},
\end{eqnarray}
where $\Omega_{s}=e_{s}B/m_{s}c$ is the Larmor frequency of species $s$, $\hat{b}=\mathbf{B}/B$ is the unit vector in the direction of magnetic field and $\mathcal{F}_{s}$ is the characteristic function
\begin{eqnarray}\label{eq:1.5}
\mathcal{F}_{s}(\mathbf{k},\mathbf{k}^{v};t)=\int\int d\mathbf{r}\;d\mathbf{v}\exp(i\mathbf{k}\cdot\mathbf{r}+i\mathbf{k}^{v}\cdot\mathbf{v})F_{s}(\mathbf{r},\mathbf{v};t),
\end{eqnarray}
where we have denoted the wave-vector by $\mathbf{k}$ and the corresponding wave vector for the velocity as $\mathbf{k}^v$. We can rewrite the kinetic equation by identification of time derivatives of the wave vectors as
\begin{eqnarray}\label{eq:1.6}
\frac{d\mathcal{F}_{s}}{dt}=\frac{\partial \mathcal{F}_{s}}{\partial t}+\frac{d\mathbf{k}^{v}}{dt}\frac{\partial\mathcal{F}_{s}}{\partial \mathbf{k}^{v}}+\frac{d\mathbf{k}}{dt}\frac{\partial\mathcal{F}_{s}}{\partial \mathbf{k}}=0.
\end{eqnarray}
We use the method of characteristics on the Equation (\ref{eq:1.4}) and (\ref{eq:1.6}) whereby we find that the characteristics are
\begin{eqnarray}\label{eq:1.7}
\frac{\partial \mathcal{F}_{s}}{\partial t}=-D|\mathbf{k}^{v}|^{\alpha} \mathcal{F}_{s},\\
\frac{d\mathbf{k}^{v}}{dt}=-\mathbf{k}+\Omega_{s}(\mathbf{k}^{v}\times \hat{b})+\nu\mathbf{k}^{v},\\
\frac{d\mathbf{k}}{dt}=0.
\end{eqnarray}
Following the method used in Ref. \cite{chechkin2002,chechkin2000} the solution corresponding to the homogenous and steady state system in Fourier space is
\begin{eqnarray}\label{eq:2.14}
\mathcal{F}_{s}(\mathbf{k}^{v}, t)=e^{-\frac{D}{\alpha\nu}(|\mathbf{k}^{v}_{\bot}|^{\alpha}+|\mathbf{k}^{v}_{\parallel}|^{\alpha})}.
\end{eqnarray}
In order to find the solution in real space we compute the inverse Fourier transform of Equation (\ref{eq:2.14})
\begin{eqnarray}\label{eq:2.15}
F_{s}(\mathbf{r},\mathbf{v})=C(\mathbf{r})\int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}e^{-\frac{D}{\alpha\nu}(|\mathbf{k}^{v}_{\bot}|^{\alpha}+|\mathbf{k}^{v}_{\parallel}|^{\alpha})}.
\end{eqnarray}
We define a new variable $\mathcal{D}=\frac{D}{\nu}$ where coefficient $D$ is given by the expression in Equation (\ref{eq:1.3}). $C(\mathbf{r})$ is a normalization factor which remains to be defined. Taking the inverse Fourier transform of the Equation (\ref{eq:2.15}) for $\alpha=2$ we get
\begin{eqnarray}\label{eq:2.16}
F_{s}(\mathbf{r},\mathbf{v})=\frac{C(\mathbf{r})}{\mathcal{D}}e^{-(\frac{v_{\bot}^2+v_{\parallel}^2}{4\mathcal{D}})}.
\end{eqnarray}
The unknown normalization factor $C$ can be determined by comparing the integrals of the Maxwellian distribution and our distribution. In comparison the Maxwellian distribution is defined as
\begin{eqnarray}\label{eq:2.17}
F^{M}_{s}(\mathbf{r},\mathbf{v})=\frac{n_{s}(\mathbf{r})}{(\sqrt{\pi}V_{T,s}(\mathbf{r}))^3}e^{-(v_{\bot}^2+v_{\parallel}^2)/V_{T,s}^2(\mathbf{r})},
\end{eqnarray}
where $V_{T,s}(\mathbf{r})=\sqrt{2T_{s}(\mathbf{r})/m_{s}}$ is the thermal velocity of species $s$. By integrating the Maxwellian distribution over the velocity space we find the density as
\begin{eqnarray}\label{eq:2.18}
\int d\mathbf{v}F^{M}_{s}(\mathbf{r},\mathbf{v})=2\pi\int_{0}^{\infty}v_{\bot}dv_{\bot}\int_{-\infty}^{\infty}dv_{\parallel}\frac{n_{s}(\mathbf{r})}{(\sqrt{\pi}V_{T,s}(\mathbf{r}))^3}e^{-(v_{\bot}^2+v_{\parallel}^2)/V_{T,s}^2(\mathbf{r})}=n_{s}(\mathbf{r}),
\end{eqnarray}
whereas performing the same integration of the expression in Equation(\ref{eq:2.16}) we obtain
\begin{eqnarray}\label{eq:2.19}
\int d\mathbf{v}F_{s}(\mathbf{r},\mathbf{v})=2\pi\int_{0}^{\infty}v_{\bot}dv_{\bot}\int_{-\infty}^{\infty}dv_{\parallel}\frac{C(\mathbf{r})}{ \mathcal{D}}e^{-(\frac{v_{\bot}^2+v_{\parallel}^2}{4\mathcal{D}})}=2\pi^{3/2}\sqrt{2\mathcal{D}} C(\mathbf{r}).
\end{eqnarray}
We can now compare the two results obtained in Equations (\ref{eq:2.18}) - (\ref{eq:2.19}) and we find the following relation
\begin{eqnarray}\label{eq:1.21}
C(\mathbf{r})=\frac{n_{s}(\mathbf{r})}{2\pi^{3/2}\sqrt{2\mathcal{D}}}.
\end{eqnarray}
The distribution function can now be determined by replacing this expression into Equation (\ref{eq:2.16}) for $C(\mathbf{r})$ yielding
\begin{eqnarray}\label{eq:1.22}
F_{s}(\mathbf{r},\mathbf{v})=\frac{n_{s}(\mathbf{r})}{2\pi^{3/2}\mathcal{D}\sqrt{2\mathcal{D}}} e^{-(\frac{v_{\bot}^2+v_{\parallel}^2}{4\mathcal{D}})}.
\end{eqnarray}
We can easily recover the Maxwellian distribution in Equation (\ref{eq:2.17}) by setting $\alpha=2$ in the definition for $D$ in Equation (\ref{eq:1.3}) and using that $\Gamma(3)=2$. Note that for a general $\alpha$, the equilibrium distribution is as follows
\begin{eqnarray}\label{eq:2.24.1}
F_{s}(\mathbf{r},\mathbf{v})=\frac{n_{s}(\mathbf{r})}{2\pi^{3/2}\sqrt{2\mathcal{D}}} \int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}e^{-\frac{\mathcal{D}}{\alpha}(|\mathbf{k}^{v}_{\bot}|^{\alpha}+|\mathbf{k}^{v}_{\parallel}|^{\alpha})},
\end{eqnarray}
where
\begin{eqnarray}\label{eq:2.25.1}
\mathcal{D}=\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)},
\end{eqnarray}
and we have introduced a generalized thermal velocity as
\begin{eqnarray}\label{eq:2.25.2}
V_{T,s}^{\alpha}=\frac{2^{\alpha-1}T_{\alpha}}{m_{s}^{\alpha-1}}.
\end{eqnarray}
The generalized equilibrium distribution including the effects of the fractional velocity derivative in Equation (\ref{eq:2.24.1}) becomes
\begin{eqnarray}\label{eq:2.24}
F_{s}(\mathbf{r},\mathbf{v})=\frac{n_{s}(\mathbf{r})}{2\pi^{3/2}(\Gamma(1+\alpha))^{-1/2}\sqrt{2V_{T,s}^{\alpha}}}
\int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}e^{-\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)\alpha}(|\mathbf{k}^{v}_{\bot}|^{\alpha}+|\mathbf{k}^{v}_{\parallel}|^{\alpha})}.
\end{eqnarray}
We will now determine the dispersion relation for density gradient driven drift waves including the effects of the fractional velocity differential operator.
\section{The dispersion relation}
In order to quantify the non-local effects on drift waves induced by the fractional differential operator we will determine the dispersion relation for density gradient driven drift modes. We start by formulating the linearized gyro-kinetic theory where the particle distribution function, averaged over gyro-phase is of the form (see Ref. \cite{Balescu1991})
\begin{eqnarray}\label{eq:2.26}
f_{s}(\mathbf{r},\mathbf{v})=F_{s}(\mathbf{r},\mathbf{v})+(2\pi)^{-4}\times\int\int d\mathbf{k}\;d\omega\exp(i\mathbf{k}\cdot\mathbf{r}-i\omega t)\delta f^{s}_{\mathbf{k},\omega}(\mathbf{v}).
\end{eqnarray}
We assume that the turbulence is purely electrostatic and neglect magnetic field fluctuations $(\delta \mathbf{B}=0)$. For small deviations from the local equilibrium we find the linearized gyro-kinetic equation of the form
\begin{eqnarray}\label{eq:2.27}
(-\omega+k_{\parallel}v_{\parallel})\delta f^{s}_{\mathbf{k},\omega}(v_{\parallel},v_{\bot})+(\omega-\omega_{*s})\frac{e_{s}}{T_{s}}J_{0}(|\Omega_{s}|^{-1}k_{\bot}v_{\bot})F_{s}(x,\mathbf{v})\delta \phi_{\mathbf{k},\omega}=0,
\end{eqnarray}
where $\omega_{*s}=\frac{cT_{s}}{e_{s}B}k_{y}\cdot\frac{d\; ln\;n(x)}{d x}$ is the drift wave frequency of species $s$, and we assumed that the space dependence of $F_{s}$ is only in the $x$ direction perpendicular to the magnetic field as well as for the density gradient. In the equation above, $J_{0}$ is the Bessel function of order zero, $v_{\parallel}$ is the parallel velocity, $v_{\bot}\equiv (v_{x}^{2}+v_{y}^{2})^{1/2}$ is the perpendicular velocity and hence we write the total speed as $v=(v_{\bot}^{2}+v_{\parallel}^{2})^{1/2}$. Inserting the expression for $F_{s}$ from the Equation (\ref{eq:2.24}) in Equation (\ref{eq:2.27}) and rearranging the terms we find the perturbed distribution $\delta f_{\mathbf{k},\omega}$ as
\begin{eqnarray}\label{eq:2.28}
\delta f^{s}_{\mathbf{k},\omega}(v_{\parallel},v_{\bot})=
-\frac{e_{s}}{T_{s}}[\frac{\omega-\omega_{*s}}{k_{\parallel}v_{\parallel}-\omega}]J_{0}(|\Omega_{s}|^{-1}k_{\bot}v_{\bot})\delta \phi_{\mathbf{k},\omega}
\frac{n_{s}(\mathbf{r})}{2\pi^{3/2}(\Gamma(1+\alpha))^{-1/2}\sqrt{2V_{T,s}^{\alpha}}} \times\nonumber\\
\int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}e^{-\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)\alpha}(|\mathbf{k}^{v}_{\bot}|^{\alpha}+|\mathbf{k}^{v}_{\parallel}|^{\alpha})}.
\end{eqnarray}
Here, the wave vector perpendicular to magnetic field is $k_{\bot}=(k^2_{x}+k^2_{y})^{1/2}$. The gyro-kinetic Equation (\ref{eq:2.28}) is complemented with Poisson equation for the electric potential. For fluctuations with wave vectors much smaller than the Debye wave vector, the Poisson equation becomes the quasi-neutrality condition
\begin{eqnarray}\label{eq:2.29}
\sum_{s} e_{s}\delta n^{s}_{\mathbf{k},\omega}=0,
\end{eqnarray}
where the density fluctuation is related to the distribution function through
\begin{eqnarray}\label{eq:2.31}
\delta n^{s}_{\mathbf{k},\omega}=-\frac{e_{s}}{T_{s}}n_{s}\delta\phi_{\mathbf{k},\omega} + \int d\mathbf{v}
J_{0}(|\Omega_{s}|^{-1}k_{\bot}v_{\bot})\delta f^{s}_{\mathbf{k},\omega}(v_{\parallel},v_{\bot}).
\end{eqnarray}
In the above equation we have separated the adiabatic response (first term on the right hand side) from the non-adiabatic response (second term on the right hand side). We have to keep in mind that the density $n_{s}$ coming from the $F_{s}(x,\mathbf{v})$ in the adiabatic response is also given by Equation (\ref{eq:2.24}) and for a general $0\le\alpha\le2$ the adiabatic response can be different than that calculated by Maxwellian distribution of Equation (\ref{eq:2.17}). Using the quasi-neutrality condition (\ref{eq:2.29}) we find the dispersion equation which determines the eigenfrequencies as a function of the wave vector, $\omega=\omega(\mathbf{k})=\omega_{r}(\mathbf{k})+i\gamma(\mathbf{k})$. In the simplest case we consider a plasma consisting of electrons and a single species of singly charged ions with the equal temperatures. For the density fluctuation therefore we have
\begin{eqnarray}\label{eq:2.32}
\delta n^{s}_{\mathbf{k},\omega}=-n_{s}(\mathbf{r})\frac{e_{s}}{T_{s}}\delta\phi_{\mathbf{k},\omega}[M^{ad,s}+M^{s}_{\mathbf{k},\omega}].
\end{eqnarray}
Therefore, the dispersion equation as in the Ref. \cite{Balescu1991} is
\begin{eqnarray}\label{eq:2.33}
M^{ad,e}+M^{e}_{\mathbf{k},\omega}=-M^{ad,i}-M^{i}_{\mathbf{k},\omega},
\end{eqnarray}
where
\begin{eqnarray}\label{eq:2.34}
M^{ad,s}=\int d\mathbf{v}
\frac{1}{2\pi^{3/2}(\Gamma(1+\alpha))^{-1/2}\sqrt{2V_{T,s}^{\alpha}}}
\int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}e^{-\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)\alpha}(|\mathbf{k}^{v}_{\bot}|^{\alpha}+|\mathbf{k}^{v}_{\parallel}|^{\alpha})},
\end{eqnarray}
gives the adiabatic contribution, and
\begin{eqnarray}\label{eq:2.35}
M^{s}_{\mathbf{k},\omega}=\int d\mathbf{v}[\frac{\omega-\omega_{*s}}{k_{\parallel}v_{\parallel}-\omega}]J_{0}(b_{s}v_{\bot}/V_{Ts})\times\nonumber\\
\frac{1}{2\pi^{3/2}(\Gamma(1+\alpha))^{-1/2}\sqrt{2V_{T,s}^{\alpha}}}
\int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}e^{-\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)\alpha}(|\mathbf{k}^{v}_{\bot}|^{\alpha}+|\mathbf{k}^{v}_{\parallel}|^{\alpha})},
\end{eqnarray}
gives the non-adiabatic contribution. Here, $b_{s}=k_{\bot}V_{T,s}/\Omega_{s}$. If we take $\alpha=2$ in the Equation(\ref{eq:2.33}) we recover the dispersion equation for a Maxwellian distribution as in the Ref. \cite{Balescu1991}.
\subsection{Adiabatic response}
First, we may analyze the contribution from the adiabatic parts of the dispersion relation only by ignoring all fluctuations, yielding \begin{eqnarray}\label{eq:2.37}
|M^{ad,e}|=|M^{ad,i}|.
\end{eqnarray}
In addition, utilizing the quasi-neutrality condition while neglecting the density gradient in the system we have $n_{i}=n_{e}$, therefore $\alpha_{e}$ and $\alpha_{i}$ becomes connected through Equation (\ref{eq:2.37}). This indicates that the deviation from a Maxwellian distribution described by $\alpha$ for electrons and ions becomes dependent on each other. We will get back to this relation in later sections.
\section{Deviations from a Maxwellian distribution function}
We will now turn our attention to the problem of solving the dispersion relation described by Equation (\ref{eq:2.33}). In order to solve this dispersion equation we use the method proposed in Ref. \cite{Balescu1991} with the difference that here we have to perform additional integrations over $\mathbf{k}^{v}$. We have
\begin{eqnarray}\label{eq:3.1}
M^{s}_{\mathbf{k},\omega}=\frac{\omega-\omega_{*,s}}{|k_{\parallel}|V_{T,s}}Z(\xi_{s})\Gamma(b_{s}),
\end{eqnarray}
where the plasma dispersion function is
\begin{eqnarray}\label{eq:3.2}
Z(\xi_{s})=\frac{V_{T,s}}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\int_{-\infty}^{\infty}du[\frac{\Phi(v_{\parallel})}{u-\xi_{s}-i\sigma}],
\end{eqnarray}
with $u=v_{\parallel}/V_{Ts}$, $\xi_{s}=\omega/(|k_{\parallel}|V_{Ts})$ and the function $\Phi(v_{\parallel})$ is
\begin{eqnarray}\label{eq:3.3}
\Phi(v_{\parallel})=\frac{1}{\sqrt{2(\Gamma(1+\alpha))^{-1/2}\sqrt{2V_{T,s}^{\alpha}}}}
\int \frac{d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{1/2}}e^{-i\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel}}e^{-\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)\alpha}(|\mathbf{k}^{v}_{\parallel}|^{\alpha})}.
\end{eqnarray}
The integral over $v_{\bot}$ can be written in a general way as
\begin{eqnarray}\label{eq:3.4}
\Gamma(b_{s})=2V_{T,s}^2\int_{0}^{\infty}dw w \Psi_{s}(b_{s}w)\Phi(v_{\bot}),
\end{eqnarray}
where $w=v_{\bot}/V_{Ts}$, $\Psi_{s}=J_{0}^2(b_{s}v_{\bot}/V_{Ts})$ and,
\begin{eqnarray}\label{eq:3.5}
\Phi(v_{\bot})=\frac{1}{\sqrt{2(\Gamma(1+\alpha))^{-1/2}\sqrt{2V_{T,s}^{\alpha}}}}
\int \frac{d\mathbf{k}_{\bot}^{v}}{(2\pi)}e^{-i\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}}e^{-\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)\alpha}(|\mathbf{k}^{v}_{\bot}|^{\alpha})}.
\end{eqnarray}
The analytical solutions for integrals over $\mathbf{k}^{v}$ with an arbitrary $\alpha$ in the Equations (\ref{eq:3.3}) and (\ref{eq:3.5}) requires rather tedious calculations. Instead we consider an infinitesimal deviation of the form $\alpha=2-\epsilon$, where $0\le\epsilon\ll 2$ and expand the terms depending on $\alpha$ in the Equations (\ref{eq:3.3}) and (\ref{eq:3.5}) around $\epsilon=0$ as follows
\begin{eqnarray}\label{eq:3.6}
\frac{1}{\sqrt{(\Gamma(1+\alpha))^{-1/2}\sqrt{V_{T,s}^{\alpha}}}} e^{-\frac{V_{T,s}^{\alpha}}{\Gamma(1+\alpha)\alpha}(|k^{v}|^{\alpha})}=
\frac{2^{1/4}e^{-\frac{1}{4}V_{T,s}^{2}|k^{v}|^{2}}}{\sqrt{V_{T,s}}}+\epsilon\Lambda(k^{v})+\mathcal{O}[\epsilon^2],
\end{eqnarray}
where
\begin{eqnarray}\label{eq:3.7}
\Lambda(k^{v})=\frac{e^{-\frac{1}{4}V_{T,s}^{2}|k^{v}|^{2}}}{2^{11/4}\sqrt{V_{T,s}}}\{
-3+2{\gamma_E}-4V_{T,s}^2|k^{v}|^2+2 {\gamma_E} V_{T,s}^2|k^{v}|^2\nonumber\\
+2{\log}[V_{T,s}]+2V_{T,s}^2 {\log}[V_{T,s}]|k^{v}|^2+2V_{T,s}^2|k^{v}|^2 {\log}[|k^{v}|]\}.
\end{eqnarray}
Here, we have used the Euler-Mascheroni constant $\gamma_E = 0.57721$. The first term in Equation (\ref{eq:3.6}) will produce
\begin{eqnarray}\label{eq:3.3.1}
\Phi(u)=\frac{e^{-u^2}}{V_{T,s}^{3/2}},\;\;\;\;\;\;\; \mbox{and} \;\;\;\;\;\;\;\;\Phi(w)=\frac{e^{-w^2}}{V_{T,s}^{3/2}}
\end{eqnarray}
which give the Maxwellian adiabatic response
\begin{eqnarray}\label{eq:3.1.1}
M^{ad,s}=1.
\end{eqnarray}
By using the expansion defined by the expression (\ref{eq:3.6}) in Equations (\ref{eq:2.34}) and (\ref{eq:2.34}), the adiabatic and non-adiabatic part of the dispersion relation $M^{ad,s}$ and $M^{s}_{\mathbf{k},\omega}$ are as follows
\begin{eqnarray}\label{eq:3.8}
M^{ad,s}=1+(2\pi\int_{-\infty}^{\infty}dv_{\parallel}\int_{0}^{\infty}dv_{\bot}v_{\bot}\times\nonumber\\
\frac{1}{2\sqrt{2}\pi^{3/2}} \int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}\Lambda(k_{\bot}^{v})\Lambda(k_{\parallel}^{v}))\epsilon+\mathcal{O}[\epsilon]^2=1+\epsilon W^{ad,s}.\nonumber\\
\end{eqnarray}
and
\begin{eqnarray}\label{eq:3.9}
M^{s}_{\mathbf{k},\omega}=2\pi\int_{-\infty}^{\infty}dv_{\parallel}\int_{0}^{\infty}dv_{\bot}v_{\bot}[\frac{\omega-\omega_{*s}}{k_{\parallel}v_{\parallel}-\omega}]\Psi_{s}(b_{s}v_{\bot}/V_{Ts})\times\nonumber\\
\frac{1}{(\sqrt{\pi}V_{T,s}(\mathbf{r}))^3}e^{-(v_{\bot}^2+v_{\parallel}^2)/V_{T,s}^2(\mathbf{r})}+\nonumber\\
(2\pi\int_{-\infty}^{\infty}dv_{\parallel}\int_{0}^{\infty}dv_{\bot}v_{\bot}[\frac{\omega-\omega_{*s}}{k_{\parallel}v_{\parallel}-\omega}]\Psi_{s}(b_{s}v_{\bot}/V_{Ts})\times\nonumber\\
\frac{1}{2\sqrt{2}\pi^{3/2}} \int \frac{d\mathbf{k}_{\bot}^{v}d\mathbf{k}_{\parallel}^{v}}{(2\pi)^{3/2}}e^{-i(\mathbf{k}_{\bot}^{v}\mathbf{v}_{\bot}+\mathbf{k}_{\parallel}^{v}\mathbf{v}_{\parallel})}\Lambda(k_{\bot}^{v})\Lambda(k_{\parallel}^{v}))\epsilon+\mathcal{O}[\epsilon]^2=N^{s}_{\mathbf{k},\omega}+\epsilon W^{s}_{\mathbf{k},\omega}.\nonumber\\
\end{eqnarray}
Inserting these relations we may rewrite the dispersion relation (\ref{eq:2.33}) in the form
\begin{eqnarray}\label{eq:3.10}
(1+N^{e}_{\mathbf{k},\omega})+\epsilon (W^{ad,e}+W^{e}_{\mathbf{k},\omega})=-(1+N^{i}_{\mathbf{k},\omega})-\epsilon (W^{ad,i}+W^{i}_{\mathbf{k},\omega}).
\end{eqnarray}
The first terms on the right and left hand sides generate the usual contributions to the dispersion equation as in Ref. \cite{Balescu1991} and the terms proportional to $\epsilon$ generate the non-Maxwellian contributions where we have
\begin{eqnarray}\label{eq:3.11}
N^{s}_{\mathbf{k},\omega}=\frac{\omega-\omega_{*,s}}{|k_{\parallel}|V_{T,s}}Z(\xi_{s})\Gamma(b_{s}),
\end{eqnarray}
with the usual plasma dispersion function $Z(\xi_{s})$ written as
\begin{eqnarray}\label{eq:3.12}
Z(\xi_{s})=\frac{1}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\int_{-\infty}^{\infty}du e^{-u^2}[\frac{1}{u-\xi_{s}-i\sigma}],
\end{eqnarray}
and
\begin{eqnarray}\label{eq:3.13}
\Gamma(b_{s})=2\int_{0}^{\infty}dw w e^{-w^2}\Psi_{s}(b_{s}w).
\end{eqnarray}
The effects of the fractional velocity derivative can be boiled down to a non-Maxwellian contribution of the form
\begin{eqnarray}\label{eq:3.14}
W^{s}_{\mathbf{k},\omega}=\frac{\omega-\omega_{*,s}}{|k_{\parallel}|V_{T,s}}Z_{\epsilon}(\xi_{s})\Gamma_{\epsilon}(b_{s}),
\end{eqnarray}
where the non-Maxwellian plasma dispersion function is given by
\begin{eqnarray}\label{eq:3.15}
Z_{\epsilon}(\xi_{s})=\frac{V_{T,s}}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\int_{-\infty}^{\infty}du[\frac{\Phi(v_{\parallel})}{u-\xi_{s}-i\sigma}],
\end{eqnarray}
with the function $\Phi(v_{\parallel})$ being
\begin{eqnarray}\label{eq:3.16}
\Phi(v_{\parallel})=\frac{1}{2^{3/4}}\int \frac{dk^{v}_{\parallel}}{(2\pi)^{1/2}}\exp(-ik^{v}_{\parallel}v_{\parallel})\Lambda(k_{\parallel}^{v}).
\end{eqnarray}
It is important to note that the deviation from Maxwellian is different for the different species (electrons and ions). In the rest of Sec. 4, we will quantify the deviations. The non-Maxwellian contribution to Equation (\ref{eq:3.4}) is
\begin{eqnarray}\label{eq:3.17}
\Gamma_{\epsilon}(b_{s})=2V_{T,s}^2\int_{0}^{\infty}dw w \Psi_{s}(b_{s}w)\Phi(v_{\bot}),
\end{eqnarray}
where
\begin{eqnarray}\label{eq:3.18}
\Phi(v_{\bot})=\frac{1}{2^{3/4}}\int \frac{dk^{v}_{\bot}}{(2\pi)}\exp(-ik^{v}_{\bot}v_{\bot})\Lambda(k_{\bot}^{v}).
\end{eqnarray}
To extimate the non-Maxwellian contribution we need to determine the inverse Fourier transforms of the Equations (\ref{eq:3.16}) and (\ref{eq:3.18}) resulting in
\begin{eqnarray}\label{eq:3.18.1}
\Phi(z)=\frac{1}{8V_{T,s}^{3/2}}e^{-z^2} \nonumber \\
\left\{-4(-2+{\gamma_E})z^2+(-7+4\; {\gamma_E})+ 2 \log[V_{T,s}]+2e^{z^2}{_1 F_1}[\frac{3}{2},\frac{1}{2},-z^2]\right\}
\end{eqnarray}
with $z=\{u,w\}$ and ${_1 F_1}[a;b;z]$ denoting Kummer's confluent hypergeometric function. Therefore we can write
\begin{eqnarray}\label{eq:3.18.2}
W^{ad,s}= \frac{2V_{T,s}^3}{\sqrt{\pi}} \int_{-\infty}^{\infty}du\int_{0}^{\infty}wdw \Phi(u)\Phi(w).
\end{eqnarray}
By inserting typical values for the plasma parameters from Ref. \cite{BalescuBook} we find the velocities as $V_{T,e}=5.93\times 10^{9} [cm/s]$ and $V_{T,i}=1.38\times 10^{8} [cm/s]$ and we obtain
\begin{eqnarray}\label{eq:3.18.3}
W^{ad,e}=33.724\;\;\;\;\;\;\;\;\;,W^{ad,i}=23.6591.
\end{eqnarray}
Following the adiabatic condition in Equation (\ref{eq:2.37}) and the expanded dispersion relation in Equation (\ref{eq:3.10}) we obtain the following ratio between the non-Maxwellian contributions
\begin{eqnarray}\label{eq:3.18.4}
\frac{\epsilon_{i}}{\epsilon_{e}}=\frac{W^{ad,e}}{W^{ad,i}}=1.42541.
\end{eqnarray}
This relation means that if there is a deviation of the distribution function from the Maxwellian for plasma electrons, the deviation from the Maxwellian for ions will be $\sim 1.4$ larger.
\section{Solutions of the dispersion relation}
We will solve the dispersion relation in terms of expansions of the plasma dispersion function by noting that the drift waves are defined in the frequency range $|k_{\parallel}|V_{Ti}\ll\omega \ll |k_{\parallel}|V_{Te}$ in evaluating Equations (\ref{eq:3.12}) and (\ref{eq:3.15}). We define the expansion parameter for electrons in powers of $\xi_{e}=\omega/(|k_{\parallel}|V_{Te})\ll 1$ and for ions we expand it in powers of $\xi_{i}^{-1}=(|k_{\parallel}|V_{Ti})/\omega\ll 1$, respectively. The Maxwellian dispersion function $Z(\xi_{s})$ has the same definition as in Ref. \cite{Balescu1991}
\begin{eqnarray}\label{eq:3.19}
Z(\xi_{e})=\frac{1}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\int_{-\infty}^{\infty}du e^{-u^2}[\frac{1}{u-\xi_{e}-i\sigma}]
=-2\xi_{e}+\frac{4\xi_{e}^3}{3}+i\sqrt{\pi}(1-\xi_{e}^2)+\mathcal{O}[\xi_{e}^4],
\end{eqnarray}
whereas the non-Maxwellian plasma dispersion function $Z_{\epsilon}(\xi_{e})$ becomes
\begin{eqnarray}\label{eq:3.20}
Z_{\epsilon}(\xi_{e})=\frac{V_{T,e}}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\int_{-\infty}^{\infty}du[\frac{\Phi(u)}{u-\xi_{e}-i\sigma}]
=\frac{V_{T,e}}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\nonumber\\
\int_{-\infty}^{\infty}du\Phi(u)[\frac{1}{u-i\sigma}+\frac{\xi_{e}}{(u-i\sigma)^2}+\frac{\xi_{e}^2}{(u-i\sigma)^3}+\frac{\xi_{e}^3}{(u-i\sigma)^4}+\mathcal{O}[\xi_{e}^4]].
\end{eqnarray}
For ions, using the expansion in powers of $\xi_{i}^{-1}$ we can rewrite the above integrals as a function of the expansion parameter as
\begin{eqnarray}\label{eq:3.21}
Z(\xi_{i})=\frac{1}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\int_{-\infty}^{\infty}du e^{-u^2}[\frac{1}{u-\xi_{i}-i\sigma}]
=-\xi_{i}^{-1}-\frac{1}{2}\xi_{i}^{-3}+\mathcal{O}[\xi_{i}^{-5}],
\end{eqnarray}
and the non-Maxwellian $Z_{\epsilon}(\xi_{i})$ becomes
\begin{eqnarray}\label{eq:3.22}
Z_{\epsilon}(\xi_{i})=\frac{V_{T,i}}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\int_{-\infty}^{\infty}du[\frac{\Phi(u)}{u-\xi_{i}-i\sigma}]
=\frac{V_{T,i}}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\nonumber\\
\int_{-\infty}^{\infty}du\Phi(u)[\frac{1}{(-\xi_{i}-i\sigma)}-\frac{u}{(\xi_{i}+i\sigma)^2}+\frac{u^2}{(-\xi_{i}-i\sigma)^3}-\frac{u^3}{(\xi_{i}+i\sigma)^4}+\mathcal{O}[\xi_{i}^{-5}]].
\end{eqnarray}
We can now evaluate he Maxwellian integrals of the forms $\Gamma(b_{e})$ and $\Gamma(b_{i})$ assuming $\Psi_{e}=1$, $\Psi_{i}=J_{0}^2(b_{i}v_{\bot}/V_{Ti})$ we get
\begin{eqnarray}\label{eq:3.23}
\Gamma(b_{e})=2\int_{0}^{\infty}dw w e^{-w^2}=1,
\end{eqnarray}
and
\begin{eqnarray}\label{eq:3.24}
\Gamma(b_{i})=2\int_{0}^{\infty}dw w e^{-w^2}\Psi_{e}(b_{i}w)=e^{-b_{i}/2}\mathcal{I}_{0}(b_{i}),
\end{eqnarray}
where $\mathcal{I}_{0}$ denotes modified Bessel function of the zeroth order. The final result will be found after evaluating the non-Maxwellian $\Gamma_{\epsilon}(b_{e})$ and $\Gamma_{\epsilon}(b_{i})$ are given as
\begin{eqnarray}\label{eq:3.25}
\Gamma_{\epsilon}(b_{e})=2V_{T,e}^2\int_{0}^{\infty}dw w \Phi(w)=4.8\times 10^{5},
\end{eqnarray}
and
\begin{eqnarray}\label{eq:3.26}
\Gamma_{\epsilon}(b_{i})=2V_{T,i}^2\int_{0}^{\infty}dw w \Psi(b_{i}w)\Phi(w)=6.1\times 10^{4},
\end{eqnarray}
where we have used $V_{T,e}=5.93\;10^{9} [cm/s]$, $V_{T,i}=1.38\;10^{8} [cm/s]$ and $b_{i}=0.1$. Finally we can summarize different terms in the dispersion relation (\ref{eq:3.10}) as
\begin{eqnarray}\label{eq:3.27}
N^{e}_{\mathbf{k},\omega}=(\xi_{e}-\bar{\omega}_{*,e})(-2\xi_{e}+\frac{4\xi_{e}^3}{3}+i\sqrt{\pi}(1-\xi_{e}^2)),\nonumber\\
N^{i}_{\mathbf{k},\omega}=(\xi_{i}-\bar{\omega}_{*,i})(-\xi_{i}^{-1}-\frac{1}{2}\xi_{i}^{-3})e^{-b_{i}/2}\mathcal{I}_{0}(b_{i}),\nonumber\\
W^{e}_{\mathbf{k},\omega}=(\xi_{e}-\bar{\omega}_{*,e})Z_{\epsilon}(\xi_{e})\Gamma_{\epsilon}(b_{e}),\nonumber\\
W^{i}_{\mathbf{k},\omega}=(\xi_{i}-\bar{\omega}_{*,i})Z_{\epsilon}(\xi_{i})\Gamma_{\epsilon}(b_{i}),
\end{eqnarray}
where $\bar{\omega}_{*,s}=\omega_{*,s}/|k_{\parallel}|V_{T,s}$.
Note that the non-Maxwellian contributions in Equations (\ref{eq:3.20}), (\ref{eq:3.22}), (\ref{eq:3.25}) and (\ref{eq:3.26}) have been calculated numerically. By utilizing the found values of the integrals above we rewrite the dispersion relation (\ref{eq:3.10}) as follows
\begin{eqnarray}\label{eq:3.28}
(1+\epsilon_{e} W^{ad,e})+(\xi_{e}-\bar{\omega}_{*,e})\{-2\xi_{e}+\frac{4\xi_{e}^3}{3}+i\sqrt{\pi}(1-\xi_{e}^2)+\epsilon_{e} Z_{\epsilon}(\xi_{e}) \Gamma_{\epsilon}(b_{e})\}=\nonumber\\
-(1+\epsilon_{i} W^{ad,i})-(\xi_{i}-\bar{\omega}_{*,i})\{(-\xi_{i}^{-1}-\frac{1}{2}\xi_{i}^{-3})e^{-b_{i}/2}\mathcal{I}_{0}(b_{i})+\epsilon_{i} Z_{\epsilon}(\xi_{i})\Gamma_{\epsilon}(b_{i})\}
\end{eqnarray}
where $W^{ad,s}$ are given in Equation (\ref{eq:3.18.3}) and we will use the ratio between $\epsilon_{e}$ and $\epsilon_{i}$ from Equation (\ref{eq:3.18.4}).
\section{Results and discussion}
We have derived a dispersion relation for drift waves driven by a density gradient in a shear-less slab geometry with constant magnetic field where the small deviation from a Maxwellian distribution is described by $\epsilon$. Here we will determine the quantitative effects on the real frequency and growth rate as a function of this deviation. We start by assuming that we have adiabatic electrons for which the dispersion Equation (\ref{eq:3.28}) is,
\begin{eqnarray}\label{eq:4.1}
2+\epsilon_{i} (2.35\;W^{ad,e}+ W^{ad,i})=\nonumber\\
-(\xi_{i}-\bar{\omega}_{*,i})\{(-\xi_{i}^{-1}-\frac{1}{2}\xi_{i}^{-3})e^{-b_{i}/2}\mathcal{I}_{0}(b_{i})+\epsilon_{i} Z_{\epsilon}(\xi_{i})\Gamma_{\epsilon}(b_{i})\}.
\end{eqnarray}
After rearranging the terms in the above equation we finally get the following relation for $\epsilon_{i}$:
\begin{eqnarray}\label{eq:4.1.1}
\epsilon_{i} =\frac{-2\xi_{i}^3+[\xi_{i}^3+0.5\xi_{i}-\bar{\omega}_{*,i}\xi_{i}^2-0.5\bar{\omega}_{*,i}]e^{-b_{i}/2}\mathcal{I}_{0}(b_{i})}{W^{ad,tot}\xi_{i}^3+(\bar{\omega}_{*,i}\xi_{i}^3-\xi_{i}^4)Z_{\epsilon}(\xi_{i})\Gamma_{\epsilon}(b_{i})}
\end{eqnarray}
where $W^{ad,tot}=2.35\;W^{ad,e}+ W^{ad,i}$. This relation gives the possible deviation of the equilibrium PDF from the Maxwellian PDF for a given plasma turbulence, i.e $\xi_{i}$. One has to remember that only positive values of $\mathbf{Re}[\epsilon]$ are physically meaningful.
Using the same plasma parameters as was used in Equations (\ref{eq:3.18.3}) and (\ref{eq:3.24},\ref{eq:3.26}) we compute the term $Z_{\epsilon}(\xi_{i})$, and from Equation (\ref{eq:3.22}) we get
\begin{eqnarray}\label{eq:4.2}
Z_{\epsilon}(\xi_{i})=\frac{V_{T,i}}{\sqrt{\pi}}Lim_{\sigma\rightarrow 0}\{\frac{1}{(-\xi_{i}-i\sigma)}\int_{-\infty}^{\infty}du\Phi(u)+\frac{1}{(-\xi_{i}-i\sigma)^3}\int_{-\infty}^{\infty}u^2du\Phi(u)\}\nonumber\\
=\frac{-6.5\times 10^{-9} - 3.8 \times 10^{-9} \xi_{i}^2}{\xi_{i}^3}.
\end{eqnarray}
Here, those integrations omitted resulted in zero contributions and rewriting Equation (\ref{eq:4.1}) by using these explicit values results in the expression for the deviation in Equation (\ref{eq:4.1.1}) we obtain
\begin{eqnarray}\label{eq:4.4}
\epsilon_{i} =\frac{-2\xi_{i}^3+[\xi_{i}^3+0.5\xi_{i}-\bar{\omega}_{*,i}\xi_{i}^2-0.5\bar{\omega}_{*,i}]e^{-b_{i}/2}\mathcal{I}_{0}(b_{i})}{66.3\xi_{i}^3-39.2\xi_{i}+39.2\bar{\omega}_{*,i}+23.0 \bar{\omega}_{*,i}\xi_{i}^2}
\end{eqnarray}
Figure \ref{fig1} shows $\epsilon_{i}$ from Equation (\ref{eq:4.4}) where $\xi_{i}=\omega+i \gamma$. Here, the values of $\omega, \gamma$ are normalized to $|k_{\parallel}|V_{T,i}$. We have assumed parameter values $b_{i}=0.1$, $k_{\parallel}=10^{-3}$ and $\bar{\omega}_{*,i}=-7.1\times 10^{2}$ with $d\;ln\;n/dx=1$. It is found that there is a threshold in the growth rate $\gamma$ close to $\gamma = 0.7$ and that increasing to $1.0$ only increases the deviation from a Maxwellian from 0 to 0.03. It should be noted that $\epsilon$ increases the excess kurtosis of the distribution function by a similar amount thus a quite small deviation from a Maxwellian can have a rather significant impact.
In figure \ref{fig2}, the mode growth rate as a function of $\epsilon_{i}$ is shown. Note that in this figure the values of growth rate are the solutions of the Equation (\ref{eq:4.4}) for a given $\epsilon_{i}$ while in the figure \ref{fig1} we solve Equation (\ref{eq:4.4}) for $\epsilon$ at a given $\xi_{i}$. As the dispersion equation is of 3rd order in $\bar{\omega}$ three possible solutions exist, however we are only interested in the solutions with non-zero imaginary value, $\gamma>0$ corresponding to unstable situations. It is shown in figure \ref{fig2} that a deviation of $\epsilon_{i}=0.01$ yield an increase of about $20\%$ in the growth rate. Furthermore, the growth rate increases almost linearly with increasing $\epsilon_{i}$ and such an increase in the growth rate will lead to a significant increase in the level of anomalous flux.
\begin{figure}[tbp]
\begin{center}
\epsfig{figure=Figure1.eps, width=10cm,height=8cm,clip=}
\end{center}
\caption{$\epsilon$ as a function of $\omega$ and $\gamma$ where $\xi_{i}=\omega+i \gamma$. We have assumed $b_{i}=0.1$, $k_{\parallel}= 10^{-3}$ and $\bar{\omega}_{*,i}=-7.1\times 10^{2}$ with $d\;ln\;n/dx=1$.}
\label{fig1}
\end{figure}
\begin{figure}[tbp]
\begin{center}
\epsfig{figure=Figure2.eps, width=10cm,height=8cm,clip=}
\end{center}
\caption{$\gamma$ as a function of $\epsilon$. The same plasma parameters as in the figure \ref{fig1} are used.}
\label{fig2}
\end{figure}
In summary, we have derived a dispersion relation for density gradient driven linear drift waves including the effects coming from the inclusion of a fractional velocity derivative in the Fokker-Planck equation in the case of constant magnetic field and a shear-less slab geometry. The solutions of this Fokker-Planck equation are the alpha-stable distributions. It has not yet been shown that in a direct way one can derive the alpha-stable distribution function \cite{BalescuBook,montroll} from the classical form of collision operator \cite{Gatto}. One way may be to construct a new type of collisional operator by considering a fractal phase space and reformulate the collision operator on this new space. However, such a discussion is outside the scope of the present paper. Interestingly enough, we note that non-local effects are observed in non-linear collisionless fluid simulations of plasma turbulence where the non-local transport showing Levy features are induced by the interaction of the non-linear terms in the dynamical equations \cite{negrete2005}. The non-local features of non-linear fluid models are indicated by recent analytical theories using path-integral methods to derive probability density fucntions of fluxes \cite{anderson1}.
The fractional derivative is represented with the Fourier transform containing a fractional exponent that we are able to connect to the deviation from a Maxwellian distribution described by $\epsilon$. The characteristics of the plasma drift wave are fundamentally changed, i.e. the values of the growth-rate $\gamma$ and real frequency $\omega$ are significantly altered. A deviation from the Maxwellian distribution function alters the dispersion relation for the density gradient drift waves such that the growth rates are substantially increased and thereby may cause enhanced levels of transport.
{\em Acknowledgements}
The authors would like to thank professor T. F\"ul\"op for her helpful comments. This work was funded by the European Communities under Association Contract between EURATOM and
{\em Vetenskapsr{\aa}det}.
\section*{References}
|
\section{Introduction}
Two years ago Ho\v rava proposed a gravity theory (now commonly referred to as Ho\v rava--Lifshitz gravity or simply Ho\v rava gravity) that has some realistic hope to be a UV completion of general relativity \cite{Horava:2009uw}. This is to be achieved by adding to the gravitational action higher-order spatial derivatives without adding higher-order time derivatives. This procedure can lead to a modification of the graviton propagator which renders the theory power-counting renormalizable, without increasing the number of time derivatives in the field equations \cite{Horava:2009uw, power-count1, power-count2}.
Clearly such a theory cannot treat space and time on the same footing, and is naturally constructed in terms of a preferred foliation. This foliation can be described by a scalar field, which is an extra degree of freedom of the theory with respect to general relativity. The theory is not invariant under the full set of diffeomorphisms, but it can still be invariant under the more restricted foliation-preserving diffeomorphisms, $t\to\tilde{t}(t)$, $x^i\to\tilde{x}^i(t,x^i)$. It is convenient to consider the Arnowitt--Deser--Misner decomposition of spacetime
\begin{equation}
{\mathrm{d}} s^2 = - N^2 c^2 {\mathrm{d}} t^2 + g_{ij}({\mathrm{d}} x^i + N^i {\mathrm{d}} t) ({\mathrm{d}} x^j + N^j {\mathrm{d}} t).
\end{equation}
Defining the extrinsic curvature as
\begin{equation}
K_{ij} = {1\over2N} \left\{ \dot g_{ij} - \nabla_i N_j - \nabla_j N_i \right\},
\end{equation}
the action of the theory is of the form
\begin{equation}
\label{action}
S=\frac{M_{\rm pl}^2}{2}\int {\mathrm{d}}^dx \,{\mathrm{d}} t \, N \sqrt{g} \left\{ K^{ij} K_{ij} - \lambda K^2 +{\cal V}\right\}\, ,
\end{equation}
where Latin indices run from $1$ to $3$, $M_{\rm pl}$ is the Planck mass, $g$ is the determinant of the spatial metric $g_{ij}$, while $\lambda$ is a dimensionless running coupling, and ${\cal V}$ is the part of the Lagrangian that does not contain any time derivatives. In fact, invariance under foliation-preserving diffeomorphisms requires that ${\cal V}$ does not include any terms containing the shift $N^i$, but is instead constructed only with the lapse $N$, the induced metric $g_{ij}$, and their spatial derivatives.
Power counting renormalizability, on the other hand, requires that ${\cal V}$ include terms with up to $2z$ derivatives, where $z\geq d$, and $d$ is the number of spatial dimensions \cite{Horava:2009uw, power-count1, power-count2}. When $d=3$, one should have at least $z=3$, {\em i.e.}~6th-order, operators in ${\cal V}$. Additionally, radiative corrections are expected to generate all possible terms compatible with the symmetries of the theory. This leads to a very large number of terms that one needs to include in $\cal V$. Restricting the lapse to be a function of time only, $N=N(t)$, as has been suggested in reference~\cite{Horava:2009uw} in order to match the reduced symmetry of the theory, leads to a much smaller and tractable number of terms \cite{Sotiriou:2009bx, Sotiriou:2009xxx} --- the so-called projectable version of the theory, see references \cite{Weinfurtner:2010hz,Mukohyama:2010xz} for reviews. However, in this version of the theory, as well as in other versions with different restrictions, such as detailed balance \cite{Horava:2009uw}, the extra scalar degree of freedom exhibits undesirable dynamical behaviour, such as instabilities, over-constrained evolution, and strong coupling at low energies \cite{Sotiriou:2009xxx,bunch1, bunch2, bunch3, bunch4, Kobakhidze:2009zr,bunch5,bunch6,bunch7,bunch8,bunch9,Wang:2010uga}. (We will not consider in this article models where the action has been modified in order to respect extra symmetries, such as the model with an extra local $U(1)$ symmetry proposed in reference~\cite{Horava:2010zj}. See reference \cite{Sotiriou:2010wn} for a brief review including all of the various versions of Ho\v rava--Lifshitz gravity.)
It turns out that the scalar mode is much better behaved if no projectability or detailed balance restriction is imposed \cite{Blas:2009qj}. Then ${\cal V}$ has the general form
\begin{equation}
\label{genV}
{\cal V}=\xi \, R+\eta \,a^i a_i+\frac{1}{M_\star^2}L_4+\frac{1}{M_\star^4}L_6\,,
\end{equation}
where $R$ is the Ricci scalar of $g_{ij}$,
\begin{equation}
a_i=\partial_i \ln N\,,
\end{equation}
while $\xi$ and $\eta$ are dimensionless couplings, $M_\star$ is a new mass scale, and $L_4$ and $L_6$ include all possible terms of 4th and 6th order in spatial derivatives respectively. A cosmological constant can also be added to ${\cal V}$ but we neglect it here for simplicity as it is not important for our discussion. The presence of the $a^i a_i$ term, in addition to the standard $R$ term, is enough to alleviate the aforementioned problems at low energies.
Strong coupling persists even in this version of the theory and, though it is pushed up to high energies, it still constitutes a potential threat for UV completeness \cite{Papazoglou:2009fj,Kimpton:2010xi}. However, it can be altogether avoided by assigning a specific hierarchy between the scales $M_{pl}$ and $M_\star$ \cite{Blas:2009ck}.
There is still a long way to go before one can say with confidence whether Ho\v rava's proposal is a truly viable UV complete gravity theory. Renormalizability beyond power-counting has not been demonstrated, and the renormalization group flow of the various running couplings is not yet known (despite the fact that infrared viability hinges on the hope that various parameters will run sufficiently rapidly to desired values). Furthermore, there are various phenomenological aspects of the theory which have not yet been studied, which will lead to new constraints ({\em e.g.}~coupling to matter, equivalence principle violations, etc.). Despite these limitations, for the moment Ho\v rava--Lifshitz gravity definitely seems to be an interesting candidate for a UV complete gravity theory, one which deserves further study.
A purely technical difficulty of the theory is the very large number of terms that one needs to consider as part of $L_4$ and $L_6$. One way to deal with this problem without imposing restriction to the action would be to study the theory in less that 3 spatial dimensions. Since renormalizability requires $z=d$ (at least), taking $d<3$ would also reduce the number of spatial derivatives one would have to allow, so it would also consequently reduce drastically the number of higher-order operators. Therefore, one could consider studying lower-dimension Ho\v rava--Lifshitz gravity in an attempt to gain insight into the $3+1$ dimension theory.
As discussed, the problem with imposing restrictions to the action in order to reduce the number of higher-order operators is essentially that it alters the dynamics of the theory. There is no particular reason to expect that this is not also going to be the case when one reduces the number of spatial dimensions. Indeed, we already know for a fact that the dynamics of general relativity changes drastically when the number of spatial dimensions drops below $3$. Our goal here is precisely to examine to which extent the dynamics of lower-dimensional Ho\v rava--Lifshitz gravity resembles that of the $3+1$ dimensional theory. This will allow us to gauge how much we can learn for the latter from the former. In what follows, we will separately consider $1+1$ and $2+1$ Ho\v rava--Lifshitz gravity, we will construct the full actions and discuss their dynamics, utilizing (and extending to lower dimensions) also the equivalence \cite{Jacobson:2010mx} between Einstein-aether theory \cite{Jacobson:2000xp,Jacobson:2008aj} and the low energy limit of Ho\v rava--Lifshitz gravity.
Note that lower-dimensional models of Ho\v rava--Lifshitz gravity are also interesting in their own right. Renormalizable gravity theories, even in less that $3+1$ dimensions, are not so easy to find. Such models could potentially be used as duals to non-relativistic quantum field theories in the context of the AdS/CFT correspondence, or as theories describing gravity on worldsheets of strings or worldvolumes of branes \cite{Horava:2008ih}. See also reference~\cite{Horava:2011gd} for further examples.
\section{Ho\v rava--Lifshitz gravity in $1+1$ dimensions}
\subsection{Setup and most general action}
We start by considering the simplest case of $1+1$ dimensions. The action of the theory in this case is drastically simplified as 1-dimensional space cannot have intrinsic curvature. At the same time the extrinsic curvature is actually a scalar, given by
\begin{equation}
K= {1\over2N} \left\{ \dot g_{11} - 2\nabla_1 N_1\right\},
\end{equation}
Finally, we have $z=d=1$, so ${\cal V}$ needs to contain space-covariant terms we can construct with $g_{ij}\to g_{11}$, and $a_i\to a_1$, with up to 2 spatial derivatives only. Thus, the action has the simple form
\begin{equation}
\label{1daction}
S=\frac{M_{\rm pl}^2}{2}\int {\mathrm{d}} x \, {\mathrm{d}} t \, N \sqrt{g_{11}} \left\{ (1 - \lambda) \; K^2 +\eta \;g^{11}a_1a_1\right\}\, .
\end{equation}
When $\lambda=1$ and $\eta=0$, which corresponds to the values of these parameters in general relativity, the theory becomes trivial as expected. We will not attempt to derive field equations for this theory. Instead, in the next section we will establish its dynamical equivalence with Einstein-aether theory in 2 dimensions, by applying the results of reference~\cite{Jacobson:2010mx}. We will then use this equivalence to discuss the dynamics.
\subsection{Equivalence with Einstein-aether theory and dynamics}
\label{1+1equiv}
The action for Einstein-aether theory \cite{Jacobson:2000xp,Jacobson:2008aj} is
\begin{equation} \label{S}
S^{4d}_{\ae} = \frac{1}{16\pi G_{\ae}}\int {\mathrm{d}}^{4}x\, \sqrt{-\tilde{g}}~ (-\tilde{R} -M^{\alpha\beta\mu\nu} \tilde{\nabla}_\alpha u_\mu \tilde{\nabla}_\beta u_\nu)
\end{equation}
where Greek indices run from $0$ to $3$, $\tilde{R}$ is the 4-dimensional Ricci scalar of the spacetime metric $\tilde{g}_{ab}$, $\tilde{g}$ is the determinant of that metric, $ \tilde{\nabla}_\alpha$ the associated covariant derivative, and
\begin{equation}
M^{\alpha\beta\mu\nu} = c_1 \tilde{g}^{\alpha\beta}\tilde{g}^{\mu\nu}+c_2\tilde{g}^{\alpha\mu}\tilde{g}^{\beta\nu}+c_3 \tilde{g}^{\alpha\nu}\tilde{g}^{\beta\mu}+c_4 u^\alpha u^\beta \tilde{g}_{\mu\nu}\,,
\end{equation}
(a tilde will be used to denote a spacetime metric, as opposed to an induced spatial metric, irrespectively of dimensionality). $G_{\ae}$ has dimensions of an inverse mass squared, whereas the parameters $c_i$ are dimensionless. Additionally, the aether field, $u_\mu$ is forced to satisfy the constraint $u_\mu u^\mu=1$. Here we assume that this constraint is imposed implicitly by allowing only variations that respect it. Alternatively, one could impose it explicitly, by the use of a Lagrange multiplier. It has been shown in reference~\cite{Jacobson:2010mx} that, once the extra restriction that the aether is hypersurface orthogonal has been imposed, this aether action is dynamically equivalent to the low-energy limit of Ho\v rava--Lifshitz gravity in $3+1$ dimensions, {\em i.e.}~to the action given in equation~(\ref{action}), \emph{but without the higher-order operators $L_4$ and $L_6$}. Given also the unit vector constraint on the aether, this restriction amounts locally to the requirement that there exists a function such that
\begin{equation}
\label{ho}
u_\alpha=\frac{\partial_\alpha T}{\sqrt{\tilde{g}^{\mu\nu}\; \partial_\mu T \partial_\nu T}}\,,
\end{equation}
However, one could choose to work in a gauge where $T$ is identified with the time coordinate $t$. Then
\begin{equation}
\label{hou}
u_\mu=\delta_{\mu T}\; (g^{TT})^{-1/2}=N\delta_{\mu T}\,.
\end{equation}
If one uses equation~(\ref{hou}) to replace $u_\mu$ in equation~(\ref{S}), the low-energy limit of the action of Ho\v rava--Lifshitz gravity in $3+1$ dimensions is recovered, that is, the action given in equation~(\ref{action}) without the higher-order operators $L_4$ and $L_6$, and with the following correspondence of parameters:
\begin{eqnarray}
\label{HLpar}
&&\quad\frac{1}{8\pi M_{pl}^2 \,G_{\ae}}=\xi=\frac{1}{1-c_{13}},\\\nonumber\\&& \lambda=\frac{1+c_2}{1-c_{13}},\quad\qquad \eta=\frac{c_{14}}{1-c_{13}},\nonumber
\end{eqnarray}
where $c_{ij}=c_i+c_j$.
Clearly, this equivalence will apply also apply in 2 dimensions (or indeed any number of dimensions). The action of 2-dimensional Einstein-aether theory is \cite{Eling:2006xg}
\begin{equation} \label{S2}
S^{2d}_{\ae} = -\frac{1}{16\pi G_{\ae}}\int {\mathrm{d}}^{2}x\; \sqrt{-\tilde{g}}L^{2d}_{\ae} ,
\end{equation}
where
\begin{equation}
L^{2d}_{\ae}=\frac{1}{2}c_{14} F^{\alpha\beta}F_{\alpha\beta}+c_{123}(\tilde{\nabla}_\alpha u^\alpha)^2\,,
\end{equation}
and we have ignored total divergences (such as the $\tilde{R}$ term). Greek indices now take the values $0$ and $1$ only, and
\begin{equation}
F_{\alpha\beta}=\tilde{\nabla}_\alpha u_\beta-\tilde{\nabla}_\beta u_\alpha\,.
\end{equation}
In 2 dimensions any vector field is always locally hypersurface orthogonal and, thus, so is the aether. It is straightforward to show that, after choosing $T$ as the time coordinate,
\begin{eqnarray}
&&\tilde{\nabla}_\alpha u^\alpha=-K\,,\qquad F^{\alpha \beta}F_{\alpha\beta}=-g^{11}a_1 a_1\,.
\end{eqnarray}
Then, the action in equation~(\ref{S2}) becomes the action of Ho\v rava--Lifshitz gravity in $1+1$ dimensions, as given in equation~(\ref{1daction}), with the parameter correspondence as given in equation~(\ref{HLpar}).
As was mentioned earlier, 4-dimensional Einstein-aether theory with the extra restriction of the aether being hypersurface orthogonal is equivalent to the low-energy limit of Ho\v rava--Lifshitz gravity in $3+1$ dimensions, not the full theory. However, in Ho\v rava--Lifshitz gravity in $1+1$ dimensions no operators with more than two spatial derivatives need to be considered anyway, as naive power counting renormalizability demands $z=d=1$. Therefore, the complete $1+1$ Ho\v rava theory turns out to be equivalent to 2-dimensional Einstein-aether theory. Additionally, it is worth stressing once more that in 2 dimensions the aether is necessarily hypersurface orthogonal, so no further restriction on the Einstein-aether side is needed either.
In reference~\cite{Eling:2006xg} the full set of solutions of $2$-dimensional Einstein-aether theory has been found. Though nontrivial, unlike general relativity in 2 dimensions, the theory does not posses any local degrees of freedom. The equivalence presented above can be used to turn these solutions of Einstein-aether theory into solutions of Ho\v rava--Lifshitz gravity in $1+1$ dimensions --- then similar conclusions can be made for the dynamics of the latter. Based on this, although it might constitute an interesting model for a 2-dimensional quantum gravity theory, we see that Ho\v rava--Lifshitz gravity in $1+1$ dimensions is drastically different from its $3+1$ counterpart.
It is also worth mentioning that naive power-counting renormalizability should be taken with a grain of salt in $1+1$ dimensions. Since there are no local degrees of freedom the perturbative power-counting arguments do not (strictly speaking) apply. However, again due to this lack of local degrees of freedom, the theory is likely to still be non-perturbatively renormalizable, like 2+1 general relativity.
\subsection{$1+1$ projectable theory as a limiting case}
\label{proj1+1}
As mentioned in the Introduction, our main interest is the most general Ho\v{r}ava--Lifshitz gravity without any restrictions, as the more restricted versions tend to have viability and consistency problems. However, it is rather trivial for one to obtain the most general action and understand the dynamics of projectable Ho\v rava--Lifshitz gravity as a limiting case of what has been presented above. The difference between the projectable version and the version we considered here is that in the former the lapse is forced to be a function of time only, {\em i.e}~$N=N(t)$, at the level of the action.
In $1+1$ dimensions requiring $N=N(t)$ would make the last term in eq.~(\ref{1daction}) vanish, so the most general action in the projectable version is
\begin{equation}
\label{1dactionp}
S_p=\frac{M_{\rm pl}^2}{2}\int {\mathrm{d}} x \, {\mathrm{d}} t \, N \sqrt{g_{11}} (1 - \lambda) \; K^2 \, .
\end{equation}
Clearly, provided that $\lambda\neq 1$ and possibly modulo a sign, the factor $(1-\lambda)$ can be absorbed in a redefinition of $M_{\rm pl}$. So, the crucial difference between projectable Ho\v rava--Lifshitz gravity and general relativity in $1+1$ dimensions is simply the fact that $N=N(t)$, and nothing more.
\section{Ho\v rava--Lifshitz gravity in $2+1$ dimensions}
\subsection{Setup and most general action}
We now turn our attention to 2 spatial dimensions. The action can take the form given in equation~(\ref{action}), with $K_{ij}$ being the extrinsic curvature of the 2-dimensional spatial hypersurfaces. What remains is to determine ${\cal V}$. We have $z=d=2$, so ${\cal V}$ needs to contain all space-covariant terms we can construct with $g_{ij}$ and $a_i$, with up to 4 spatial derivatives. Some of these terms, however, can be eliminated if one takes into account the following:
\begin{itemize}
\item
In 2 dimensions we know that
\begin{equation}
R_{abcd}=\frac{1}{2}(g_{ac}g_{db}-g_{ad}g_{cb}) R,
\end{equation}
so all curvature invariants can be expressed in terms of the Ricci scalar.
\item Various terms differ only by a total divergence.
\item The vector $a_i$ is the gradient of a scalar.
\end{itemize}
Once all of the above have been taken into account one can, without any loss of generality, write
\begin{eqnarray}
{\cal V}&=&\xi\,R+\eta\, a^i a_i+g_1 \,R^2+g_2\, \nabla^2R+g_3\,(a^ia_i)^2\nonumber\\
&&+g_4\, Ra^ia_i+g_5 a^2 (\nabla\cdot a) + g_6 (\nabla\cdot a)^2 \nonumber\\
&&+ g_7 (\nabla_i a_j) (\nabla^i a^j).
\end{eqnarray}
The $g_i$ couplings are not dimensionless, but have dimensions of an inverse mass squared ({\em i.e.}~we have absorbed the mass scale $M_\star$ into these couplings). Again, we can add a cosmological constant which we neglect here for simplicity. The most general action in $2+1$ dimension then has the form
\begin{eqnarray}
\label{2daction}
S&=&\frac{M_{pl}^2}{2}\int {\mathrm{d}}^2 x \, {\mathrm{d}} t \, N\sqrt{g} \Big\{K^{ij}K_{ij}-\lambda K^2+\xi R +\eta\, a_i a^i \nonumber\\
&&\qquad\qquad+g_1 \,R^2+g_2\, \nabla^2R+g_3\,(a^ia_i)^2\nonumber\\&&\qquad\qquad+g_4\, Ra^ia_i+g_5 a^2 (\nabla\cdot a) + g_6 (\nabla\cdot a)^2 \nonumber\\&&\qquad\qquad+ g_7 (\nabla_i a_j) (\nabla^i a^j)\Big\}.
\end{eqnarray}
We can obtain the field equations by varying the action with respect to the lapse $N$, the shift $N^i$, and the induced metric $g_{ij}$.
Variation with respect to the lapse yields
\begin{eqnarray}
\label{eqlapse}
&&K^{ij}K_{ij}-\lambda K^2+\xi R+ \eta\left[2 \nabla\cdot a +a^ia_i\right]\\&&-g_1 R^2-g_2 \nabla^2 R+g_3\left[4\nabla_i(a_ja^j a^i)+3(a_ia^i)^2\right]\nonumber\\&&+g_4 \left[2 \nabla_i(R a^i) +Ra^ia_i\right]+g_5\big\{a^ia_i \nabla\cdot a\nonumber\\&&+2\nabla\cdot (a[\nabla\cdot a])-[\nabla^2(Na^ia_i)]/N\big\}\nonumber\\&&-g_6\big[3 (\nabla\cdot a)^2 + 2 \nabla^2 (\nabla\cdot a) + 2 a^2 (\nabla\cdot a) \nonumber\\&&+4 (a\cdot\nabla) (\nabla\cdot a)\big]
-g_7\big[ (\nabla_i a_j) (\nabla^i a^j) \nonumber\\&&-2 (\nabla_i a_j) a^i a^j -4 \nabla^i ( a^j (\nabla_i a_j) ) - 2 \nabla_i \nabla^2 a^i\big]\,
=0.\nonumber\\\nonumber
\end{eqnarray}
Variation with respect to the shift yields
\begin{equation}
\label{eqshift}
\nabla_i\pi^{ij}\equiv \nabla_i\left\{K^{ij}-\lambda Kg^{ij}\right\}=0\,.
\end{equation}
\def{\mathcal{L}}{{\mathcal{L}}}
Variation with respect the spatial metric $g_{ij}$ yields
\begin{widetext}
\begin{eqnarray}
\label{piequation}
&&\frac{1}{\sqrt{g}} \; \partial_t ( \sqrt{g} \; \pi_{ij} ) +2N \left(K_i^{\phantom{a}l}K_{lj} - \lambda K K_{ij}\right)
- {1\over2} N \left(K^{ij}K_{ij}-\lambda K^2\right) g^{ij}
- N \pi^{ij} (\nabla_m N^m)- {\mathcal{L}}_{\vec N} (N\pi_{ij})
\nonumber\\
&& +2 N N_{(i}\pi_{j)m}a^m-\xi \frac{1}{N}[\nabla_i\nabla_j -g_{ij}\nabla^2]N +\eta\,\left[a_i a_j -\frac{1}{2} a^l a_lg_{ij}\right]+\frac{1}{2} g_1R^2 g_{ij} -2 g_1\frac{1}{N}[\nabla_i\nabla_j -g_{ij}\nabla^2] (NR)\nonumber\\&&-g_2 \frac{1}{N}[\nabla_i\nabla_j -g_{ij}\nabla^2] (\nabla^2 N) -g_2 a_{(i} \nabla_{j)} R
+g_2 {\nabla_m(NRa^m)\over2N} g_{ij}+2 g_3 a^2 a_i a_j -{1\over2} g_3(a^2)^2 g_{ij}+g_4
R a_i a_j\nonumber\\&&
-g_4 \frac{1}{N}[\nabla_i\nabla_j -g_{ij}\nabla^2] (N a^i a_i) - {1\over2} g_4 Ra^2 g_{ij} +g_5 a_i a_j [(\nabla\cdot a) - a^2]
-{1\over2}g_5 \left\{ (\nabla_i a^2) a_j + (\nabla_j a^2) a_i \right\}\nonumber\\&&
+ {1\over2} g_5\left\{ (a^2)^2+ a^m\nabla_m a^2 \right\} g_{ij}-2g_6 (\nabla\cdot a) a_i a_j -g_6\left\{ (\nabla_i [\nabla\cdot a]) a_j + (\nabla_j i [\nabla\cdot a) a_i \right\}\nonumber\\
&&
+ g_6\left[a^2 (\nabla\cdot a) + a\cdot\nabla (\nabla\cdot a) +{1\over2} (\nabla\cdot a)^2 \right]g_{ij}
-{1\over2} g_7 (\nabla_m a_n) (\nabla^m a^n) g_{ij} - 2 g_7(a\cdot\nabla a_{(i}) a_{j)} -2 g_7(\nabla^2 a_{(i}) a_{j)}\nonumber\\
&&
+
g_7 a^2 (\nabla_i a_j) + g_7\nabla\cdot( [\nabla_i a_j] a)=0\,,
\end{eqnarray}
\end{widetext}
where $ {\mathcal{L}}_{\vec N}$ denotes the Lie derivative along the vector $N^i$.
Though rather lengthy, these equations are perhaps more manageable than they seem. Equations~(\ref{eqlapse}) and (\ref{eqshift}) do not contain any time derivatives of the lapse or the shift and are, therefore, constraints. The only dynamical equation is~(\ref{piequation}) and the only dynamical variable is $g_{ij}$. However, applying the uniformization theorem
\cite{poincare}, since $g_{ij}$ is a 2-dimensional metric and the theory is invariant under transformation of the sort $x^i\to\tilde{x}^i(t,x^i)$, there is enough gauge freedom to set
\begin{equation}
\label{gauge}
g_{ij}=\Omega^2\; g^c_{ij}\,,
\end{equation}
where $g^c_{ij}$ denotes the metric of a constant curvature spherical, Euclidean or hyperbolic 2-dimensional space. This gauge choice would turn equation~(\ref{piequation}) into a dynamical equation for the conformal factor $\Omega$. We will not proceed further to explicitly write the equation in this specific gauge, as the current discussion suffices to convincingly argue that the theory propagates only a single scalar degree of freedom.
\subsection{Linearization and dynamics}
To provide further support for our claim that there is only a single scalar degree of freedom, we linearize the theory around flat 2+1 Minkowski space. This will also allow us to determine the linearized propagator for the scalar, and get a deeper insight into the dynamics. We start with the action given in equation~(\ref{2daction}), and we perturb to quadratic order. We have
\begin{equation}
N=1+\alpha, \qquad\qquad N_i= 0 + n_i,
\end{equation}
and we make the gauge choice
\begin{equation}
g_{ij}=e^{2\zeta} \delta_{ij}.
\end{equation}
The terms $(a^ia_i)^2$, $Ra^ia_i$ and $a^2 (\nabla\cdot a)$ will not contribute to this order (since by construction $a_i=R=0$ in the background). On the other hand, $(\nabla\cdot a)^2$ and $(\nabla_i a_j) (\nabla^i a^j)$ give the same (non-trivial) contribution to this order, so we define (with some foresight) $g_{67}=(g_6+g_7)/2$. The quantity $K_{ij}$ appears only quadratically in the action, so we only need to compute it to first order:
\begin{eqnarray}
&&K^{(1)}_{ij}=\dot{\zeta} \delta_{ij}-\partial_{(i} n_{j)}\,,\\
&&K^{(1)}= 2 \dot{\zeta} -\partial_i n^i \,.
\end{eqnarray}
The quadratic action then takes the form
\begin{eqnarray}
S&=&M_{pl}^2\int {\mathrm{d}}^2 x \, {\mathrm{d}} t \, \Big\{(1-2\lambda)\dot{\zeta}^2- (1-2\lambda)\dot{\zeta} (\partial_i n^i)\nonumber\\
&&\quad\qquad+\frac{1}{4}(1-2\lambda) (\partial_i n^i)^2
+{1\over4} n^i \partial^2 n_i -\xi \alpha \partial^2 \zeta \nonumber\\
&&\quad\qquad+\frac{\eta}{2} (\partial_i\alpha)(\partial^i\alpha)
+2 g_1(\partial^2\zeta)^2\nonumber\\
&&\quad\qquad-2 g_2 (\partial^2 \alpha) (\partial^2 \zeta)+g_{67}(\partial^2 \alpha)^2\Big\},
\end{eqnarray}
where $\partial^2=\delta^{ij}\partial_i\partial_j$.
We start by varying with respect to $n_i$. This variation yields
\begin{equation}
\label{neq}
{1\over2}(1-2\lambda)\partial_i (\partial_j n^j) +{1\over2} \partial^2 n_i =(1-2\lambda) \partial_i \dot{\zeta}\,.
\end{equation}
Taking the divergence leads to the equation
\begin{equation}
\label{interimni}
(1-\lambda)\partial^2 (\partial_i n^i) =(1-2\lambda) \partial^2 \dot{\zeta}\,,
\end{equation}
which, (given suitable regular boundary conditions), can be trivially integrated to give
\begin{equation}
\label{interimbeta}
\partial_i n^i=\frac{1-2\lambda}{1-\lambda} \dot{\zeta}\,.
\end{equation}
Re-inserting this into equation~(\ref{neq}) one gets
\begin{equation}
\partial_i (\partial_j n^j) -\partial^2 n_i =0\,.
\end{equation}
In particular, this implies
\begin{equation}
\partial^2 n_{[i,j]} =0\,,
\end{equation}
which in turn (given suitable regular boundary conditions) implies that
\begin{equation}
\label{eqni}
n_{[i,j]} =0 \quad \Rightarrow \quad n_i =\partial_i \beta\,.
\end{equation}
That is, $n_i$ has to be the gradient of a scalar, which leaves no room for vector perturbations. From equation~(\ref{interimbeta}) we get
\begin{equation}
\label{eqbeta}
\partial^2\beta=\frac{1-2\lambda}{1-\lambda} \; \dot{\zeta}\,.
\end{equation}
We now move on to the variation with respect to $\alpha$. This yields
\begin{equation}
-g_2\partial^4 \zeta+2g_{67} \,\partial^4 \alpha-\xi \,\partial^2\zeta -\eta \,\partial^2\alpha=0.
\end{equation}
Again, imposing some regularity conditions, this can be solved to give
\begin{equation}
\label{eqalpha}
\alpha=-\frac{\xi +g_2 \partial^2}{\eta-2 g_{67}\partial^2} \; \zeta\,.
\end{equation}
We can now use equations~(\ref{eqni}), (\ref{eqbeta}) and (\ref{eqalpha}) to integrate out the non-dynamical fields $n_i$ and $\alpha$ in favour of the dynamical field $\zeta$. The quadratic action then reads
\begin{eqnarray}
S=M_{pl}^2\int {\mathrm{d}}^2 x \,{\mathrm{d}} t \,\Big\{\frac{1}{2} \; \frac{1-2\lambda}{1-\lambda}\; \dot{\zeta}^2+\zeta D \zeta\Big\}\,,
\end{eqnarray}
where
\begin{equation}
D\equiv \frac{\xi^2+2\left[2\eta g_1+\xi g_2\right]\partial^2+\left[g_2^2-4 g_1(g_6+g_7)\right]\partial^4}{2(\eta- (g_6+g_7)\partial^2)} \; \partial^2\,.
\end{equation}
The dispersion relation for the scalar is then given by a rational polynomial
\begin{equation}
\label{dispersion}
\frac{1}{2}\;\frac{1-2\lambda}{1-\lambda} \; \omega^2=
\frac{P_1(k)}{P_2(k)} \,,
\end{equation}
where
\begin{eqnarray}
P_1(k)&=&\xi^2 k^2-2\left[2\eta g_1+\xi g_2\right]k^4\nonumber\\&&+\left[g_2^2-4 g_1 (g_{6}+g_{7})\right]k^6\\
P_2(k)&=&2(\eta+(g_{6}+g_{7}) k^2)\,.
\end{eqnarray}
Note that this rational polynomial dispersion relation is qualitatively similar to that of the scalar mode in $3+1$ dimensions \cite{Blas:2009qj}. Classical stability requires this action to have the correct relative sign between the kinetic and the potential term, that is, it requires
\begin{equation}
\frac{1-\lambda}{1-2\lambda}\;\frac{1}{\eta}>0\,.
\end{equation}
Since we have only one excitation, positivity of energy is not really a concern, as it can be controlled by flipping the overall sign of the action. This is not the case in $3+1$ dimensions, where there is a spin-2 graviton as well, and one needs to separately require that its kinetic term has the same sign as the kinetic term of the scalar, if neither of the two fields is to be a ghost.
In the low-momentum limit we have
\begin{equation}
\label{eqalpha2}
\alpha\approx -\frac{\xi}{\eta} \; \zeta\,, \qquad D \approx \frac{\xi^2}{2\eta} \; \partial^2\,.
\end{equation}
That is, $\alpha$ is given algebraically in terms of the conformal factor of the spatial metric $\zeta$, which now satisfies a standard linear dispersion relation
\begin{equation}
\label{dispersion2}
\frac{1}{2}\;\frac{1-2\lambda}{1-\lambda} \; \omega^2 \approx
\frac{\xi^2 k^2}{2\eta}\,,
\end{equation}
with (low momentum) phase velocity
\begin{equation}
c_\zeta=\frac{\xi}{\sqrt{\eta}}\sqrt{\frac{1-\lambda}{1-2\lambda}}\,.
\end{equation}
Clearly, the limit to general relativity is far from trivial. In that limit we have $g_i\to 0$, $\eta\to 0$ and $\lambda\to 1$, so the behaviour of the mode depends strongly on how $\eta/(\lambda-1)$ scales as both numerator and denominator approach zero. Note that in general relativity, where $g_i=0$, $\eta= 0$ and $\lambda= 1$ {\em a priori}, equation~({\ref{interimni}) is enough to render the dynamics of $\zeta$ trivial.
What about the dynamics beyond the quadratic order? One could straightforwardly follow the lines of reference~\cite{Papazoglou:2009fj} (or reference~\cite{Kimpton:2010xi}) and derive the cubic interactions for the scalar. However, we will not attempt this here, simply because the calculation is not sensitive to the dimensionality of space. Thus, there is no reason to believe that the result will differ in any way (at least qualitatively) from that obtained in the case of $3+1$ dimensions. It is, therefore, expected that the scalar will exhibit strong coupling at some scale determined by the magnitudes of $\lambda$ and $\eta$ \cite{Papazoglou:2009fj,Kimpton:2010xi}. It is also expected that choosing the right hierarchy between $M_{pl}$ and $M_\star$ will alleviate this strong coupling, just as in the $3+1$ case \cite{Blas:2009ck}. The only difference here is that, since in 2+1 dimensions there are no experimental constraints but only consistency requirements, one now has the freedom to choose the magnitude of $\lambda$ and $\eta$ so as to push the strong coupling scale beyond the energies at which the higher-order operators become important.
To sum up, the theory in $2+1$ dimensions clearly has a fundamental difference from its counterpart in $3+1$ dimensions, namely the absence of a spin-2 graviton, as one might have expected. However, it does possess a scalar degree of freedom with non-trivial dynamics. Additionally, this degree of freedom exhibits (qualitatively) the same dynamical behaviour as does the scalar degree of freedom in $3+1$ dimensional Ho\v rava--Lifshitz gravity. Therefore, it seems that Ho\v rava--Lifshitz gravity in $2+1$ dimensions is a good theoretical playground for exploring scalar-mode related aspects of Ho\v rava--Lifshitz gravity in $3+1$ dimensions.
\subsection{Equivalence with Einstein-aether theory and covariant formulation}
As one can readily see from the discussion in section \ref{1+1equiv} the equivalence between Ho\v rava--Lifshitz gravity and Einstein-aether theory with a hypersurface orthogonal aether does not really hinge on the dimensionality. In fact, in $2+1$ dimensions it will go through exactly in the same way as in $3+1$ dimensions. However, unlike the special case of $1+1$ dimensions, where the action of Ho\v rava--Lifshitz gravity does not contain any higher-order terms, here this equivalence will be limited to the low-energy limit only, and not hold for the complete action.
Nonetheless, restricted though it may be, this equivalence is suggestive on its own of a suitable extension to all energies. One could follow this series of steps:
\begin{enumerate}
\item
Construct the most general (fully covariant) extension of Einstein-aether theory with up to four derivatives, imposing the additional restriction that the aether be hypersurface orthogonal.
\item
Observe that the part of the action that is second-order in derivatives is equivalent to the low-energy limit of Ho\v rava--Lifshitz gravity, and that if there is a more general equivalence it will necessarily have to work order by order.
\item
Use the prescription from the second-order equivalence to identify how to implement the preferred foliation of Ho\v rava--Lifshitz gravity to this higher-order Einstein-aether theory.
\item
Compare the higher-order terms of the two theories order by order, and identify the parameter matching that will lead to the desired equivalence.
\end{enumerate}
The number of covariant terms which are 4th-order in derivatives in 3-dimensional Einstein-aether theory is quite large and, therefore, we will not attempt to apply the algorithm described above and present the result explicitly (doing so at this stage seems also to be of little practical significance). However, as a point of principle, the existence of this algorithm provides a method of ``covariantization'', alternative to the use of projection operators as suggested in reference~\cite{Germani:2009yt}. More importantly, it allows one to make the following interesting observations.
The number of 4th-order terms in equation~(\ref{2daction}) is far smaller than the number of covariant terms which are 4th-order in derivatives in 3-dimensional Einstein-aether theory (even after hypersurface orthogonality of the aether has been imposed). Indeed, the latter generically contains higher-order time derivatives, whereas the former is carefully constructed not to. This will be reflected in the matching of parameters in step 4: several of the couplings of the 4th-order terms in Einstein-aether theory will have to be exactly tuned to satisfy specific algebraic relations in order for the higher-order time derivatives to cancel out. This will reduce the number of independent couplings to exactly the number of independent couplings of the corresponding Ho\v rava--Lifshitz gravity. Though this seems perfectly feasible, it implies that, reformulated as a covariant theory, Ho\v rava--Lifshitz gravity would seem to be unnaturally fine-tuned, even though this does not really seem to be the case, (at least not according to power-counting renormalizability arguments).
Viewed from a perspective of the corresponding Einstein-aether theory the same observations sound perhaps more striking: When constructing such a theory with terms up to a certain order in derivatives, selectively neglecting to exclude a term, or relating its coefficient with another term's coefficient, would certainly be considered unnatural fine tuning. (Based on this observation, standard Einstein-aether theory includes all possible terms which are second-order in derivatives.) However, the specific fine tuning that leads to the corresponding Ho\v rava--Lifshitz gravity is \emph{not} actually unnatural. Remarkably this is exactly the choice that removes higher-order time derivatives, which would otherwise be very worrisome. (Note that it is not the lack of Lorentz symmetry {\em per se} that make this specific fine tuning natural, as any generic choice of parameters would anyway lead to a Lorentz violating theory.)
Clearly, all of the above observations are not particular to 2 spatial dimensions. Therefore, the whole discussion applies unmodified to the phenomenologically more interesting $3+1$ Ho\v rava--Lifshitz gravity.
\subsection{$2+1$ projectable theory as a limiting case}
We now consider the limit to the $2+1$ dimensional projectable version of the theory, same as we did in section \ref{proj1+1} for $1+1$ dimensions.
In $2+1$ dimensions, starting from eq.~(\ref{2daction}) and setting $N=N(t)$ yields
\begin{equation}
\label{2dactionp}
S_p=\frac{M_{pl}^2}{2}\int {\mathrm{d}}^2 x \, {\mathrm{d}} t \, N\sqrt{g} \left[K^{ij}K_{ij}-\lambda K^2+g_1 \,R^2\right].
\end{equation}
Note that, in addition to all terms involving $a_i$ vanishing, the $g_2 \nabla^2 R$ term in eq.~(\ref{2daction}) can also be neglected in the projectable case, as it becomes a boundary term. Furthermore, in view of the Gauss-Bonnet theorem in 2 space dimensions, the $\xi R$ term is a total divergence which becomes a pure boundary term when $N=N(t)$. Hence, it should also be discarded.
We will refrain from presenting the field equations for the projectable theory here, but they can be straightforwardly derived from eqs.~(\ref{eqlapse}), (\ref{eqshift}) and (\ref{piequation}), by setting $N=N(t)$ and discarding terms proportional to $g_2$ and $\eta$. Some extra care is needed in the case of the equation for the lapse, which will now have to be turned to into a global, instead of a local constraint.
What is perhaps of more interest is to obtain the linearized dispersion relation for the projectable theory as a limiting case. Note that taking the limit in which $\eta\to 0$, $g_i\to 0$ for $i\geq 2$ would not be correct, as it would not actually enforce the constraint $N=N(t)$. Instead, if we view the projectable model as a limit of the non-projectable model, then the right limit is $\eta\to \infty$ (this would force $\alpha\to 0$ in the linearized theory). The dispersion relation then becomes
\begin{equation}
\label{dispersionp}
\frac{1}{2}\;\frac{1-2\lambda}{1-\lambda} \; \omega^2= -2 g_1 k^4\,,
\end{equation}
Remarkably, there is no $k^2$ term in this dispersion relation. This can of course be traced back to the fact that in 2 dimensions the Ricci scalar $R$ is actually a total divergence.
This property is specific to 2 dimensions, making $2+1$ dimensional projectable Ho\v rava--Lifshitz gravity rather special and non-representative. In fact, in $3+1$ dimensions, a $k^2$ term is present in the dispersion relation of both the spin-2 and the scalar mode \cite{Sotiriou:2009xxx}.
\section{Conclusions}
We have studied (non-projectable) Ho\v rava--Lifshitz gravity in $1+1$ and $2+1$ dimensions. The $1+1$ theory includes only second-order operators and is, as might have been expected, dynamically equivalent to 2-dimensional Einstein-aether theory. Given that the dynamics of the latter is well known, we can easily infer that Ho\v rava--Lifshitz gravity in $1+1$ dimensions, though non-trivial, does not have any local degrees of freedom. Therefore, its dynamics (or lack thereof) is qualitatively very different from the $3+1$, theory. In fact, in this regard, $1+1$ Ho\v rava--Lifshitz gravity actually shares features of $2+1$ Einstein gravity.
We then moved on to $2+1$ Ho\v rava--Lifshitz gravity could also be considered very different from the $3+1$ theory as it does not have any spin-2 mode. It does, however, have a propagating scalar mode, so it is far from being dynamically trivial. Additionally, if viewed from a different perspective, it actually bears remarkable similarity with the $3+1$ theory, as the propagating scalar mode has, qualitatively, the same dynamical behaviour (rational polynomial dispersion relation, strong coupling, etc.) as the scalar mode in the $3+1$ theory.
Based on the above, one can easily infer that the study of Ho\v rava--Lifshitz gravity in $1+1$ dimensions cannot provide much insight into the properties of the $3+1$ theory. On the other hand, studying Ho\v rava--Lifshitz gravity in $2+1$ dimensions could provide a simpler and tractable setting for studying the properties of the scalar mode in the $3+1$ theory. Additionally, it could help understand properties of the $3+1$ theory which are not strictly (and solely) dependent on the type of the propagating field, {\em e.g.}~renormalization properties. Clearly, both $1+1$ and $2+1$ theories are interesting in their own right as well.
In addition to the above, we also used the $2+1$ theory and its equivalence with 3-dimensional Einstein-aether theory as an example to discuss a possible covariant formulation. We provided an algorithm for writing Ho\v rava--Lifshitz gravity as a higher-order Einstein-aether theory. Starting from this algorithm, we were able to argue that Ho\v rava--Lifshitz gravity would appear to be unnaturally fine-tuned when seen as a covariant theory, even though this does not seem to be the case according to power-counting renormalizability arguments. Put differently, fine tuning a higher-order Einstein-aether theory to eliminate higher-order time derivatives is remarkably not unnatural, provided that the aforementioned renormalizability arguments are robust beyond power counting.
Lastly, we also briefly considered the projectable version of Ho\v rava--Lifshitz gravity, where the lapse function is assumed to be space-indepedent, as a limiting case of the results we had already obtained. In $1+1$ dimensions, the action simply reduces to that of general relativity but with the projectability constraint. In $2+1$ dimensions, the action is that of general relativity with an extra $R^2$ term and the projectability constraint. There is a propagating scalar degree of freedom satisfying a dispersion relation of the form $\omega^2\propto k^4$. The absence of the $k^2$ is particular to $2+1$ dimensions making the $2+1$ theory non-representative (and Lorentz-violating at all scales).
\section*{Acknowledgments}
TPS and SW were supported by Marie Curie Fellowships. MV was supported by the Marsden Fund administered by the Royal Society of New Zealand. We further acknowledge partial support via a FQXi travel grant.
|
\section{Introduction}
\label{intro} Condensation phenomena emerge in various physical
contexts, to name a few, the well-known Bose-Einstein condensation
(BEC) in dilute atomic gases~\cite{Bose, BEC95, Bradley95}, jamming
in traffic flow~\cite{Evans96, Chowdhury00}, wealth condensation in
macroeconomi\-es~\cite{Burda02}, and condensation in zero-range
process (ZRP, see e.g., recent review~\cite{Evans05} and references
therein).
Since the pioneered research on complex networks~\cite{Watts98, BA,
BA00, Reviews, Reviews02, Reviews03, Books, Books04, Books03}, in
the last decade, condensation phenomena, i.e., condensation of links
(or edges) in complex networks has also been widely
discussed~\cite{Bouchaud00, Krapivsky00, Barabasi01, Burda01,
Godreche01, Bauer02, Berg02, Rodgers02, Doro03a, Doro05, Doro03b,
Farkas04, Noh05a, Noh05b, Ohkubo05a, Ohkubo05b, Ohkubo05c, Ohkubo07,
Godreche05}. In the context of complex networks,
the condensation phase corresponds to the situation that a single
node captures a macroscopic finite fraction of total links/edges. It
has been found that condensation phenomena can occur in both growing
and non-growing complex networks. The condensation phase transitions
occurring in non-growing networks~\cite{Burda01, Bauer02, Berg02,
Doro03a, Doro05, Doro03b, Farkas04} are formally equivalent to that
in balls-in-boxes model~\cite{Bialas97}, and hence has been well
studied to a large extent via methods of equilibrium statistical
mechanics. While for growing complex networks, the appearance of the
condensation phase transitions during the dynamical evolution of the
network are particularly interesting due to its out-of-equilibrium
characteristics.
The tasks in this paper are two folds: first, we merge two important models on
this regard, the growing network with nonlinear preferential attachment (we will
refer to ``nonlinear model''~\cite{Krapivsky00} from now on), and the fitness
model~\cite{fitness} into a unifying framework--the nonlinear fitness model. We
then argue that the condensation phase transition appearing in fitness model and
that in nonlinear model stems from different mechanisms; Second, particularly
interesting, we reveal a novel phase structure in the model and this may increase
our understanding on the non-equilibrium phase transitions in dynamical evolution
of complex networks.
The nonlinear model is defined as follows. At each time step $t$, the newly-added
node created $m$ directed links to ones of the earlier existing nodes with $k$-link,
according to a probability, say, $\Pi$, that is proportional to some ``connection
kernel'' $k^{\gamma}$, $\Pi \propto k^{\gamma}$. Here the exponent $\gamma \geq 0$
reflects the tendency of preferential linking to a popular node and hence controls
the preferential attachment. In Ref.~\cite{Krapivsky00}, P. L. Krapivsky {\it et al.}
had discussed the cases of different choices on exponent $\gamma$, i.e., $\gamma =
1$, $\gamma < 1$ and $\gamma > 1$, for growing complex networks with connection
probability,
\begin{equation}
\Pi^s = \frac{k_i^{\gamma}}{\Sigma_j k_j^{\gamma}}.
\end{equation}
They proved that the number of nodes with $k$ links, $N_k$, follows a power law
distribution in the case that $\gamma$ closes to unity. While in the case of
$\gamma < 1$, the distribution shows a stretched exponential form. For $\gamma > 1$,
that is so called the super-linear case, the model exhibits a condensation phase
transition. Especially when $\gamma > 2$, there exists a limiting situation where
the most connected node links to almost all the other nodes in the network, this
corresponds to a ``winner-takes-all'' phenomenon. In this case, the degree of the
most connected node follows $k_{max} \sim t$.
A similar condensation phase, or so-called ``winner-takes-all'' phase, also appears
in fitness model of the complex networks~\cite{fitness, Barabasi01, Rodgers02}. In
fitness model of growing networks, a {\it fitness} parameter, $\eta_i$, which
represents an internal superiority of the $i$-th node, is introduced and chosen
randomly from some distribution. As a result, the well-known Barab\'asi-Albert
scale-free network~\cite{BA} is generalized to the Bianconi-Barab\'asi (B-B) fitness
model of complex network. One may assign the $i$-th node the fitness $\eta_i$
according to its ``energy level'' $\varepsilon_i$ which satisfy some distribution
$g(\varepsilon)$ through the relation
\begin{equation}
\eta_i = e^{-\beta \varepsilon_i},
\label{etai}
\end{equation}
where $\beta$ can be identified as inverse temperature, i.e., $\beta = 1/T$. In this
growing complex network with fitness, the connection probability $\Pi^{f}$ that a new
node connects one of its $m$ links to an existing node $i$ at each time step $t$ is
defined by
\begin{equation}
\Pi^{f} = \frac{\eta_i k_i}{\sum_j \eta_j k_j},
\label{Pfi}
\end{equation}
where $k_i$ is the degree (the number of links occupied by one node) of node
$i$. One may introduce a partition function $Z_{t}$ as
\begin{equation}
Z_{t} = \sum_{j = 1}^t e^{-\beta \varepsilon_j} k_j(\varepsilon_j, t, t_j).
\end{equation}
This model can be solved in a mean-field approximation and the condensation phase
transition process is described by the ``chemical potential'' $\mu$ formally defined
as follows,
\begin{equation}
e^{-\beta \mu} = \lim_{t \rightarrow \infty} \frac{\overline{Z_{t}}}{mt},
\label{muf}
\end{equation}
where $\overline{Z_{t}}$ is the partition function $Z_t$ averaged over some normalized
distribution $g(\varepsilon)$~\cite{Barabasi01}.
The chemical potential $\mu$ is determined by the self-consistent equation
\begin{equation}
I(\beta, \mu) = \int d \varepsilon g(\varepsilon) n(\varepsilon) = 1 ,
\label{selfc}
\end{equation}
where $n(\varepsilon)$ is the occupation number, i.e., the number of
links attached by the preferential attachment mechanism to nodes with
``energy'' $\varepsilon$. Very interestingly, it was proved~\cite{Barabasi01}
that $n(\varepsilon) = 1/(e^{\beta(\varepsilon - \mu)} - 1)$, this
is nothing but Bose-Einstein (BE) statistics. When $\mu < 0$, the
network is in so-called ``fit-get-rich'' (FGR) phase. While in the
thermodynamic limit $t \rightarrow \infty$ if $I(\beta, 0) < 1$,
i.e., the self-consistent equation (\ref{selfc}) has no solution,
and hence a BEC phase transition occurs. There exists a critical
temperature $T_C = 1/\beta_C$ such that $I(\beta, 0) < 1$ for $T <
T_C$. This condensation phase transition was demonstrated as well by
numerical simulations in Ref.~\cite{Barabasi01}.
Both the fitness model and nonlinear model of the complex network
experience the condensation phase transition during their dynamical
evolutions, two questions are naturally arisen: whether or not the
condensation phase transitions in these models have some common
underlying relationship~\cite{Albert02, Goltsev08}, and the role
that different preferential attachment mechanism in either model
plays during the phase transition of the network. In order to answer
such questions, and as well, to explore the possible phase structure
in out-of-equilibrium evolution in complex networks, we propose the
nonlinear fitness model of complex networks which includes the
nonlinear model and the fitness model as its appropriate limits,
as we will show in the following.
\section{The nonlinear fitness growing network}
\label{sec:1}
It is clear that the preferential attachment mechanism controls the network
topology in the nonlinear network and the fitness parameter $\eta_i$ does in
the network with fitness. To bridge the divergence between the two models and
to give a general description of the phase structure, we employ a new connection
probability, $\Pi^{sf}_i$, as follows,
\begin{equation}
\Pi^{sf} = \frac{\eta_i k^{\gamma}_i}{\sum_j \eta_j k^{\gamma}_j}.
\label{Pi1}
\end{equation}
This new probability includes the original connection probability $\Pi^{s}$ in
the nonlinear model and $\Pi^{f}$ in the fitness model, respectively, as its
proper limits: in the limit $\gamma \rightarrow 1$ one goes back to the
preferential attachment probability of the fitness model, and in the limit
$\eta \rightarrow 1$, one recovers the nonlinear model of growing network.
This attachment mechanism now addresses the dynamical evolution of the network.
In order to identify the condensation phase transition of this nonlinear fitness
model of network, similar to the Bianconi-Barab\'asi method, (see Eqn. (\ref{muf})),
we formally define the chemical potential $\mu$~\cite{Barabasi01},
\begin{equation}
\mu = - \frac{1}{\beta} \lim_{t \rightarrow \infty} \ln \frac{\overline{Z^c_t}}{mt}.
\label{mu}
\end{equation}
where, $\beta$ is again the inverse temperature, $m$ is the number of newly added
links at each time step $t$, and
$\overline{Z^c_t}=\sum_{j = 1}^t e^{-\beta \varepsilon_j} k^\gamma (\varepsilon_j, t, t_j)$
is the partition function in our nonlinear fitness model, $t \rightarrow \infty$ plays
the role of thermodynamic limit.
The corresponding rate equation is,
\begin{equation}
\frac{\partial k_i(\varepsilon_i, t, t_i)}{\partial t} = \frac{e^{-\beta
\varepsilon_i} k_i^{\gamma} (\varepsilon_i, t, t_i)}{\overline{Z^c_t}}.
\end{equation}
Note that in $\gamma \rightarrow 1$ limit, the rate equation of B-B fitness model
is restored and one naturally expects a BEC phase transition at low enough
temperature~\cite{Barabasi01}. However, in general, the chemical potential
$\mu (\gamma, T)$ is a function of both temperature $T$ (or the fitness $\eta$)
and the nonlinear exponent $\gamma$, i.e., the phase structure of this nonlinear
fitness network is controlled by these two parameters, instead only one parameter
in either model ($T$ in the original B-B fitness model or $\gamma$ in the nonlinear
one). These parameters both affect the phase transition during the dynamic evolution
of the network. This fact leads to a more complicated behavior of $\mu$ in current
model than that in the fitness or the nonlinear model alone. Here we apply the method
of rate equation of {\it degree}, used in the B-B fitness model~\cite{Barabasi01},
rather than that of {\it connectivity distribution}, used in the nonlinear
model~\cite{Krapivsky00}. This is due to the fact that the fitness parameter $\eta_i$
is different for each node $i$, i.e., $\eta$ is a local, not a global parameter. This
suggests that the fitness parameter $\eta$ and the exponent parameter $\gamma$ which
respectively controls the preferential attachment mechanism are different in nature --
$\eta_i$ reflects an ``inner'' property of the $i$th node, while $\gamma$ controls the
global evolution of the network. This important difference between the B-B fitness
model and the nonlinear model leads to the different dynamical evolution result of the
complex networks.
\section{Numerical simulations and discussions}
\label{sec:2}
Similar to the original B-B fitness model, we identify the non-condensation-condensation
phase transition by the change of sign of the chemical potential $\mu (\gamma, T)$, i.e.,
when $\mu$ experiences a change from negative value to positive value within some regime,
that change implies the critical point of the corresponding phase transition. In addition,
in principle, the phase transition occurs in the thermodynamic limits $t \rightarrow \infty$.
However, in general situations, taking such thermodynamical limit is not realistic, one has
to resort to the numerical simulations. We numerically compute the chemical potential $\mu$
according to Eqn.(\ref{mu}).
Throughout our simulations we fix the number of links per node (per time step) $m = 2$. For
simplicity, we take energy level distribution $g(\varepsilon) = C \varepsilon$ with
normalization constant $C = 2$, and the total number of time steps $t = 10^3$, average over
$100$ runs. The main numerical simulation results are plotted in a three dimensional (3d)
figure (see Fig.~\ref{fig1}), in which three axes of the figure are exponent $\gamma$,
temperature $T$ and chemical potential $\mu$, respectively. In this ``phase diagram", one
can see that the role that the temperature $T$ or exponent $\gamma$ plays in the formation
of such a phase structure during the dynamical evolution of the network. For instance, for
$\gamma = 1$, when one lowers the temperature, a transition from FGR phase (high temperature
phase) to BEC phase (low temperature phase) occurs~\cite{Barabasi01} as what we expect.
Similarly, if one fixes the temperature, say $T = 2.0$, a phase transition would also occur
when the exponent $\gamma$ is varied (between $\gamma = 0$ and $\gamma = 1.5$ in
Fig.~\ref{fig1}). However, one should keep in mind that the current phase transition is
different with that in nonlinear model because no fitness (temperature) effect was taken into
consideration in the latter model. On the other hand, in both cases, the signal of the phase
transition lies on the change of sign of the chemical potential. Note that we do not take
absolute value of the chemical potential, which is different with Ref.~\cite{Barabasi01}.
We see that in Fig.~\ref{fig1} the chemical potential $\mu$ experiences a change from
positive values to negative ones when the exponent $\gamma$ and/or the temperature $T$ vary
in some regime. Now the whole phase structure is richer than that in the B-B fitness model
and in the nonlinear model, respectively, e.g., the BEC phase and the FGR phase in the B-B
fitness model now are only parts of the new phase diagram in $\gamma = 1$ limit.
\begin{figure}
\includegraphics[width=8cm]{fig1}\\
\caption{(Color online) Three dimensional illustration of the phase structure
in the nonlinear fitness complex network (Top), and its projection on the
$\gamma - T$ plane (Bottom). In top panel, it shows the chemical potential
$\mu$ as a function of temperature $T$ and the nonlinear exponent $\gamma$,
as defined in Eqn. (\ref{mu}). In this simulation we take the number of links
per node (per time step) $m = 2$, and energy level distribution
$g(\varepsilon) = 2 \varepsilon$ (normalized), and the total number of time
steps $t = 10^3$, average over $100$ runs. \label{fig1}}
\end{figure}
The situation is more clear when we plot the chemical potential $\mu$ as a
function of exponent $\gamma$ in Fig.~\ref{fig2}, and of temperature $T$ in
Fig.~\ref{fig3}, respectively. These figures display the different ``cross
section" views of 3d ``phase diagram" of Fig.~\ref{fig1}. Some new features of
the phase transition can be read out from these plots. In Fig.~\ref{fig2}, we
show the chemical potential $\mu$ versus the exponent $\gamma$ with temperature
$T$ fixed. It can be seen from the figure that as temperature increases, the
critical value of exponent for phase transition $\gamma_C$ decreases, e.g., for
$T = 0.2$, $\gamma_C \sim 1.2$, while for $T = 5.0$, $\gamma_C \sim 0.6$, roughly.
This tells us that the transition of phase occurs with weaker tendency of
preferential linking at relatively higher temperature.
\begin{figure}
\includegraphics[width=8cm]{fig2}\\
\caption{(Color online) The chemical potential $\mu (\gamma, T)$ as a function of
the exponent $\gamma$ but with temperature fixed, in a linear-linear
scale. The four curves shown in the panel correspond to $T = 0.2$
(open triangles, solid line), $T = 0.8$ (open circles, dashed line),
$T = 2.0$ (open squares, dotted line), $T = 5.0$ (open diamonds,
dot-dashed line), respectively. All symbols are connected by lines
for eye guidance. \label{fig2}}
\end{figure}
\begin{figure}
\includegraphics[width=8cm]{fig3}\\
\caption{(Color online) Similar to Fig.~\ref{fig2}, the chemical potential $\mu
(\gamma, T)$ as a function of temperature but with exponent $\gamma$
fixed, in a linear-log scale. The four curves shown in the panel
correspond to $\gamma = 0.3$ (open triangles, solid line), $\gamma =
0.7$ (open circles, dashed line), $\gamma = 1.0$ (open squares,
dotted line), $\gamma = 2.0$ (open diamonds, dot-dashed line),
respectively. Note that for $\gamma = 0.3$, the values of $\mu$-s
are all positive, while for $\gamma = 2.0$, the values of $\mu$-s
are all negative. \label{fig3}}
\end{figure}
In Fig.~\ref{fig3} we plot the chemical potential $\mu$ versus the temperature
$T$ with exponent $\gamma$ fixed in a linear-log scale. For data points with
$\gamma = 0.7$ (open circles, dashed line) and $\gamma = 1$ (open squares, dotted
line), the condensation phase transition is very similar to what happens in
Fig.~\ref{fig2}: the critical temperature $T_C$ of the phase transition decreases
as the exponent $\gamma$ increases, e.g., for $\gamma = 0.7$, $T_C \sim 3.0$, while
for $\gamma = 1$, $T_C \sim 0.8$. The latter case recovers the results of B-B BEC
phase transition, as it should. Fig.~\ref{fig3} also displays some new, interesting
aspects: the values of the chemical potential with $\gamma = 0.3$ (open triangles,
solid line) are all positive and those with $\gamma = 2$ are all negative. This
implies that the phase transition disappears eventually as $\gamma$ reaches some
specific critical value, and hence results in a single phase structure -- either a
FGR phase (with negative chemical potential), or a condensation phase (with positive
chemical potential). However, one can not always use this as an identification of
transition between the condensation phase and the non-condensation phase, this
point is illustrated in Fig.~\ref{fig4}.
\begin{figure}
\includegraphics[width=8cm]{fig4}\\
\caption{(Color online) The occupation ratios of the most connected node,
$k_{max}/mt$, plotted as a function of temperature $T$, and of
different nonlinear exponent $\gamma$. The five curves in the panel,
from bottom to top, represent $\gamma = 0.3$ (open triangles, solid
line), $\gamma = 0.7$ (open circles, dashed line), $\gamma = 1$
(open squares, dotted line), $\gamma = 1.3$ (stars, solid line),
$\gamma = 2$ (open diamonds, dot-dashed line), respectively. Note
that the data points with $\gamma = 1$ are consistent with B-B's
results (see the corresponding curve of Fig.~\ref{fig3}. in
Ref.~\cite{Barabasi01}). \label{fig4}}
\end{figure}
We plot the occupation ratio, $k_{max}/mt$, of {\it the most connected} node, in
Fig.~\ref{fig4}. At the first sight from Fig.~\ref{fig4}, it seems that the network
is in a condensation phase at low enough temperature $T < T_C$ ($T_C$ is the critical
temperature), and in a non-condensation phase at high temperature, $T > T_C$. But note
that as the exponent $\gamma$ increases, the condensation phase appears in even
higher temperature region. Especially when $\gamma = 2$, the network enters a
condensation phase regardless of the value of temperature. This supports the
conclusion of nonlinear growing network in Ref.~\cite{Krapivsky00}. However, in addition
to the appearance of the condensation phase in original B-B fitness model and the
nonlinear model with $\gamma > 1$, we conclude from Fig.~\ref{fig4} that they are still
{\it different} phases because the chemical potential is positive ($\mu > 0$) for the
B-B fitness model at $T < T_C$, while negative ($\mu < 0$) for the nonlinear model. The
competitive relationship between $e^{-\beta \varepsilon_i}$ and $k^{\gamma}$ in partition
function play important roles in the dynamical evolution of corresponding complex networks
and hence the formation of such a novel phase structure. When the former factor
$e^{-\beta \varepsilon_i}$ is dominant, the network experiences a condensation phase
transition if $T < T_C$, and its chemical potential is positive, $\mu > 0$; while the
prevailing role of the factor $k^{\gamma}$ leads to the opposite side: the chemical
potential of the condensation phase is always negative, $\mu < 0$. It is also clear that,
from above figure, for $\mu > 0$ and $\gamma = 1$, the network with fitness preferential
attachment is condensed on the node with the {\it lowest} energy level (or the fittest
node) only if $T < T_C$ in the thermodynamic limit; for $\mu < 0$, the network experiences
a ``winner takes all'' phenomenon when $\gamma$ is large enough (e.g., $\gamma > 1.7$).
Due to aforementioned competitive relationship between two factors $e^{-\beta \varepsilon_i}$
and $k^{\gamma}$, the chemical potential of the non-condensation phase is changed, for
instance, see the curve with $\gamma = 0.3$ in Fig.~\ref{fig3}, though one has an overall
positive $\mu$ at high temperature, no condensation phase appears. The different phase
structures identified by the chemical potential $\mu$ stem from different dominant factors
of preferential attachment mechanism, they are distinct in nature.
In addition, as is well known, a significantly different manifestation of BEC
and FGR phases is that the fraction of the total number of links connected to the most
connected node tends to zero in thermodynamic limit for the latter and tends to a finite
value for the former. In current combined model, Fig.~\ref{fig4} shows that there exists a
similar situation for $\mu>0$ when the factor $\eta$ dominates the attachment mechanism.
while when the factor $\gamma$ dominates the attachment, the ratio tends to zero in
thermodynamic limit for $\mu<0$, which displays a FGR phase. However, note that if $\gamma$
is large enough, this occupation ratio can also approach to a finite value and a
condensation phase appears even for $\mu<0$.
A major difference between the condensations of fitness model and non-linear model
is that the condensation in the former occurs on the node with the lowest energy level (or
the highest fitness), instead in the latter model it only occurs on the first node of the
network. However, topologically it is hardly able to tell whether or not a graph has reached
condensation due to the non-linear preferential attachment or the BEC due to the fitness,
since in both condensation phases, the important topological parameter, i.e., the cluster
coefficient (CC) approaches to the same limit. If one trace the node with the lowest energy
level, then this node is more likely to be the {\it most connected} node in low temperature
region, since with large parameter $\beta=1/T$, the fitness parameter $\eta$ dominates the
dynamical evolution in combined attachment mechanism. As temperature increases, the
probability of the node with the lowest energy level being the most connected node is getting
smaller due to the impact of fitness $\eta$, as a result, in high temperature region, the
factor $k^{\gamma}$ dominates the attachment. This is shown in Fig.~\ref{fig5}, in which the
occupation ratios of the node with the lowest energy level, $k_{\varepsilon}/mt$, is plotted;
as well as that of the most connected node, $k_{max}/mt$, as function of temperature.
The former exhibits evidently larger fluctuations compared to the occupation ratio
of the most connected node $k_{max}/mt$, since the node with the lowest energy level is not
necessary the most connected node. As a result, in high temperature region, in which
$k^{\gamma}$ dominates the attachment ($\eta \rightarrow 1$), fluctuations are relatively
small, since as a global parameter, $\gamma$ does not depend on the local fitness
of a node, such that the earliest presented node could be the most connected node for large
$\gamma$, regardless the local fitness for $T>T_C$ ($T_C$ is the critical temperature); in
low temperature region, ($T<T_C$), fluctuations of the occupation ratio of a node with the
lowest energy level suggests that the emergence of condensation comes from the competition
of two sides -- the local fittest node (with fitness attachment) and the earliest presented
node (with super-linear preferential attachment), and this competition between makes
fluctuations larger and larger (e.g., see $\gamma = 2$ in Fig.~\ref{fig5}). This demonstrates
how both attachment factors $\eta$ and $\gamma$ control the evolution of the network
simultaneously.
\begin{figure}
\includegraphics[width=8cm]{fig5}\\
\caption{(Color online) The occupation ratios of the most connected node,
$k_{max}/mt$, and that of the node with the lowest energy level,
$k_{\varepsilon}/mt$, plotted as a function of temperature $T$, and
of different nonlinear exponent $\gamma$'s. Data with hollow symbols
and solid lines correspond to $k_{max}/mt$, represent $\gamma = 0.7$
(open circles), $\gamma = 1$ (open squares), $\gamma = 2$ (open
diamonds), respectively. Data with solid symbols and non-solid lines
correspond to $k_{\varepsilon}/mt$, represent $\gamma = 0.7$ (solid
circles, dashed line), $\gamma = 1$ (solid squares, dot-dashed
line), $\gamma = 2$ (solid diamonds, dotted line), respectively.
\label{fig5}}
\end{figure}
\section{Conclusions}
\label{sec:3}
In conclusion, in this paper we analyze the condensation phase transitions in a unifying
framework which includes both the nonlinear model and the fitness model as its appropriate
limit. The goodness of this new framework is to, on one hand, bridge the differences
between the nonlinear model and the fitness model; On the other hand, allow us to identify
the different roles played by different factors in current model. Under appropriate limits
(e.g., $\gamma \rightarrow 1$ or $\eta \rightarrow 1$), the typical phase structures of the
original B-B fitness model and the nonlinear model are recovered respectively.
Through the numerical simulations of this nonlinear fitness model, we show a novel 3d phase
diagram. We employ the Bianconi--Barab\`asi method and numerically compute the chemical
potential to identify the critical points ($T_C$ and $\gamma_C$) of the phase transitions
of the networks. Our results and analysis directly answer the questions that asked in the
beginning, i.e., the condensation phase transitions in both the B-B fitness model and the
nonlinear model are in fact distinct in nature. As well, the preferential attachment
mechanism in our nonlinear fitness model depends on two factors: $k^{\gamma}$ and
$e^{-\beta \varepsilon_i}$ (or $\eta_i$). Both factors affect the phase structure of the
network and hence the network topologies during the evolution of the network. We reveal
that the competitive relationship between these two factors leads to the condensation phase
transitions, and hence transitions between different network topologies. We hope the current
study may increase our understanding to the non-equilibrium critical phenomena, particularly
the condensation phase transitions in the evolution of complex networks.
\acknowledgments We acknowledge to the Initiative Plan of Shanghai Education Committee
(Project No. 10YZ76) and the Scientific Research Foundation for the Returned Overseas
Chinese Scholars, State Education Ministry (SRF for ROCS, SEM) for their support.
|
\section{Introduction}
Hidden shift problems have been studied in quantum computing as they
provide a framework that can give rise to new quantum algorithms. The
hidden shift problem was first introduced and studied in a paper by
van Dam, Hallgren and Ip \cite{vDHI:2003} and is defined as follows.
We are given two functions $f$, $g$ that map a finite group $G$ to
some set with the additional promise that there exists an element
$s\in G$, the so-called shift, such that for all $x$ it holds that
$g(x) = f(x+s)$. The task is to find $s$. Here the group $G$ is
additively denoted, but the problem can be defined for non-abelian
groups as well. The great flexibility in the definition allows to
capture interesting problems ranging from algebraic problems such as
the shifted Legendre symbol \cite{vDHI:2003}, over geometric problems
such as finding the center of shifted spheres \cite{CSV:2007,Liu:2009}
and shifted lattices \cite{Regev:2004}, to combinatorial problems such
as graph isomorphism \cite{CW:2007}.
Notable here is a well-known connection between the hidden subgroup
problem for the dihedral group, a notoriously difficult instance which
itself has connections to lattice problems and average case subset sum
\cite{Regev:2004} and a hidden shift problem over the cyclic group
$\mathbb{Z}_n$ where the functions $f$ and $g$ are injective
\cite{Kuperberg:2005,MRRS:2007,CvD:2007}. It is known \cite{FIMSS:2003,Kuperberg:2005} that the hidden
shift problem for {\em injective} functions $f,g: G \rightarrow S$
that map from an abelian $G$ to a set $S$ is equivalent to hidden subgroup
problem over the semi-direct product between $G$ and $\mathbb{Z}_2$, where the action of $\mathbb{Z}_2$ on $G$ is
given by the inverse. We would like to point out that the functions studied here are Boolean functions (i.e., $G=\mathbb{Z}_2^n$) and therefore
far from being injective. Even turning them into injective {\em quantum}
functions, as is possible for bent functions \cite{Roetteler:2010},
seems not to be obvious in this case.
Another recent example of a
non-abelian hidden shift problem arises in a reduction used to argue
that the McEliece cryptosystems withstands certain types of quantum
attacks \cite{DMS:2010}.
In this paper we confine ourselves to the abelian case and in
particular to the case where $G=\mathbb{Z}_2^n$ is the Boolean hypercube. The
resulting hidden shift problem for Boolean functions, i.e., functions
that take $n$ bits as inputs and output just $1$ bit, at first glance
looks rather innocent. However, to our knowledge, the Boolean case was
previously only addressed for two extreme cases: a) functions which
mark precisely one element and b) functions which are maximally apart
from any affine Boolean function (so-called bent functions). In case
a), the problem of finding the shift is the same as unstructured
search, so that the hidden shift can be found by Grover's algorithm
\cite{Grover:96} and the query complexity is known to be tight and is
given by $\Theta(\sqrt{2^n})$.
In case b) the hidden shift can be discovered in one query using an
algorithm that was found by one of the co-authors
\cite{Roetteler:2010}, provided that the {\em dual} of the function
can be computed efficiently, where the definition of the dual is via
the Fourier spectrum of the function which in this case can be shown
to be flat in absolute value. If no efficient implementation of the
dual is known then still a quantum algorithm exists that can identify
the hidden shift in $O(n)$ queries. The present paper can be thought
of as a generalization of this latter algorithm to the case of Boolean
functions other than those having a flat spectrum. This is motivated
by the quite natural question of what happens when the extremal
conditions leading to the family of bent functions are relaxed. In this paper we
address the question of whether there is a broader class of functions for
which hidden shifts of a function can be identified.
The first obvious step in direction of a generalization is actually a
roadblock: Grover's search problem~\cite{Grover:96} can also be cast
as a hidden shift problem. In this case the corresponding class of
Boolean functions are the \emph{delta functions}, i.e., $f, g:
\{0,1\}^n \rightarrow \{0,1\}$, where $g(x)=f(x+s)$ and $f(x)$ is the
function that takes value $1$ on input $(0, \ldots, 0)$ and $0$
elsewhere and $g(x)$ is the function that takes the value $1$ on input
$s$ and $0$ elsewhere. Grover's algorithm \cite{Grover:96} allows to
find $s$ in time $O(\sqrt{2^n})$ on a quantum computer (which is also
the fastest possible \cite{BV:97}).
Thus, the following situation emerges for the quantum and the classical query complexities of these two extremal cases: for bent functions the classical query complexity\footnote{Note that the query complexity depends crucially on how the functions $f$ and $g$ can be accessed: the stated bounds hold for the case where $f$ and $g$ are given as black-boxes. If $f$ is a {\em known} bent function, then it is easy to see that the classical query complexity becomes $O(n)$.} is $\Omega(\sqrt{2^n})$ and the quantum query complexity\footnote{A further improvement is possible in case the so-called {\em dual} bent function
$\widetilde{f}$ is accessible via another black-box: in this case
the quantum query complexity becomes constant
\cite{Roetteler:2010}.} is $O(n)$. For delta functions the classical query
complexity is $\Theta(2^n)$ and the quantum query complexity is
$\Theta(\sqrt{2^n})$.
For a general Boolean function the hidden shift problem can be seen as
lying somewhere between these two extreme cases. This is somewhat
similar to how the so-called weighing matrix
problem~\cite{WeighingMatrices} interpolates between the
Bernstein-Vazirani problem~\cite{BV:97} and Grover search, and how the
generalized hidden shift problem \cite{CvD:2007} interpolates between
the abelian and dihedral hidden subgroup problems. However, apart from
these two extremes, not much is known about the query complexity of
the hidden shift problem for general Boolean functions.
The main goal of this work was to understand the space between these
two extremes. We show that there is a natural way to ``interpolate''
between them and to give an algorithm for each Boolean function whose
query complexity depends only on properties of the Fourier spectrum of
that function.
\bigskip
\paragraph{\bf Prior work.} As far as hidden shifts of Boolean
functions are concerned, besides the mentioned papers about the bent
case and the case of search, very little was known. The main technique
previously used to tackle hidden shift problem was by computing a
suitable convolution. However, in order to maintain unitarity, much of
target function's features that we want to compute the convolution
with had to be ``sacrificed'' by requiring the function to become
diagonal unitary, leading to a renormalization of the diagonal
elements, an issue perhaps first pointed out by \cite{CurtisMeyer}.
No such renormalization is necessary if the spectrum is already flat
which corresponds to the case of the Legendre symbol \cite{vDHI:2003}
(with the exception of one special value at 0) and the case of
bent functions which was considered in \cite{Roetteler:2010}.
\paragraph{\bf Our results.} We introduce a quantum algorithm that allows
us to sample from vectors that are perpendicular to the hidden shift
$v$ according to a distribution that is related to the Fourier
spectrum of the given Boolean function $f$. If $f$ is bent, then this
distribution is uniform which in turn leads to a unique
characterization of $v$ from $O(n)$ queries via a system of linear
equations. For general $f$ more queries might be necessary and
intuitively the more concentrated the Fourier spectrum of $f$ is, the
more queries have to be made: in the extreme case of a ($\pm 1$
valued) delta function $f$ the spectrum is extremely imbalanced and
concentrated almost entirely on the zero Fourier coefficient which
corresponds to the case of unstructured search for which our algorithm
offers no advantage over Grover's algorithm. For general $f$ we give
an upper bound on the number of queries in terms of the {\em
influence} $\gamma_{f}$ of the function $f$, where the influence is defined
as $\gamma_{f} = \min_{v} (\PR[x]{f(x)\neq f(x+v)})$.
From a simple application of the Chernoff bound it follows that it is
extremely unlikely that a randomly chosen Boolean function will give
rise to a hard instance for our quantum algorithm. This in turn gives
rise to our main result of the paper:
\bigskip
\noindent
{\bf Theorem 2 (Average case exponential separation).} {\em Let $({\cal O}_f,
{\cal O}_g)$ be an instance of a Boolean hidden shift problem (BHSP)
where $g(x)=f(x+v)$ and $f$ and $v$ are chosen uniformly at random.
Then there exists a quantum algorithm which finds $v$ with bounded error using $O(n)$
queries and in $O(\mathrm{poly}(n))$ time whereas any classical algorithm needs
$\Omega(2^{n/2})$ queries to achieve the same task.}
\smallskip
This result can be interpreted as an exponential quantum-classical
separation for the time and query complexity of an average case problem. Finally,
we would like to comment on the relationship between the problem
considered in this paper and the abelian hidden subgroup problem. It
is interesting to note, yet not particularly difficult to see, that the
case of a hidden shift problem for bent functions can be reduced to
that of an abelian hidden subgroup problem. The hiding function in
this case is a quantum function, i.\,e., it takes values in the set
of quantum sets rather than just basis states. For the case of a
non-bent function, including the cases of random functions considered
here, the same direct correspondence to the hidden subgroup problem over
an abelian group no longer exists, i.\,e., even though there is no obvious
group/subgroup structure present in the function $f$, the algorithm
can still identify the hidden shift $v$.
\section{Preliminaries}
\begin{definition}[Boolean Hidden Shift Problem]\label{d_BHSP}
Let $n \geq 1$ and let $f,g : \mathbb{Z}_2^n\to\mathbb{Z}_2$ be two Boolean functions such that the following conditions hold:
\begin{itemize}
\item if for some $t\in \mathbb{Z}_2^n$ it holds that $f(x)\equiv f(x+t)$ then $t=0$;
\item for some $s\in \mathbb{Z}_2^n$ it holds that $g(x)\equiv f(x+s)$.
\end{itemize}
If $f$ and $g$ are given by two oracles $O_f$ and $O_g$, we say that
the pair $(O_f,O_g)$ defines an instance of a hidden shift problem
(BHSP) for the function $f$. The value $s\in\mathbb{Z}_2^n$ that satisfies
$g(x)\equiv f(x+s)$ is the solution of the given instance of the
BHSP.
\end{definition}
We also consider the $\{+1,-1\}$-valued function $F$ corresponding to
the function $f$ and view it as a function over $\mathbb{R}$, that is,
\begin{align}
F:\mathbb{Z}_2^n\to\mathbb{R}:x\mapsto (-1)^{f(x)}.
\end{align}
The arguments of these functions are assumed to belong to $\mathbb{Z}_2^n$,
and their \emph{inner product} is defined accordingly, i.e., $
\ip{u}{v}=\bigoplus_{i=1}^n u_i\cdot v_i. $ We also denote by
$\chi_u(\dots)$ the elements of the standard Fourier basis
corresponding to $\mathbb{Z}_2^n$, that is, $\chi_u(v)=(-1)^{\ip uv}$ for
every $u,v\in\mathbb{Z}_2^n$.
We will see that the complexity of the BHSP\ depends on the notion of
influence.
\begin{definition}[Influence]
For any Boolean function $f$ over $\mathbb{Z}_2^n$ and $n$-bit string $v$,
we call $\gamma_{f,v}=\PR[x]{f(x)\neq f(x+v)}$ the influence of $v$ over $f$,
and $\gamma_{f}=\min_{v}\gamma_{f,v}$ the minimum influence over $f$.
\end{definition}
The following lemma relates the influence over a Boolean function $f$
to the Fourier spectrum of its $\{+1,-1\}$-valued analog $F$, see
also \cite[Fact 11, p. 14]{Gopalan2009}.
\begin{lemma}\label{lem:influence}
$\gamma_{f,v}=\sum_{u:\ip vu=1} \abs{\widehat F(u)}^2.$
\end{lemma}
We give a proof of this lemma in Appendix \ref{app:influence} for completeness.
\section{Our algorithm}
\begin{theorem}\label{thm:complexity}
There exists a quantum algorithm that solves an instance of BHSP\
defined over the function $f$ using expected $O(n/\sqrt{\gamma_{f}})$
oracle queries. The algorithm takes expected time polynomial in the
number of queries.
\end{theorem}
\begin{figure}
\centerline{
\unitlength0.75pt%
\begin{picture}(10,50)(0,0)
\put(-25,0){\makebox(0,0)[l]{$\ket{0}^{\otimes n}$}}
\put(-25,40){\makebox(0,0)[l]{$\ket{0}$}}
\end{picture}%
\begin{picture}(16,50)(0,0)
\put(0,0){\line(1,0){16}}
\put(6,-7){\line(1,3){5}}
\put(0,40){\line(1,0){16}}
\end{picture}%
\begin{picture}(30,50)(0,0)
\put(0,-15){\framebox(30,30){$H^{\otimes n}$}}
\put(0,40){\line(1,0){30}}
\end{picture}%
\begin{picture}(10,50)(0,0)
\multiput(0,0)(0,40){2}{\line(1,0){10}}
\end{picture}%
\begin{picture}(30,50)(0,0)
\put(0,-15){\framebox(30,70){$O_f$}}
\end{picture}%
\begin{picture}(10,50)(0,0)
\put(0,0){\line(1,0){10}}
\put(0,40){\line(1,0){10}}
\end{picture}%
\begin{picture}(30,50)(0,0)
\put(0,25){\framebox(30,30){$Z$}}
\put(0,0){\line(1,0){30}}
\end{picture}%
\begin{picture}(10,50)(0,0)
\put(0,0){\line(1,0){10}}
\put(0,40){\line(1,0){10}}
\end{picture}%
\begin{picture}(30,50)(0,0)
\put(0,-15){\framebox(30,70){$O_g$}}
\end{picture}%
\begin{picture}(10,50)(0,0)
\multiput(0,0)(0,40){2}{\line(1,0){10}}
\end{picture}%
\begin{picture}(30,50)(0,0)
\put(0,-15){\framebox(30,30){$H^{\otimes n}$}}
\put(0,40){\line(1,0){30}}
\end{picture}%
\begin{picture}(16,50)(0,0)
\put(0,0){\line(1,0){16}}
\put(6,-7){\line(1,3){5}}
\put(0,40){\line(1,0){16}}
\end{picture}%
\begin{picture}(20,50)(0,0)
\put(5,20){\makebox(0,0)[l]{$\left. \rule{0mm}{1.0cm} \right\}$}}
\put(13,20){\makebox(0,0)[l]{{ measure}}}
\end{picture}%
}
\vspace{2mm}
\caption{Quantum circuit for the {\bf Sampling Subroutine}.\label{fig:sampling}}
\end{figure}
\begin{proof}
The algorithm relies on the {\bf Sampling Subroutine} described in
Fig.~\ref{fig:sampling}, where $H$ denotes the standard Hadamard
gate, $Z$ is a phase gate acting on one qubit as
$Z:\ket{b}\mapsto (-1)^b\ket{b}$, and $O_f$ is the oracle for
$f$ acting on $n+1$ qubits as $O_f:\ket{b}\ket{x}\mapsto\ket{b\oplus
f(x)}\ket{x}$ (similarly for $O_g$). The algorithm works as
follows:
\algorithm{ {\bf Quantum algorithm}
\begin{enumerate}
\item Set $i=1$
\item\label{step:sampling} Run the {\bf Sampling Subroutine}. Denote
by $(b_i,u_i)$ the output of the measurement.
\item If $\mathsf{Span}\{u_k|k\in{[i]}\}\neq\mathbb{Z}_2^n$, increment $i\rightarrow
i{+}1$ and go back to step~\ref{step:sampling}. Otherwise set $t=i$
and continue.
\item\label{step:system} Output ``$s$'', where $s$ is the unique
solution of
\begin{align*}
\begin{cases}
\ip{u_1}s=b_1;\\
\quad\dots\\
\ip{u_t}s=b_t.
\end{cases}
\end{align*}
\end{enumerate}
}
Obviously, this algorithm makes $O(t)$ quantum queries to the oracles
and its complexity is polynomial in $t+n$. The quantum state before the measurement is
\begin{align}
\ket{0}\ket{0}^{\otimes n}
&\stackrel{H^{\otimes n}}{\longmapsto}\frac{1}{\sqrt{2^n}}\sum_x\ket{0}\ket{x}
\stackrel{O_f}{\longmapsto} \frac{1}{\sqrt{2^n}}\sum_x\ket{f(x)}\ket{x}
\stackrel{Z}{\longmapsto}\frac{1}{\sqrt{2^n}}\sum_x(-1)^{f(x)}\ket{f(x)}\ket{x}\nonumber\\
&\stackrel{O_g}{\longmapsto}\frac{1}{\sqrt{2^n}}\sum_x(-1)^{f(x)}\ket{f(x)\oplus g(x)}\ket{x}\nonumber\\
&\phantom{\longmapsto}=\frac{1}{\sqrt{2^n}}\ket{0}\sum_x\frac{F(x)+F(x+s)}{2}\ket{x}+\frac{1}{\sqrt{2^n}}\ket{1}\sum_x\frac{F(x)-F(x+s)}{2}\ket{x}\nonumber\\
&\stackrel{H^{\otimes n}}{\longmapsto}
\ket{0}\sum_u\frac{1+\chi_u(s)}{2}\widehat F(u)\ket{u}
+\ket{1}\sum_u\frac{1-\chi_u(s)}{2}\widehat F(u)\ket{u}.\label{m_D}
\end{align}
Its measurement therefore always returns a pair
$(b_i,u_i)\in\{0,1\}\times\{0,1\}^n$ where $\ip{u_i}s=b_i$. Moreover,
since by construction $\mathsf{Span}\{u_i|i\in{[t]}\}=\mathbb{Z}_2^n$, the system of
equations in step~\ref{step:system} accepts a unique solution that can
only be the hidden shift $s$, thus the final answer of our algorithm
is always correct.
We now show that the algorithm terminates in bounded expected time. We
need to prove that repeatedly sampling using the procedure in
step~\ref{step:sampling} yields $n$ linearly independent vectors
$u_i$, therefore spanning $\mathbb{Z}_2^n$, after a bounded expected number
of trials $t$. Let $(B,U)$ be a pair of random variables describing the measurement outcomes for the {\bf Sampling Subroutine}, and $D_f^U$ denote the marginal distribution of $U$. From the right-hand side of~(\ref{m_D}) it is clear that
\begin{align*}
D_f^U(u)\equiv\abs{\widehat F(u)}^2.
\end{align*}
Note that this distribution does not depend on $g$.
Let $d_i$ be the dimension of $\mathsf{Span}\{u_k|k\in{[i]}\}$. By
construction, we have $d_1=1,d_t=n$ and $d_{i+1}$ equals either $d_i$
or $d_{i}+1$. Let us bound the probability that $d_{i+1}=d_{i}+1$, or,
equivalently, that $u_{i+1}\notin\mathsf{Span}\{u_k|k\in{[i]}\}$. This
probability can only decrease as $d_{i}$ increases, so let us consider
the worst case where $d_{i}=n-1$. In that case, there exists some
$v\in\mathbb{Z}_2^n\setminus\{0\}$ such that $\mathsf{Span}\{u_k|k\in{[i]}\}$ is
exactly the subspace orthogonal to $v$. Then, the probability that
$u_{i+1}$ distributed according to $D_f^U$ does not lie in this subspace
(and hence $d_{i+1}=d_i+1$) is given by
\begin{align*}
\PR[u\simD_f^U]{\ip vu=1}=\sum_{u:\ip vu=1} \abs{\widehat F(u)}^2=\gamma_{f,v},
\end{align*}
which follows from Lemma~\ref{lem:influence}. Therefore, for any $i$,
the probability that $d_{i+1}=d_i+1$ is at least $ \gamma_{f}=\min_{v}{\gamma_{f,v}},
$ and the expected number of trials before it happens is at most
$1/\gamma_{f}$. Since $d_i$ must be incremented $n$ times, the expected total
number of trials $t$ is at most $n/\gamma_{f}$.
Using quantum amplitude amplification, we can obtain a quadratic
improvement over this expected running time. Indeed, instead of
repeating the {\bf Sampling Subroutine} $O(1/\gamma_{f})$ times until we
obtain a sample $u$ not in the subspace spanned by the previous
samples, we can use quantum amplitude amplification, which achieves
the same goal using only $O(1/\sqrt{\gamma_{f}})$ applications of the quantum
circuit in the {\bf Sampling Subroutine} (see~\cite[Theorem~3]{BrassardHMT02}). We therefore obtain a quantum algorithm that
solves the problem with success probability 1 and an expected number
of queries $O(n/\sqrt{\gamma_{f}})$. \hfill $\Box$
\end{proof}
In case a lower bound on $\gamma_{f}$ is known, we have the following corollary:
\begin{corollary}\label{cor:complexity-promise}
There exists a quantum algorithm that solves an instance of BHSP\
defined over the function $f$, with the promise that
$\gamma_{f}\geq\delta$, with success probability at least $1-\varepsilon$
and using at most $O(n\log(1/\varepsilon)/\sqrt{\delta})$ oracle
queries. The algorithm takes expected time polynomial in the number
of queries.
\end{corollary}
\begin{proof}
This immediately follows from Markov's inequality, since it implies
that the algorithm in Theorem~\ref{thm:complexity} will still
succeed with constant probability even when we stop after a time
$\Theta(n/\sqrt{\gamma_{f}})$ if it has not succeeded so far. \hfill $\Box$
\end{proof}
\section{Classical complexity of random instances of BHSP}
In this section we show that a uniformly chosen instance of BHSP\ is
exponentially-hard classically with high probability.
\begin{lemma}\label{l_cla_h}
A classical algorithm solving a uniformly random instance of BHSP\
with probability at least $1/2$ makes $\Omega(2^{n/2})$ oracle
queries.
\end{lemma}
\begin{proof}
Consider a classical algorithm $\mathcal{A}_{\mathrm{cla}}$ that makes $t_{\mathrm{cla}}$ queries to the
oracles $A_f$ and $A_g$ and with probability at least $1/2$ returns
the unique $s$ satisfying $g(x)\equiv f(x+s)$
(cf.~Definition~\ref{d_BHSP}). For notational convenience we assume
that $\mathcal{A}_{\mathrm{cla}}$ only makes duplicated queries $\left(f(x),g(x)\right)$.
This can at most double the total number of oracle calls.
Consider the uniform distribution of $f:\mathbb{Z}_2^n\to\mathbb{Z}_2$ and
$s\in\mathbb{Z}_2^n$, and let an input instance of BHSP\ be chosen
accordingly. Let $\left(X_1,\dots, X_{t_{\mathrm{cla}}}\right)$ be random
variables representing the queries made by $\mathcal{A}_{\mathrm{cla}}$. Then by the
correctness assumption, the values $f(X_1),g(X_1),\dots,
f(X_{t_{\mathrm{cla}}}),g(X_{t_{\mathrm{cla}}})$ can be used to predict $s$ with probability
at least $1/2$.
First we observe that if, after $k$ queries, it holds that
$X_i-X_j\neq s$ for every $i,j\in[k]$, then even conditionally on
the values of $f(X_1),g(X_1),\dots, f(X_{k}),g(X_{k})$ every
$s\notin\{X_i-X_j|i,j\in[k]\}$ has exactly the same probability to
occur. More precisely, if $S_k=\{X_i-X_j|i,j\in[k]\}$ and $E_k$ is
the event that $s\in S_k$, we have
\begin{align}
\PR{s=s_0|\neg E_k}=\frac{1}{2^n-\sz{S_k}}
\leq \frac{1}{2^n-k^2}
\end{align}
for any $s_0\notin S_k$ and $0\leq k\leqt_{\mathrm{cla}}$. In other words, modulo
``$s\notin S_k$'' the actual values of $f$ and $g$ at points
$\{X_i|i\in[k]\}$ provide no additional information about $s$, and the
best the algorithm can do in that case is a random guess, which
succeeds with probability at most $1/(2^n-k^2)$.
Now let us analyze the probability that
$S_{t_{\mathrm{cla}}}=\{X_i-X_j|i,j\in[t_{\mathrm{cla}}]\}$ contains $s$, that is,
$\PR{E_{t_{\mathrm{cla}}}}$. Since $\sz{S_{k+1}}-\sz{S_k}\leq k$,
we have by the union bound
\begin{align*}
\PR{E_{k+1}|\neg E_k}
&\leq\sum_{s_0\in S_{k+1}}\PR{s=s_0|\neg E_k} \leq \frac{k}{2^n-k^2}.
\end{align*}
Consequently,
\begin{align*}
\PR{E_{t_{\mathrm{cla}}}}\leq \frac{\sum_{k=0}^{t_{\mathrm{cla}}-1}k}{2^n-t_{\mathrm{cla}}^2}
\leq\frac{t_{\mathrm{cla}}^2}{2^n-t_{\mathrm{cla}}^2}.
\end{align*}
Finally, we can bound the probability that the algorithm succeeds
after $t_{\mathrm{cla}}$ oracle queries as
\begin{align*}
\PR{\mathcal{A}_{\mathrm{cla}}\textrm{ succeeds}}
&=\PR{\mathcal{A}_{\mathrm{cla}}\textrm{ succeeds}|E_{t_{\mathrm{cla}}}}\cdot\PR{E_{t_{\mathrm{cla}}}}\\
&\qquad +\PR{\mathcal{A}_{\mathrm{cla}}\textrm{ succeeds}|\neg E_{t_{\mathrm{cla}}}}\cdot\PR{\neg E_{t_{\mathrm{cla}}}}\\
&\leq \PR{E_{t_{\mathrm{cla}}}}+\PR{\mathcal{A}_{\mathrm{cla}}\textrm{ succeeds}|\neg E_{t_{\mathrm{cla}}}}\leq\frac{t_{\mathrm{cla}}^2+1}{2^n-t_{\mathrm{cla}}^2},
\end{align*}
which is larger than $1/2$ only if $t_{\mathrm{cla}}\in\Omega\left(2^{n/2}\right)$,
as required. \hfill $\Box$
\end{proof}
We are now ready to state our main theorem which is an exponential
quantum-classical separation for an average case problem.
\begin{theorem}[Average case exponential separation]
Let $({\cal O}_f,
{\cal O}_g)$ be an instance of a Boolean hidden shift problem (BHSP)
where $g(x)=f(x+v)$ and $f$ and $v$ are chosen uniformly at random.
Then there exists a quantum algorithm which finds $v$ with bounded error using $O(n)$
queries and in $O(\mathrm{poly}(n))$ time whereas any classical algorithm needs
$\Omega(2^{n/2})$ queries to achieve the same task.
\end{theorem}
\begin{proof}
For a fixed $v$ and randomly chosen $f$, consider the $2^{n-1}$
mutually independent events ``$f(x)=f(x+v)$''. By definition of
$\gamma_{f,v}$ and the Chernoff bound, the probability that
$\gamma_{f,v}<1/3$ is at most $e^{-\Omega(2^n)}$. Since this is
double-exponentially small in $n$ we obtain from an application of
the union bound to the $2^n$ possible values of $v$ that if
$f:\mathbb{Z}_2^n\to\mathbb{Z}_2$ is chosen uniformly at random then ${\bf
Pr}_f[\gamma_{f}<1/3]\in e^{-\Omega(2^n)}$. We now apply Corollary
\ref{cor:complexity-promise} for constant $\gamma_{f}$ to obtain a quantum
algorithm that uses at most $O(n)$ queries and outputs the correct
hidden shift $v$ with constant probability of success (i.e.,
$\varepsilon$ is chosen to be constant). Combining this with the
exponential lower bound from Lemma~\ref{l_cla_h} implies that there
is an exponential gap between the classical and quantum complexity
of the BHSP\ defined over a random Boolean function. \hfill $\Box$
\end{proof}
\section{Discussion and open problems}
We presented a quantum algorithm for the Boolean hidden shift problem
that is based on sampling from the space of vectors that are
orthogonal to the hidden shift. It should be noted that our algorithm
reduces to one of the two algorithms given in \cite{Roetteler:2010} in
case the function is a bent function. We related the running time and
the query complexity of the algorithm to the minimum influence of the
function and showed that for random functions these complexities are polynomial. This leads to
an average case exponential separation between the classical and quantum time complexity
for Boolean functions. An interesting question is whether these methods
can be generalized and adapted for the case of non-Boolean functions
also. Furthermore, we conjecture that the complexity of our quantum algorithm is optimal up to polynomial factors for any function.
\subsection*{Acknowledgments}
The authors acknowledge support by ARO/NSA under grant
W911NF-09-1-0569. We wish to thank Andrew Childs, Sean Hallgren,
Guosen Yue and Ronald de Wolf for fruitful discussions.
|
\section{Introduction}
In the last decade interdisciplinary research on complex networks resulted in spectacular development \cite{AlbertBarab}-\cite{DorMend}. It has become clear that networks constructed from diverse complex systems show remarkably similar features. Several aspects were investigated, like clustering \cite{WattsStrogatz}, the degree distribution\cite{degreedistr}, diameter \cite{diam}, \cite{ultrasmall}, spreading processes \cite{spreading}, diffusion \cite{diff}, synchronization \cite{sync}, critical phenomena \cite{crit_phenom} and game theoretical models on complex networks \cite{gametheory}.
One of the most actively researched questions about complex networks is the one of community detection \cite{SF_review}. Community detection aims at finding dense groups in graphs, like circles of friends in social networks, web pages about the same topic, or substances appearing in the same pathway in metabolic reaction networks. Perhaps the strongest motivation behind the research is that dense groups in the topology are expected to correspond to functions performed by the network, such that one can infer from pure topology to function. While the concept of communities seems intuitively plausible, attempts for an algorithmically useful definition have not been successful yet. The global characterization by modularity \cite{Q} or by random walks \cite{MAS, Infomap}, the local ``weak`` and ``strong`` definitions \cite{Radicchi}, the clique percolation approach \cite{CPM}, or the multiresolution methods \cite{RB,AFG,KSKK} have all increased out understanding of this complex problem but the proliferation of methods of community detection just indicates the difficulty of this issue \cite{SF_review}.
Unfortunately, any precise definition of communities is still lacking, giving rise to innumerable methods using different definitions. Lack of a definition also makes problematic the testing of methods; although there is progress in this issue \cite{SFbenchmark1}, \cite{SFbenchmark2}. Difficulty of the problem is increased by more subtle factors: very often communities occur on a broad scale, they can be ordered in a hierarchical manner, and they may overlap, which make their identification even harder.
After being the subject of active research for several years, it is getting clear that the following stages appear during community detection:
\begin{itemize}
\item[1] defining the term ``community'';
\item[2] finding the objects corresponding to the definition;
\item[3] determining the significance of the found communities.
\end{itemize}
Although from the theoretical perspective stage 1 is clearly a key issue, it is far from being settled. Several different propositions exists, which are evaluated mostly according to their results on a few benchmarks. This is the stage to be improved in the first place in this paper. Stage 2 is a technical issue, often consisting of some combinatorial optimization method. Its choice is usually a result of a trade-off between speed and quality. Stage 3 should give information about how surprising is the existence of a found community in the actual graph, given some characteristics of the graph like edge density or degree distribution. Although this issue also got some attention \cite{Q_0value}-\cite{stat_signif_Gfeller}, it just began to get widespread application \cite{OSLOM}.
The rest of the paper will focus on the question of definition, so a few remarks about stage 3 are made here. Most community detection methods give no information about the significance of their output, thus forcing the investigator to assume that all results are (equally) significant. This way, the community detection stages 2 and 3 are combined into a single decision whether a particular subgraph is a good enough community or not -- effectively pruning the significance test in practice. The other end of the spectrum, represented by \cite{OSLOM}, builds the definition of communities on statistical significance, which is clearly an improvement. However, it should be noted that the fitness and statistical significance of a subgraph as a community are not synonyms. Statistical significance tells us how surprising a subgraph is, while fitness talks about how close is it to the ideal community. Therefore, the two quantities are complementary and both belong to the description of a community.
\section{Local criteria for communities} \label{s:def}
A fundamental problem of community detection is to define the term ``community''. There are different approaches to this question. One is the algorithmic approach, giving a computational procedure for finding clusters. This naturally incorporates a mathematically precise definition, although different algorithms usually result in diverse definitions, and there is no theoretical framework currently to help their differentiation. Another possibility is to present a general concept, on which a precise definition can be based. In this paper, the latter approach is taken, although an algorithmic realization is also presented.
No definition of communities which is both precise and generally accepted has appeared yet. Currently the description of communities exhausts in the phrase ``nodes having more edges among themselves than to the rest of the graph'' (or equivalent forms). It can be translated roughly to ``statistically significant locally dense subgraphs``. Statistical significance is a quite precise expression, the main problem is with the term ``locally dense``. For an intuitive picture, it is quite good, but much less than directly transformable to algorithms. Although there is an implicit agreement on that clearly counterintuitive results are not permitted, even a formal list of required properties is missing. However, there are some properties which fit human intuition about locally dense subgraphs\footnote{For brevity, the words ``community'', ``group'' and ``cluster'' will be used from this point as synonyms for ``locally dense subgraph'', omitting the statistical significance from the meaning.}\footnote{It should be noted that the meaning of the term ``community'' can depend on the context; consequently a single definition may not be enough. Here the aim is to describe a particularly intuitive one.}:
\begin{description}
\item[Separation:] a good community is well-separated from the rest of the graph;
\item[Cohesion:] a good community is homogeneously well connected inside, i.e. it is hard to separate into two communities.\footnote{The term ``cohesion'' also appeared in \cite{RB}, although there it denotes a quantity with an unrelated concept.}
\end{description}
The separation criterion is quite clear, although there is an important remark: separation should be defined locally, involving only the community under investigation and its immediate neighborhood. Global methods, in which distant regions of the graph can modify a community in order to improve a global fitness value, can produce results violating the human perception about clusters. A famous example is the resolution limit of modularity \cite{reslim, reslim2}.
Although separation is a very intuitive criterion, and famous methods rely on it (see the Appendix), it is not enough in itself. Figs. \ref{fig:2cliques}-\ref{fig:sw_2cliques} illustrate that the distribution of links inside the separated region (the ``shape`` of the subgraph) also matters heavily. Application of current community detection methods to real-world networks confirms that this is a real problem, e.g. tree-like communities can occur, even when the whole network is not tree-like \cite{various_comparison}, \cite{mobilephone_comparison}.
Both separation and cohesion are required properties of communities. If one neglects cohesion, the result may contain clusters like the one on Fig. \ref{fig:2cliques}. On the other hand, if separation is not taken into account, one may end up chopping a separated subgraph until very cohesive pieces (cliques in the extreme) are obtained, like the triangles on Fig. \ref{fig:sw_2cliques}.
\begin{figure}[!h]
\begin{center}
\subfloat[]{\label{fig:2cliques}\includegraphics*[width=1.5cm]{figure_1a.pdf}}
\qquad\qquad
\subfloat[]{\label{fig:sw_2cliques}\includegraphics*[width=3cm]{figure_1b.pdf}}\\
\end{center}
\caption{Illustration of the importance of subgraph shape. The two subgraphs
have the same number of nodes and the same degrees, i.e. they differ only in the distribution of links. The figure on the left is much less cohesive than the figure on the right, although just a reorganization was applied to the links.}
\label{fig:counterexample}
\end{figure}
Given the subgraphs on Figs \ref{fig:2cliques} and \ref{fig:sw_2cliques} as proposed communities, most community detection methods' fitness values, to be reviewed in the next section, can not tell the difference between them. This is due to that most methods simply count the internal and/or external edges, which do not tell about the distribution of those edges. The reason why several methods do not fail to assign proper clusters for Fig \ref{fig:2cliques} is that they look for optimal clusters, consequently they compare configurations like Fig \ref{fig:2cliques} in one cluster and in two clusters, and splitting the two cliques into two clusters may improve the partition. But the situation is even worse. In the next Section, we will see that a number of fitness functions are more optimal for a counterintuitive clustering than for the intuitive one (e.g. joining the two cliques on Fig \ref{fig:2cliques}, like modularity for a large enough graph). It should be noted that in such a case, the proper communities might be recovered if the heuristic gets stuck in the proper local optimum, even when that is not the global optimum.
\section{Overview of the existing methods}
Here, existing community detection methods will be reviewed from the point of view of the previous section, i.e. do they conform the criterions of separation and cohesion. As mini-benchmarks, the examples on Fig. \ref{fig:counterexample} or their simple variations will be used (see the Appendix for details on individual methods). The desired output for \ref{fig:2cliques} is two communities consisting of the two cliques, while \ref{fig:sw_2cliques} should be kept in one piece. In both cases, no nodes from the rest of the graph should be included. For methods optimizing a fitness function, the globally optimal solution will be considered, for other methods, the possible solutions. These solutions will be compared to the desired ones, independently for Fig. \ref{fig:2cliques} and \ref{fig:sw_2cliques}. If a method separates the two cliques of Fig. \ref{fig:2cliques}, then it gets a ``+'', if it puts all nodes of Fig. \ref{fig:sw_2cliques} into one cluster, then it gets another ``+''. If there are multiple equally valid solutions (like for label propagation), all solutions are required to conform the preferred result. \\
For methods optimizing a function, the heuristic realizing the optimization may deviate from the global optimum, presenting worse or even better results (in terms of conformity to separation and cohesion). This will not be investigated, here the focus is on the definition of the communities (following from the choice of the fitness function), not on the practical aspects. Results for methods which can produce a single partition or cover are displayed in Table \ref{t:review}. The large number of published methods makes assembling a complete list nearly impossible. Instead, the emphasis is put on the diversity of the reviewed approaches.
\begin{table}[!h]
\begin{center}
\begin{tabular}{ c c c c }
method & cohesion test & separation test \\
& (like Fig. \ref{fig:2cliques}) & (like Fig. \ref{fig:sw_2cliques}) \\
\hline
Lancichinetti et al. \cite{Lancichinetti} & - & + \\
Labelpropagation \cite{labelprop} & - & - \\
Infomap \cite{Infomap} & - & + \\
Clique Percolation \cite{CPM} & - & - \\
Estrada \& Hatano \cite{Estrada} & - & - \\
Modularity optimization \cite{Q} & - & + \\
Donetti \& Mu\~noz \cite{DM} & - & + \\
Ronhovde \& Nussinov \cite{RN} & - & - \\
Nepusz et al. \cite{Nepusz} & + & - \\
Hofman \& Wiggins \cite{HW} & - & - \\
Hastings \cite{Hastings} & - & - \\
Newman \& Leicht \cite{NewmanLeicht} & + & - \\
Wang \& Lai \cite{NL++} & + & - \\
Bickel \& Chen \cite{BickelChen} & + & - \\
Karrer \& Newman \cite{Karrer} & + & - \\
Infomod \cite{Infomod} & - & + \\
Radicchi et al. \cite{Radicchi} & + & - \\
Chauhan et al. \cite{lambda1} & + & - \\
Evans \& Lambiotte \cite{EvansLambiotte} & - & + \\
Ahn et al. \cite{Ahn} & - & - \\
ModuLand \cite{Moduland} & - & - \\
\hline
\end{tabular}
\caption{Cohesion \& separation criterion test results. Tests were done on Fig. \ref{fig:2cliques} and \ref{fig:sw_2cliques} or similar graphs (which are described in the Appendix). + and - are assigned according to whether the fitness function of a method is more optimal for the preferred solution or not. For methods which do not optimize a fitness function, simply the possible solution(s) was (were) analyzed. See the Appendix for details on specific methods.}
\label{t:review}
\end{center}
\end{table}
There is a bunch of multiresolution methods, which possess a parameter allowing to tune the cluster sizes from $1$ (isolated nodes) to $\mathcal{O}(N)$: the multiresolution modularity of Reichardt and Bornholdt (RB) \cite{RB}, of Arenas, Fernández and Gómez (AFG) \cite{AFG}, the local fitness method of Lancichinetti, Fortunato and Kertész (LFK) \cite{Lancichinetti}, the Potts model of Ronhovde and Nussinov (RN) \cite{RN}, the Markov autocovariance stability of Delvenne, Yaliraki and Barahona (MAS) \cite{MAS}, the hierarchical likelihood method of Clauset, Moore and Newman (CNM) \cite{newman_nature}, and the Markov Cluster Algorithm of van Dongen (MCL) \cite{MCL}. Naturally, these methods are expected to find the proper community assignments both to Fig. \ref{fig:2cliques} and \ref{fig:sw_2cliques} at some parameter values. However, there is no guarantee that these values are also the proper ones for the rest of the graph. Consequently, it is not clear how a resolution parameter should be set: the natural idea is to find the longest interval of the resolution parameter value in which the community structure does not change, but when the optimal parameter value is different for different regions in the graph, the longest stable interval not necessarily reflects the optimal communities.\\
Furthermore, the fitness values do not help us to tell good clusters from bad ones, like Fig. \ref{fig:2cliques} from Fig. \ref{fig:sw_2cliques}. For most multiresolution methods (RB, AFG, LFK, RN), it is very easy to see that the fitnesses of two clusters are the same given that all nodes has the same in- and outdegrees, independently of the shape of the clusters. Note that it is also true for most single resolution methods. For MAS it is not trivial. Therefore, empirical tests were conducted to check it. According to them, Fig. \ref{fig:2cliques} was found empirically to be at least as good as \ref{fig:sw_2cliques}\footnote{In this case, only 1 link to the rest of the graph was used. Rest of the graph, represented by a single node having self-loops, was assigned 118 edges inside, resulting in a total of $L = 150$ edges. Stability values were calculated from $0.01$ to 100, the step size being $0.01$ below $1.0$ and 1 above.}. Finally, regarding MCL and CNM, they have no fitness function\footnote{CNM does have a fitness function, but it corresponds to a full hierarchical dendrogram, not to any partitions obtained by cutting the dendrogram at some point}, the only accessible quantity about the community structure is the parameter interval in which it is stable.
Finally, there are hierarchical methods, which look for series of smaller and smaller (or larger and larger) clusters hierarchically embedded into the previous ones. Similarly to multiresolution methods, they are expected to contain good clusters in the outputted hierarchy. However, when looking at a graph having a simple one-level community structure, the question how to select the proper levels of the outputted hierarchy arises. The easiest way is to use the lowest level communities. Unfortunately, it is not a reliable procedure, as the lowest-level clusters may be just parts of the communities of the optimal partition or cover (see the Appendix for details). A second idea can be to assign significance scores to the communities on different levels, in the spirit of \cite{stat_signif_Andrea}. Although this approach might reliably qualify the found communities, a new version of statistical significance taking into account the internal cohesion is required. Furthermore, one should be very careful not to impose unnecessary constraints, like prohibiting overlaps, when constructing a hierarchical method.
A further question is whether a method provides information about the shape of the found communities or not. Recent analysis of real-world networks highlights the relevance of this issue \cite{various_comparison}, \cite{mobilephone_comparison}. Several methods are based on simply counting the internal and/or external edges, or degrees at most: LFK, Labelpropagation, Infomap, modularity optimization (and equivalents), Hofman \& Wiggins, Hastings, Ronhovde \& Nussinov, Newman \& Leicht, Wang \& Lai, Bickel \& Chen, Karrer \& Newman, Infomod, Ahn et al., OSLOM. Consequently, they do not see any difference in the distribution of the links, e.g. Fig \ref{fig:2cliques} and \ref{fig:sw_2cliques} get the same fitness values. Only Clique Percolation and Radicchi et al.'s\footnote{If the stopping criterion of their heuristic is considered as part of the definition.} method have some very limited requirement about cohesion built in the definition of communities.
The conclusion is that none of the reviewed methods is able to successfully apply both the separation and the cohesion criterions. They susceptible either to glue together well-separated subgraphs or to overpartition a cohesive subgraph. Future network designs should consider cohesion as well as separation.
\section{Community detection in a two dimensional parameter space}
In this Section, a new method for community detection is introduced. Its main goal is to present a method which takes into account both criterions defined in Sec. \ref{s:def}. First, the LFK method will be reviewed, which will serve as a starting point for the new method. Then, a composite fitness will be constructed which takes into account the separation and cohesion criterions. Finally, a heuristic optimization procedure for the composite fitness will be described, which finds locally dense subgraphs on all scales, and also able to recover hierarchical structures.
\subsection{The LFK method}
The LFK method \cite{Lancichinetti} optimizes the local fitness function
\begin{equation} \label{e:LFK}
f^C=\frac{K_{in}^C}{(K_{in}^C+K_{out}^C)^{\alpha}}
\end{equation}
where $C$ denotes a subgraph, $K_{in}^C$ and $K_{out}^C$ are the total number of inside and outside degrees in $C$, respectively, and $\alpha$ is a tunable exponent for setting the size scale of the communities to be found. Running the method with large $\alpha$ values result in small clusters, small values in large clusters. The recommended range for $\alpha$ is $0.5$-$2$.
The practical implementation of the optimization works as follows. The communities are found one-by-one, independently of each other. First, a seed node is selected from which the new community will be grown. Then, the node which can best improve the fitness of the cluster is added. This addition is repeated until the fitness reaches a local optimum. After each addition, removal of nodes takes place, if the fitness can be enhanced that way. When the fitness cannot be further increased, the actual subgraph is declared a community. The growth process is repeated for all nodes as seeds, or alternatively, until the found communities cover all nodes in the graph.
Although the resolution parameter $\alpha$ can be tuned continuously, \cite{Lancichinetti} suggested that the relevant community structures should be identified by robustness to changes in $\alpha$, i.e. which have the longest interval for $\alpha$ values without change. Changes in the community structure were detected by monitoring the mean fitness of the communities, evaluated at a reference value $\alpha=1$.
\subsection{Implementing the criterions}
For the separation criterion, the following function will be applied
\begin{equation} \label{e:f_S}
f_{S}^C = \frac{K_{in}^C}{K_{in}^C+K_{out}^C}
\end{equation}
where $C$ is a subgraph, $K_{in}$ and $K_{out}$ are the sums of in-community and out-community degrees, respectively. This is the fitness of LFK \cite{Lancichinetti}, with the multiresolution parameter being set to one. For detecting hierarchical structures, a different solution will be described. Eq. \ref{e:f_S} clearly focuses on the external separation of the clusters, therefore it is suitable as an implementation of the first criterion of the communities.
For the internal cohesion criterion, a possible solution is to consider the second eigenvalue of the Laplacian matrix of the community. The Laplacian of a graph is the matrix $L=A-D$, where $A$ is the adjacency (or weight) matrix, and $D=\diag(k_i)$ is a diagonal matrix containing the degrees (strengths). Its largest eigenvalue is always 0 (corresponding to the trivial eigenvector $(1,1,\ldots,1)$). The multiplicity of the largest eigenvalue equals to the number of connected components in the graph. This gives the hint that if two distinct graphs are got connected by a single (weak) link, the Laplacian gets only a slight perturbation (compared to the case of two connected components), which splits the double degeneracy of the first eigenvalue, such that a new eigenvalue close to zero appears\footnote{The diffusion matrix was also considered, but it prefers star-like graphs too much.}. In fact, it is known that the second eigenvalue of the Laplacian measures ``how difficult is to split the graph into two large pieces'' \cite{Mohar}.
For some important special cases the second eigenvalue can be calculated:
\begin{itemize}
\item[-]for full graphs of $n$ nodes (clique), $\lambda_2=-n$
\item[-]for a star-graph, $\lambda_2=-1$, independently of $n$
\item[-]for a linear chain, $\lambda_2=-2+2\cos(\pi/n)\to0$ as $n\to\infty$
\item[-]for two $n$-sized cliques attached by a single link (having weight $\epsilon$) (like on Fig. \ref{fig:2cliques}), $\lambda_2\approx-\frac{2\epsilon}{n+2\epsilon}$, which also goes to 0 as $n\to\infty$.
\item[-]for a disconnected graph, $\lambda_2 = 0$. This may seem trivial, but most methods give a finite score for disconnected communities; it is not without precedent that such objects can be produced in reality \cite{mobilephone_comparison}. Although this problem can be avoided by a properly designed heuristic of a method, disconnected communities should be punished by definition.
\end{itemize}
Calculation for the two cliques is in the Appendix, other results can be found in \cite{Fiedler}. These cases confirm that the second eigenvalue is useful for quantifying the cohesion criterion of the definition of communities. For an illustration, on Fig. \ref{fig:lambda2_series} a few example graphs with their second Laplacian eigenvalues are shown.
\begin{figure}[!h]
\begin{center}
\subfloat[]{\label{fig:l2series_1}\includegraphics*[width=2.5cm]{figure_2a.pdf}}
\qquad\qquad\qquad
\subfloat[]{\label{fig:l2series_2}\includegraphics*[width=3cm]{figure_2b.pdf}}\\
\subfloat[]{\label{fig:l2series_3}\includegraphics*[width=2.2cm]{figure_2c.pdf}}
\qquad\qquad
\subfloat[]{\label{fig:l2series_4}\includegraphics*[width=5cm]{figure_2d.pdf}}
\end{center}
\caption{Graphs with different second Laplacian eigenvalues. $\lambda_2^{(\text{a})}=6$, $\lambda_2^{(\text{b})}=2$, $\lambda_2^{(\text{c})}=1$, $\lambda_2^{(\text{d})}=0.268$. The maximal value of $\lambda_2$ is 6 in all cases.}
\label{fig:lambda2_series}
\end{figure}
The separation fitness term $f_S$ ranges from zero to one. In order to compose it together with the cohesion fitness, the latter should also be in the interval $[0,1]$. Therefore, $\lambda_2$ needs some transformations before application as fitness.\\
As can be seen from the above examples, for the worst cases $|\lambda_2|$ is of the order of $1/n$, therefore the lowest point of the $|\lambda_2|$-scale will be set to $1/n$. The highest point is trivially given by $n$. It is reasonable to assume that most subgraphs have $|\lambda_2|=\textrm{o}(|C|)$. Furthermore, several subgraphs can have worse internal cohesion than the star graph, thus having $|\lambda_2| \in [0,1]$. To take into account these effect, $\log |\lambda_2|$ will be more useful than $\lambda_2$. So, in order to obtain a quantity between 0 and 1, the minimum will be subtracted and divided by the maximum,
\begin{equation} \label{eq:lambda2}
\begin{split}
f_{C}^C=&\frac{\log |\lambda_2| - \log 1/|C|}{\log |C| - \log 1/|C|} = \frac{1}{2}+\frac{1}{2}\frac{\log|\lambda_2|}{\log|C|}, \quad \text{if } |C|>1\\
=&\,0\quad\qquad\qquad\qquad\qquad\qquad\qquad\quad\quad\;\;\,\;\;\: \text{if }|C|=1
\end{split}
\end{equation}
where $|C|$ is the number of nodes in the community. The above measure happens to be $0.5$ if $\lambda_2$ is 1, e.g. for the star-graph. I wish to emphasize that eq. \ref{eq:lambda2} is only one possible proposition for taking into account the internal cohesion, although a promising one -- better measures may exist. The same is true for the choice of $f_{S}$.
The cohesion fitness $f_C^C$ opens the way for constructing tests assessing the performance of community detection methods regarding the cohesion of the found communities. One may generate a graph with built-in communities which separation is controlled, like in the LFR benchmark \cite{SFbenchmark1}, then randomly select pairs of clusters and increase the interconnection between the two members of each pairs to some predefined value, finally calculating $f_C^C$ of the pairs. Running the detection method and measuring the ratio of pairs not split as a function of $f_C^C$ may indicate how strongly focuses the method on cohesion.
The next question is how to combine $f_{separation}^C$ and $f_{cohesion}^C$. Thinking in a two dimensional space of $f_{S}^C$ and $f_{C}^C$, a natural approach is to get as far from the point $(0,0)$ as possible. This implies
\begin{equation} \label{e:fullfitness}
f^C=\sqrt{(f_{S}^C)^2+(f_{C}^C)^2}
\end{equation}
so the fitness is the euclidean distance from $(0,0)$. Again, this is just one possibility, better combinations may exist. E.g. the relative weight of $f_{S}$ and $f_{C}$ may be adjusted in a more well-grounded way. However, eq. \ref{e:fullfitness} is able to pass the test raised by Fig. \ref{fig:counterexample}: for Fig. \ref{fig:2cliques}, $\lambda_2^{2\text{cliques}}=0.258$, $f_C^{2\text{cliques}}=0.228$, $f^{2\text{cliques}}=0.995$ while for a single clique $\lambda_2^{1\text{clique}}=6$, $f_C^{1\text{clique}}=1$, $f^{1\text{clique}}=1.371$. For Fig. \ref{fig:sw_2cliques}, $\lambda_2^{12\text{nodes}}=3.268$, $f_C^{12\text{nodes}}=0.738$, $f^{12\text{nodes}}=1.218$, and for the best subgraph, a triangle, $\lambda_2^{\text{triangle}}=3$, $f_C^{\text{triangle}}=1$, $f^{\text{triangle}}=1.077$.
Beyond enabling one to decide whether a given subgraph is a community or not (by requiring local optimality), the above definition makes it possible to assess how good community it is. This is also possible with another definitions, e.g. by using the modularity function, but here, communities are placed on a 2-dimensional space instead of 1 dimension. This gives rise to an interesting possibility for characterizing the communities, like ``very cohesive but densely connected outwards'' or ``well-separated but poorly interconnected''. Considering Fig.\ \ref{fig:2cliques}, one may think that the latter is not really a community. But for large subgraphs, it may make sense to consider a well-separated subgraph as a community, as common sense says that large communities should be looser than small ones.
\subsection{Community detection in reality}
In this section, the details of practical implementation of the new method are discussed. Most importantly, in order to actually find the communities, a heuristic carrying out the optimization of eq. \ref{e:fullfitness} is needed. Furthermore, there is a second problem of detecting communities hierarchically embedded into each other. These two questions will be answered by a common solution.
The heuristic is based on the one of the LFK method \cite{Lancichinetti}. Among its details, the LFK heuristic contains a tunable parameter (denoted as $\alpha$), which is claimed to be able to recover communities at different hierarchical levels. Lowering this parameter $\alpha$ results in increased community sizes. Hierarchical levels are supposed to be stable against the variation of $\alpha$, so there should be long intervals for $\alpha$ for which the communities do not change. However, large graphs may lack long stable intervals, as some changes occur around any parameter value (data not shown). Therefore, a new method for investigating hierarchical structure is needed. I dropped the idea of using threshold values of $\alpha$, corresponding to community structures at different scales, which should be simultaneously valid for all communities, and I will treat each community separately.
Similarly to \cite{Lancichinetti}, each community is grown from a seed node. It is important to note that each seed node can result in a series of (successively larger) communities. Growth consists of successively including the neighboring node which increases most the fitness defined by eq. \ref{e:fullfitness}. When there is no neighboring node which inclusion can improve the fitness, the stage of node removal begins. Here, the fitness of the cluster is tried to be improved by excluding nodes from it (with the exception of the seed node, which is not permitted to be excluded). It finishes when no further removal can improve the fitness. Then, growth begins again, if possible. The grow-shrink cycle is iterated, as long as the fitness can be improved. When no improvement is possible (there is a local optimum of the fitness), the actual list of nodes is registered as a valid community. After that, the algorithm tries to find a larger community, which contains the current one. This way, hierarchical structures can be revealed. In order to do it, first the growing cluster should escape from the basin of attraction of the current local optimum. Therefore, the cluster is forced to grow, by successively including the neighboring nodes which decrease the fitness the least. After some steps of forced growth, when increasing the fitness becomes again possible, the algorithm turns back to the normal grow-shrink procedure, until a new local optimum is found, signing a new community. The cluster keeps hopping from local optimum to another local optimum until it grows so large that it contains the whole graph. Then a new growth process starts from a new seed node. At the end of its growth process, it includes the whole graph again, unless it encounters a local optimum which has been already found, i.e. the corresponding community has already been registered. In this case, the growth process is stopped. Then, another growth process starts from a not-yet-used seed node. In contrast to \cite{Lancichinetti}, all nodes in the graph are used as seed nodes, in order not to miss good communities. When the growth process beginning from the last seed node finishes, the algorithm ends, and the registered communities are written to the output.\\
There are a few additional tricks. First, if escaping from a local optimum seems to be hard, i.e. after changing from forced growth to the normal grow-shrink stage we still end up in the previous local optimum, the cluster is restored to the state where it had its maximal size (the beginning of one of the removal sessions), then 2 steps of forced growth is applied before the normal grow-shrink cycle begins. A second trick is that when judging the identity of two communities, they are considered identical if at least $80\%$ of the larger community is a subset of the smaller one\footnote{If the criterion were based on some percent of the smaller group, subset-superset pairs would be considered identical.}. In case of identity, the community which has the higher fitness is kept in the registry.\\
The algorithm, although based on the one of \cite{Lancichinetti}, differs in several points: from one seed, several communities can be reached instead of only the smallest one; node removal occurs when node addition is not possible instead of after each addition (this trick also speeds up the algorithm); seed node is not permitted to be removed; all nodes are used as seeds instead of the not-yet-covered nodes. An algorithm similar in spirit was described in \cite{clauset}.
The results in the next section are obtained using this method, unless stated otherwise explicitly. The software realizing the algorithm is available at\\ \verb+http://www.phy.bme.hu/~tibelyg/+.
\section{Test results}
Probably the most frequently used test is Zachary's karate club friendship network\cite{Zachary source}. Due to a dispute between two prominent persons (node 1 and 34), the club split into two during sociological observation, and the memberships in the new clubs are known. As the split occurred more or less along a border of two visible communities, new community detection algorithms are usually claimed to pass the test if they reproduce the split. However, the aim is the detection of \emph{topological} modules, not functional ones, so the result of the sociological study is not a strict criterion for judging the output of any community detection method. E.g., node 10 has $1-1$ links to each of the new clubs, so ``misplacing'' it (compared to the split) may not be considered as a fault. Or node 12, which attaches only to node 1, is hard to be considered as part of a ``densely interconnected'' cluster.\\
The algorithm finds $33$ groups, containing several non-relevant ones, like pairs of nodes. Therefore, a filtering procedure is required. The statistical significance of the resulting communities \cite{stat_signif_Andrea, OSLOM} is utilized for this purpose. The statistical significance can be sensitive for missing nodes \cite{stat_signif_Andrea}, therefore each cluster is allowed to be completed with the neighboring node which optimizes the statistical significance. Then the clusters are ordered according to their statistical significance. The first 3 clusters provide a single-level community structure, corresponding to 3 known communities, with 2 overlapping and 1 homeless nodes (Fig. \ref{fig:zach}, left panel). Taking a look at the subsequent clusters provides information about the multi-scale structures in the graph. The next few clusters reveal cluster cores and hierarchical decomposition of the network (Fig. \ref{fig:zach}, right panel). The statistical significance score is quite capable of distinguishing meaningful structures; there is a gap between $0.42$ and $0.81$, so setting a threshold to $0.5$ selects the multi-scale clusters which would be approved by a human investigator. There is only one exception, the almost-full-clique of nodes \{1, 2, 3, 4, 8, 14\} has significance $0.81$, which is probably the consequence of neglecting the internal cohesion by the current form of statistical significance.
\begin{figure}[!h]
\begin{center}
\includegraphics*[height=9cm]{figure_3a.pdf}
\phantom{a}
\includegraphics*[height=9cm]{figure_3b.pdf}
\end{center}
\caption{(Color online) The 3 (left) and 10 (right) best found communities of the Zachary karate club. On the right, thicknesses of lines indicate the ordering of the statistical significance values (running from $0.002$ to $0.42$, plus $0.81$ for the dashed line-bordered community). Note that node 12 is contained only by large communities.}
\label{fig:zach}
\end{figure}
\begin{figure}[!h]
\begin{center}
\includegraphics*[height=6cm]{figure_4.pdf}
\end{center}
\caption{(Color online) Positions of the Zachary communities on the $f_{S}$-$f_{C}$ plane. Small groups tend to cluster at North, and large groups at East.}
\label{fig:zach_f1-f2}
\end{figure}
The currently most advanced class of benchmarks was introduced by \cite{SFbenchmark1}. In these so-called LFR benchmarks, the network size and edge density are freely adjustable, and more importantly, the node degrees and the community sizes are distributed according to power-law distributions, with tunable exponents. Communities are defined through a prescribed ratio of inter-community links for each node (mixing ratio, $\mu$), similarly to the preceding GN benchmark class \cite{Q}. Generalizations for weighted and directed networks, and for overlapping communities also exist \cite{SFbenchmark2}.\\ A wide-scale comparison of different community detection methods using the LFR benchmark was done by \cite{comparison}. For the ease of comparison, the parameter values of \cite{comparison} are applied here: the networks consist of 1000 nodes, the average degree is 20, the maximal degree is 50, the exponent of the degree distribution is -2 and the exponent of the community size distribution is -1. There are two types of networks, for the S type the community sizes are between 10 and 50 (``small'') and for the B type they are between 20 and 100 (``big''). In \cite{comparison}, networks of 5000 nodes were also investigated. Due to the large computational time, they are omitted here\footnote{it does not mean that a single 5000-sized graph is too large, however, a few hundred of them are}. Also for computational time considerations, the detecting algorithm stopped growing the communities over a predefined size, 120 for the S case and 220 for the B case. All measurement values are obtained from runs on 10 different networks.\\
Similarity of the built-in and the obtained community structures are quantified by a variant of the normalized mutual information (NMI), which is able to handle overlapping communities \cite{Lancichinetti}. This is the similarity measure applied by \cite{comparison}\footnote{Both the LFR benchmark and the generalized normalized mutual information are freely available from the authors' websites, \texttt{http://sites.google.com/site/santofortunato/inthepress2} and \texttt{http://sites.google.com/site/andrealancichinetti/software}}.\\
Selecting the most relevant communities from the abundant output was done similarly to the previous case. The clusters were completed by 1 neighboring node, if that improved the statistical significance, and sorted with respect to the statistical significance scores. The clusters containing at least 1 uncovered node were accepted one by one until all nodes were covered.\\
To see the potential of the new method, and check the effect of the output-filtering, the communities corresponding best to the built-in original ones were also selected from the algorithm's output. The results are plotted on Fig. \ref{fig:SFtest} (a). The filtered results are similar to the ones of the lower performing algorithms in \cite{comparison}, while optimal selection provides much better scores, although still not as good as the best methods.
The large difference between the optimal and the statistical significance-based results is quite surprising, especially in the light of the fact that statistical significance in itself is able to provide excellent results on the LFR benchmark \cite{OSLOM}.\\
The algorithm was also tested on networks with overlapping communities. In this case, clusters having significance score below $0.1$ were accepted, similarly to \cite{OSLOM}. Fig. \ref{fig:SFtest} (b) shows that the effect of the imperfect output-filtering is again very large, an ideal selection scheme would allow very good results. This is not surprising, as other algorithms based on the one of \cite{Lancichinetti} also give excellent results on overlapping communities \cite{GCE}.\\
\begin{figure}[!h]
\begin{center}
\subfloat[]{\label{fig:SFtest-a}\includegraphics*[width=0.47\textwidth, trim=0.4cm 0cm 0.7cm 0cm, clip]{figure_5a.pdf}}
\hspace{0.5cm}
\subfloat[]{\label{fig:SFtest-b}\includegraphics*[width=0.47\textwidth, trim=0.4cm 0cm 0.7cm 0cm, clip]{figure_5b.pdf}}
\end{center}
\caption{(Color online) Results on the LFR benchmark. Panel (a) corresponds to unweighted, undirected and non-overlapping tests, while panel (b) corresponds to overlapping tests. Overlapping tests were done at two different values of the mixing parameter, at $\mu = 0.1$, $0.3$. For both panels: full symbols and lines correspond to the applied filtering, empty symbols with dotted lines correspond to perfect output filtering.}
\label{fig:SFtest}
\end{figure}
Finally, the new method was applied to a word association graph built from the University of South Florida Free Association Norms \cite{wa}. Here, nodes are words and edges show that some people associated the corresponding two words. The network has 5018 nodes with mean degree $\langle k \rangle = 22.0$. It is a frequently used example of overlapping community structure \cite{OSLOM}, \cite{CPM}. Although edge weights are accessible, the algorithm was applied to the unweighted version of the network. As an illustration, low-level communities around the word \emph{bright} are plotted on Fig. \ref{fig:wa}. An interesting effect is the appearance of \emph{overlapping edges}, due to the heavy overlap in the network.
\begin{figure}[!h]
\begin{center}
\includegraphics*[height=10cm]{figure_6.pdf}
\end{center}
\caption{(Color online) Communities around \emph{bright}, on the first hierarchical level. Color denotes communities. Gray shows overlapping nodes and edges. Black edges are between different communities.}
\label{fig:wa}
\end{figure}
In conclusion, although selecting the relevant communities from the output is not an already solved task, the algorithm gives good results on the Zachary karate club, and performs reasonably on the LFR benchmarks. It should be noted however, that due to the internal cohesion criterion, this algorithm's output is not intended to perfectly match benchmarks like GN and LFR, which define communities solely on the basis of external separation. An additional observation is reported here: on GN benchmark graphs\footnote{results are omitted, as the presented LFR benchmark is a generalization of the GN.} with nodes having exactly the prescribed in- and out-degrees, at large mixing ratios communities deviating from the built-in ones but having better-than-designed mixing ratios were found. Note that the new method does not optimize just for external separation, so even better ``spontaneous'' communities may exist. This phenomenon, although not being a huge surprise, raises the question how to judge precisely a community detection method's output at large mixing ratios, as the known community structure may not be trusted to 100\%.
\section{Discussion \& Conclusions}
An important aspect of all community detection methods is the \emph{running time}. In the case of the new method described above, the time requirement is as follows. Starting a new community from each node contributes a factor of $N$ to the CPU time. Evaluating the eigenvalues of a community $C$ plus one extra node takes $2/3\:(|C|+1)^3$. Assuming that $C$ has $\text{const}\cdot\langle k\rangle\cdot|C|$ neighboring nodes (i.e., on average, each node has a constant fraction of its neighbors outside $C$), running time can be estimated as
\begin{equation}
T \approx N\cdot\sum_{|C|=1}^{|C|_{\text{max}}} \text{const}\cdot\langle k\rangle\cdot|C|\cdot\;(|C|+1)^3 \approx N\cdot\text{const'}\cdot\langle k\rangle\cdot|C|_{\text{max}}^5
\end{equation}
A naive estimate for $|C|_{\text{max}}$ would be $N$. However, as more and more community growing processes finish, the newly started communities are expected to terminate in a previously discovered community earlier and earlier, on average. Of course, some communities will reach $|C|=N$. Therefore,
\begin{equation}
T \propto N^{5+\delta},\qquad \delta \in[0,1]
\end{equation}
which is huge and clearly denies the analysis of even medium-sized graphs ($\mathcal{O}(10^4)$ nodes) without further improvements. Note that graphs of thousands of nodes may be manageable, like the word association graph shown above, which took 56 hours on a single CPU. One possibility is to choose the initial seed more intelligently, starting communities from promising seeds. \cite{GCE} achieved good results in this aspect. An intelligent seed selection is also important if the number of communities in a cover is larger than $N$, or if some communities have only overlapping nodes -- in this case, it may happen that all growth processes miss a certain community.\\
Other important question is the applicability of an advanced eigenvalue solver. Arpack++ \cite{arpack++} and SLEPc \cite{slepc} were tried. The experience was that -- despite their good asymptotic performance in the large matrix limit -- for the occurring several small subgraphs the overhead of these complicated machineries was so large that made the final running time much higher than those obtained with the QR-decomposition algorithm.
\\
Doing optimization in a multi-parameter space is a nontrivial task, because different parameters can lie in different ranges. Therefore, an important direction for future research is to investigate the best combination of the parameters in the fitness function, based on the evaluation of empirical data.
\\
Finally, filtering the relevant communities from the found ones is also a challenging task. The natural approach is to apply statistical significance, which should be applied even if filtering was not needed. However, deciding the threshold significance value is not necessarily trivial in all cases. Furthermore, the current form of statistical significance accounts only for the separation of the community, not for its internal cohesion. This manifests itself e.g. in the low score of the almost-full-clique subgraph in the Zachary karate club (the dark purple group on Fig. \ref{fig:zach}). As the main advantage of the fitness function of eq. \ref{e:fullfitness} is the inclusion of cohesion, it would be important to develop a statistical significance taking it into account.
\emph{Conclusions}. The community detection problem currently suffers from two fundamental deficiencies. First, there is no definition of community which is precise enough to allow constructing community finding methods. Second, thorough testing a proposed algorithm is problematic, not independently from the previous difficulty. I attempted to improve both issues.
In this paper, I proposed a formal list of required properties for locally dense subgraphs, taking a step towards an applicable definition of the term ``community''. Two properties, external separation and internal cohesion (``shape'') were named. External separation has already been applied by some of the community detection methods, and also by benchmarks. Internal cohesion was not considered explicitly earlier. No current method was found which satisfactorily applies both criterions. I demonstrated on simple examples that both properties are necessary; discarding either of them leads to counterintuitive results. Beyond allowing to construct new methods, these two criterions can also be used as a basis for testing existing ones. They also allow the characterization of a community by two independent quantities, instead of a single scalar.
I proposed a new composite fitness function which takes the two criterions into account. For the quantification of the internal cohesion, the second eigenvalue of the Laplacian matrix is applied, which provides appropriate results on characteristic graphs like cliques or chains. I also proposed a heuristic, by redesigning the LFK heuristic \cite{Lancichinetti}, which can find overlapping locally dense subgraphs of all scales, producing much less output than multiresolution methods but with less restrictions than imposed by assuming a hierarchical structure. Runs on the Zachary network and LFR benchmarks showed that the method is able to provide the expected results. Overlapping communities can be detected especially efficiently, similarly to other LFK-based heuristics \cite{GCE}. However, significant improvements are yet to be implemented; e.g. reducing the running time, finding a more effective filtering procedure for the output, or fine-tuning the relative weight of the separation and the cohesion terms in the fitness function.
\section{Acknowledgments}
I wish to thank János Kertész for several useful suggestions. I also thank the Eötvös University for the access to its HPC cluster, and Andrea Lancichinetti, who proposed the Arpack++ package and was extremely helpful about his software. Thanks are due to the authors of the OSLOM \cite{OSLOM}, LFR benchmark \cite{SFbenchmark1}, overlapping mutual information \cite{Lancichinetti}, Radatools \cite{radatools} and linegraph-creator \cite{EvansLambiotte} softwares for making their code publicly available. I am grateful for the referees for several comments which significantly improved the manuscript. Financial support from EU's 7th Framework Program's FET-Open to ICTeCollective project no. 238597 is acknowledged.
\newpage
\section*{Appendix}
|
\section{Introduction}
Accretion on to supermassive black holes (SMBHs) located in the centre of active galaxies provides a significant contribution to the
energy irradiated over cosmic times. The spectral shape of the X-ray background and its progressive resolution into individual sources
indicate that most active galactic nuclei (AGN) are heavily obscured by large column densities of dust and gas (e.g. Fabian \&
Iwasawa 1999; Gilli, Comastri \& Hasinger 2007). According to the AGN unification model (Antonucci 1993), the differences observed
in the optical spectra of AGN can be ascribed to the presence of a dusty torus surrounding the central engine: in nearly face-on objects,
classified as type~1, favourable lines of sight allow the exploration of the very inner regions and the detection of broad emission lines on
top of a strong optical/ultraviolet (UV) continuum. On the contrary, such spectral features are absent whenever the symmetry axis of the
system lies close to the plane of the sky and the dusty screen blocks the direct nuclear light, so that only the high-ionization narrow
lines originating from the outer regions are visible in type~2 objects as signatures of the underlying accretion activity. As the dust affects
the optical spectral properties and classification, the amount of gas along the line of sight suppresses the X-ray emission through
photoelectric absorption and Compton scattering. Compton-thick sources (i.e. those with $N_\rmn{H} \ga 10^{24}$~cm$^{-2}$)
are difficult to unveil at high redshift ($z \sim 1$--3) even with the deepest X-ray surveys (e.g. Alexander et al. 2003). None the less,
the absorbed optical to soft X-ray primary radiation is reprocessed by the intervening material and re-emitted at longer wavelengths,
driving a significant luminosity in the mid- and far-infrared (IR). Indeed, compelling evidence for the existence around $z \sim 2$ of a
vast population of Compton-thick AGN among the IR galaxies has been found, by selecting through various criteria those sources
showing excess mid-IR emission with respect to the predictions relative to the star formation component alone (Daddi et al. 2007; Fiore
et al. 2008; Treister et al. 2009; Bauer et al. 2010). \\
The local counterparts of the IR systems harbouring the most obscured nuclear activity at high redshift are the so-called Ultraluminous
infrared galaxies (ULIRGs, $L_\rmn{IR} \sim L_\rmn{bol} > 10^{12} L_{\sun}$), which rival optically-bright quasars as the most
powerful sources in the nearby Universe with their huge mid- and far-IR emission, due to the dust reprocessing of higher-frequency
radiation (Sanders \& Mirabel 1996). It is now well-established that the hidden source of the primary radiation field inside
ULIRGs is a combination of extreme star formation activity (starburst, SB) and highly obscured accretion, whose clear detection has
represented a long-standing challenge since the discovery of these objects (e.g. Genzel et al. 1998; Laurent et al. 2000).
Therefore, two crucial points are to disentangle the AGN and SB
components and to assess their contribution to the luminosity of both the individual sources and the local ULIRG population as a whole.
These tasks are fundamental not only in order to understand the nature of local ULIRGs themselves, but also to shed light on the interplay
and mutual feedback between star formation and BH accretion, which are basic ingredients of galaxy formation and evolution. \\
Taking advantage of the unprecedented data quality provided by the Infrared Spectrograph (IRS; Houck et al. 2004) onboard the
\textit{Spitzer Space Telescope} (Werner et al. 2004) we have recently developed an AGN/SB decomposition method for local ULIRGs
based on 5--8~$\mu$m rest-frame spectroscopy (Nardini et al. 2008). The key physical motivation that prompts the selection of this narrow
wavelength range for our analysis is the enhancement of the intrinsic AGN over SB brightness ratio (by a factor of $\sim$10--50; Nardini
et al. 2009, 2010) for equal bolometric luminosity, due to the intense emission from the hot dust layers exposed directly to the AGN.
This enables the detection of even faint or heavily obscured AGN components that are missed at different wavelengths.
Our spectral decomposition relies on a twofold observational evidence: at the highest luminosities the 5--8~$\mu$m average spectra of SB-
and AGN-dominated sources are clearly different, and their dispersion is very limited (Brandl et al. 2006; Netzer et al. 2007). It is therefore
possible to reproduce the AGN and SB contributions with fixed templates, only allowing for variable screen-like absorption of the AGN hot
dust continuum occurring in the colder layers of the putative torus and/or in some dusty circumnuclear cloud.\footnote{A power-law extinction
$\tau(\lambda) \propto \lambda^{-1.75}$ has been assumed (see also Draine 2003, and references therein).} Our main results can be summarized
as follows: \\
1) The AGN detection rate among local ULIRGs is $\sim$70 per cent, comparable to the amount achieved by collecting all the most
effective multiwavelength diagnostics employed so far. Remarkably, this also matches the fraction of submillimetre galaxies (Alexander et al. 2005)
and 70 $\mu$m-selected ULIRGs (Kartaltepe et al. 2010) hosting powerful AGN activity. However, 2) the main energy supply to the bolometric
luminosity of such objects comes from star formation, with an average AGN/SB power balance of $\sim$1/3. This ratio is confirmed to be
luminosity-dependent, since the typical AGN contribution rises from $\sim$10 to $\sim$60 per cent across the ULIRG luminosity range.
Consequently, the most intriguing outcome of our mid-IR analysis is 3) the identification in several sources of elusive AGN components missed
by the standard optical diagnostic tools. In particular, we find that the sources characterized by a steep continuum (which in our model is interpreted
as an effect of reddening), with deep absorption features and/or suppression of the polycyclic aromatic hydrocarbon (PAH) tracers of star
formation, do not usually show the optical line ratios of active galaxies, and occasionally are classified as pure H~\textsc{ii} regions. \\
This subclass encompasses $\sim$10--15 per cent of the local ULIRG population. Once corrected for extinction, some of these elusive AGN
components account for most of the galaxy IR luminosity, falling in the quasar -- rather than in the Seyfert -- range. As a consequence, they
represent the ideal targets for follow-up observations in the hard X-ray band, which are relevant to the quest for the most luminous, obscured
AGN (i.e. type 2 quasars) in the local Universe, and can also give deeper insight into the physical properties and geometrical structure of the
obscuring material in a ULIRG-like enviroment. \\
In this work we present the X-ray analysis of a sample of ten optically-elusive and IR-obscured AGN with quasar-like luminosity, which span a
large range with respect to the relative AGN/SB contribution, the degree of AGN obscuration and the overall IR luminosity, and are therefore broadly
representative of their parent class. New \textit{Chandra} and \textit{XMM-Newton} observations have been obtained for four of these objects,
i.e. IRAS~00397$-$1312, IRAS~01003$-$2238, IRAS~01298$-$0744 and IRAS~12127$-$1412, while archival X-ray spectra are available
for the other six buried AGN completing our sample. All the sources at issue are introduced in Section~2, with the log of the X-ray observations
and a brief description of the data reduction. In Section~3 we discuss both the existing constraints and our new results on the sources of our
X-ray sample, combining all these indications with our previous mid-IR analysis and interpreting them in the wider context of the connection
between circumnuclear star formation and BH growth in Section~4. The conclusions are drawn in Section~5. Throughout this paper distances
and luminosities have been computed assuming a $\Lambda$CDM cosmology based on the latest measures of the \textit{Wilkinson Microwave
Anisotropy Probe} ($H_0=70.5$~km~s$^{-1}$~Mpc$^{-1}$, $\Omega_\rmn{m}=0.27$ and $\Omega_\Lambda=0.73$; Hinshaw et al. 2009).
\begin{figure}
\includegraphics[width=8.5cm]{f1.eps}
\caption{The $\sim$2--14~$\mu$m rest-frame spectra (taken by \textit{Akari} and \textit{Spitzer}, and representative of the whole sample under study)
of the four sources newly observed in the X-ray band. For the sake of comparison, we also show the spectrum of a typical SB-dominated ULIRG,
IRAS~14060+2919. Each of our X-ray targets shows very faint or absent PAH emission (whose features are expected at 3.3, 6.2, 7.7, 8.6 and
11.3~$\mu$m) and broad formless and/or narrow deep absorption (due to e.g. water ice at $\sim$3.1 and 6.0~$\mu$m, carbon monoxide at
4.65~$\mu$m, hydrogenated amorphous carbons at 6.85 and 7.25~$\mu$m, silicates at 9.7~$\mu$m). These are all signatures of the unambiguous
presence of a buried compact engine and are not consistent with a SB-dominated environment. The shaded region indicates the interval used for our
AGN/SB decomposition analysis. (The \textit{Akari} spectra, courtesy of M.~Imanishi, have been rebinned for display purpose.)}
\label{f1}
\end{figure}
\begin{table*}
\caption{General properties, IR spectral parameters and X-ray observation log of the sources selected for our X-ray analysis.
(1)~\textit{IRAS} name, (2)~redshift, (3)~IR luminosity in log of solar units, (4)~optical classification, (5)~1$\sigma$ confidence
range for the AGN bolometric contribution (in per cent) from our mid-IR analysis, (6)~AGN optical depth at 6~$\mu$m, (7)~X-ray
satellite and observation identifier, (8)~observation date, (9)~net exposure in ks, (10)~references.}
\label{t1}
\begin{tabular}{cccccccccc}
\hline
Object & $z$ & $L_\rmn{IR}$ & Type & $\alpha_\rmn{bol}$ & $\tau_6$ & ObsID & Date & $T$ & Ref. \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) & (10) \\
\hline
00091$-$0738 & 0.118 & 12.27$\pm$0.03 & H~\textsc{ii} & 52--64 & 2.30 & \textit{Chandra}--10342 & 2008/11/01 & 15 & 4 \\
00182$-$7112 & 0.327 & 12.95$\pm$0.04 & LINER & 96--99 & $-$ & \textit{XMM}--0147570101 & 2003/04/16 & 10 & 2 \\
00397$-$1312 & 0.262 & 12.95$\pm$0.08 & H~\textsc{ii} & 50--58 & 0.24 & \textit{Chandra}--11532 & 2009/08/23 & 44 & 1 \\
01003$-$2238 & 0.118 & 12.33$\pm$0.06 & H~\textsc{ii} & 46--54 & 1.58 & \textit{XMM}--0605880101 & 2009/12/08 & 28 & 1 \\
01166$-$0844 & 0.118 & 12.12$\pm$0.06 & H~\textsc{ii} & 78--94 & 2.35 & \textit{Chandra}--10344 & 2008/10/31 & 15 & 1 \\
01298$-$0744 & 0.136 & 12.36$\pm$0.05 & H~\textsc{ii} & 71--80 & 1.79 & \textit{Chandra}--11531 & 2009/10/21 & 21 & 1 \\
07251$-$0248 & 0.088 & 12.41$\pm$0.04 & $-$ & 35--43 & 1.37 & \textit{Chandra}--7804 & 2006/12/01 & 15 & 1 \\
08572+3915 & 0.058 & 12.15$\pm$0.01 & LINER & 83--88 & 0.44 & \textit{Chandra}--6862 & 2006/01/26 & 14 & 3 \\
11095$-$0738 & 0.107 & 12.28$\pm$0.03 & LINER & 56--64 & 1.56 & \textit{Chandra}--10347 & 2009/04/09 & 16 & 4 \\
12127$-$1412 & 0.133 & 12.20$\pm$0.06 & LINER & 85--91 & $-$ & \textit{Chandra}--11533 & 2010/04/30 & 35 & 1 \\
\hline
\end{tabular}
\begin{flushleft}
\textit{References:}~$^1$This work; $^2$Nandra \& Iwasawa (2007); $^3$Teng et al. (2009); $^4$Teng \& Veilleux(2010).
\end{flushleft}
\end{table*}
\begin{table*}
\caption{Summary of the results of our X-ray spectral analysis. }
\label{t2}
\begin{tabular}{ccccc}
\hline
Object & $F_\rmn{X}^\rmn{AGN}$ & $F_\rmn{X}^\rmn{SB}$ & $F_\rmn{X}^\rmn{obs}$ & $\epsilon_\rmn{refl}$ \\
\hline
00091$-$0738 & 5.8 & 0.023 & $<0.003$ & $<0.05$ \\
00182$-$7112 & 2.4 & 0.011 & 0.150 & 6.0 \\
00397$-$1312 & 3.3 & 0.017 & 0.007 & $<0.3$ \\
01003$-$2238 & 7.2 & 0.019 & 0.005 & $<0.2$ \\
01166$-$0844 & 4.8 & 0.015 & $<0.001$ & $<0.02$ \\
01298$-$0744 & 6.8 & 0.021 & $<0.002$ & $<0.03$ \\
07251$-$0248 & 7.4 & 0.060 & $<0.002$ & $<0.03$ \\
08572+3915 & 33 & 0.060 & $<0.005$ & $<0.02$ \\
11095$-$0738 & 8.4 & 0.028 & 0.008 & $<0.1$ \\
12127$-$1412 & 4.1 & 0.013 & 0.040 & 1.0 \\
\hline
\end{tabular}
\begin{flushleft}
\textit{Note.} $F_\rmn{X}^\rmn{AGN}$: intrinsic 2--10~keV AGN flux estimated from our mid-IR analysis, assuming standard SEDs and UV
to X-ray corrections as described in the text; $F_\rmn{X}^\rmn{SB}$: upper limit to the 2--10~keV SB flux based on the average far-IR to X-ray
relation for star-forming galaxies; $F_\rmn{X}^\rmn{obs}$: observed 2--10~keV flux; $\epsilon_\rmn{refl}$: rough estimate of the AGN reflection
efficiency (in per cent). All the fluxes are in units of 10$^{-12}$~erg s$^{-1}$ cm$^{-2}$.
\end{flushleft}
\end{table*}
\section{Observations and Data Reduction}
In our previous work (Nardini et al. 2010) we have examined the 5--8~$\mu$m \textit{Spitzer}-IRS spectra of 164 local ULIRGs
at redshift $z \la 0.35$, separating the AGN and SB components and providing a quantitative estimate of their contribution to the bolometric
luminosity of each galaxy. This can be achieved by means of a straightforward analytical model, once the AGN emission has been corrected
for cold dust extinction by using the reddening of the hot dust template as a measure of the optical depth (Nardini et al. 2008). It turns out that the
optical classification is on average in good agreement with the findings of our mid-IR diagnostics. None the less, among the 20 ULIRGs
(out of 164) for which we assess an AGN bolometric contribution ($\alpha_\rmn{bol}$) larger than 25 per cent and an optical depth to the
AGN ($\tau_6$) larger than one at 6~$\mu$m, only three objects are optically classified as type~2 Seyferts. All the remaining entries are
classified as either low-ionization nuclear emission-line regions (LINERs) or even H~\textsc{ii} regions, revealing that a substantial fraction
of sources harbour significant but highly obscured nuclear activity. \\
We have therefore selected the optimal candidates for a supplementary spectral study of this population in the X-rays according to the
following criteria, which are applied to the whole IR sample: 1)~availability of a meaningful (i.e. $>5$~ks exposure) X-ray
observation;\footnote{The nature of the sources discussed in this work practically restricts the archival search to the present-day
X-ray observatories, with large effective area at $\sim$2--10~keV. Even in this case, in fact, a non-detection in a very short snapshot
would have a non-univocal interpretation.} 2)~estimated AGN luminosity exceeding $\sim$10$^{12} L_{\sun}$; 3)~non-Seyfert
optical classification (e.g. Veilleux, Kim \& Sanders 1999). In particular, the second requirement removes any ambiguity in the AGN
identification, since the nuclear component is always strong enough to dominate the 5--8~$\mu$m emission. The corresponding
selection consists of six objects for which X-ray observations are already available, even if only in four cases the
results have been published. Four additional sources, i.e. IRAS~00397$-$1312, IRAS~01003$-$2238, IRAS~01298$-$0744 and
IRAS~12127$-$1412, all meeting the latter two criteria, have been obtained as a part of our X-ray follow-up campaign. They were
chosen to complete the present sample, and for their remarkable expected X-ray properties: \\
\textit{(a)}~By assuming a standard quasar spectral energy distribution (SED; Elvis et al. 1994), suggesting that
$\nu L_\nu$(2500~\AA)~$\simeq$~2$\times \nu L_\nu$(6~$\mu$m), and the most recent relations between the UV and X-ray luminosities
(Lusso et al. 2010), the intrinsic 2--10~keV flux of the AGN component in IRAS~01003$-$2238 and IRAS~01298$-$0744 is expected to be
$\sim 7 \times 10^{-12}$~erg s$^{-1}$ cm$^{-2}$, ensuring a prominent detection in case of a Compton-thin column density
($N_\rmn{H} \la 5 \times 10^{23}$~cm$^{-2}$). Even in a Compton-thick prospect, allowing for a reflection efficiency of a few per cent makes
the AGN clearly detectable with the current X-ray observatories. \\
\textit{(b)}~IRAS~00397$-$1312 and IRAS~12127$-$1412 are very interesting in this context as well (see also Imanishi et al. 2008, 2010). The former
is optically classified as H~\textsc{ii} region in spite of being the most luminous ULIRG in the 1~Jy sample (Kim \& Sanders 1998), with
$L_\rmn{IR} \simeq 10^{13} L_{\sun}$: its mid-IR spectrum shows faint PAH features (Fig.~\ref{f1}), and clearly hints at the simultaneous
presence of a buried AGN roughly comparable to the SB in terms of its IR energy output. The latter is found instead near the lower end of the
ULIRG luminosity range, but its heavily absorbed continuum lacking any star formation signature implies an AGN-dominated nature. For these
two sources the predicted X-ray flux is just slightly less than the estimate given above (by a factor of $\sim$2). \\
The four $\sim$2--14~$\mu$m rest-frame spectra are fully representative of this ULIRG subclass, and are shown in Fig.~\ref{f1} in contrast with
a typical SB template, while the general information concerning all the sources and their physical parameters as derived from our mid-IR analysis
are listed in Table~\ref{t1}. Considering all the entries, this X-ray sample allows us to probe a wide range of properties with respect to obscured
activity inside ULIRGs. It is also worth emphasizing that the 2--10~keV flux obtained by ascribing all the far-IR emission\footnote{We assume
$F_\rmn{FIR} = 1.26 \times 10^{-11}(2.58 f_{60}+ f_{100})$~erg s$^{-1}$ cm$^{-2}$, where the \textit{IRAS} flux densities $f_{60}$ and $f_{100}$ are in
Jy (Helou, Soifer \& Rowan-Robinson 1985).} to the SB component and applying the far-IR to X-ray correlation of Ranalli, Comastri \& Setti (2003) lies in
the range $\sim$1--$6 \times 10^{-14}$~erg s$^{-1}$ cm$^{-2}$, which is on average $\sim$300 times\footnote{We note that our absorption correction of
the AGN flux is $\sim$10 at most ($\tau_6 \simeq 2.35$ in IRAS~01166$-$0844), hence it does not significantly affect the contrast between the AGN and
SB expected X-ray brightness.} lower than what envisaged for the AGN (see Table~\ref{t2}). The median 2--10~keV luminosity predicted for the AGN
component is $L_\rmn{X} \sim 2.5 \times 10^{44}$~erg s$^{-1}$. This is actually a conservative evaluation: if we adopt the direct 6~$\mu$m to 2--10~keV
correction found by Lutz et al. (2004a), it should be revised upwards by a factor of $\sim$2. This places our sources well above the X-ray intrinsic luminosity
of the local population of Compton-thick AGN recently identified by Goulding et al. (2011), allowing us to explore the parameter space of highly obscured AGN
in the low-redshift/high-luminosity region. \\
Concerning the unpublished X-ray data, IRAS~01003$-$2238 was observed by \textit{XMM-Newton} on 2009 December 08 for 35~ks. The EPIC pn and
MOS cameras operated in full frame mode. After the filtering of high-background periods, the useful pn exposure is $\sim$28~ks. Incidentally, a previous
snapshot ($\sim$10~ks) taken with \textit{Chandra} in 2003 yielded only 20 counts at 0.5--8.0~keV, preventing a detailed spectral modelling (Teng et al. 2005).
The other five sources, instead, were observed with the \textit{Chandra} ACIS-S detector. The basic details about all the observations are summarized in
Table~\ref{t1}. The data reduction was performed following the standard procedures, using the latest versions of the \textsc{ciao} and \textsc{sas} packages
for \textit{Chandra} and \textit{XMM-Newton} event files, respectively. The source and background spectra were extracted from circular regions with radius of
2\arcsec (\textit{Chandra}) and 30\arcsec (\textit{XMM-Newton}). In each case the X-ray emission from the source is entirely collected within the chosen
aperture, corresponding to $\sim$3--10~kpc in the \textit{Chandra} images. No clear morphological structure can be appreciated. The analysis has
been performed on unbinned spectra using the \textsc{xspec} v12.5 fitting package, and $C$-statistic has been adopted due to the low number of
counts. Quoted uncertainties and upper limits are given at the 90 per cent confidence level.
\section{Spectral Analysis and Discussion}
Since ULIRGs are in general very faint X-ray sources, only a limited number of objects in our local IR sample have been targeted
with the current X-ray observatories \textit{Chandra} (Ptak et al. 2003; Teng et al. 2005), \textit{XMM-Newton} (Franceschini et al.
2003) and \textit{Suzaku} (Teng et al. 2009). A compilation of all publicly available X-ray data for 40 ULIRGs has been recently
discussed by Teng \& Veilleux (2010). \\
Analogously to their mid-IR spectra, the X-ray spectra of ULIRGs usually show the combined signatures of AGN and SB activity; in
particular, the soft ($E < 2$~keV) emission is characterized by a diffuse thermal component associated with star formation in the
circumnuclear regions, while at higher energies the hard AGN power law is transmitted through typical column densities of the order
of $\sim$10$^{23}$~cm$^{-2}$. For Compton-thick sources a flat reflection spectrum is expected, alongside the prominent iron K line
due to fluorescent emission in the surrounding optically-thick and almost neutral gas (e.g. Matt, Brandt \& Fabian 1996). In this section
we first analyse our new observations and then review the existing X-ray spectra of the other elusive AGN in our sample, in order to
explore the properties of the obscured AGN population within local ULIRGs and obtain some useful hints about and their high-redshift
equivalents.
\subsection{IRAS~00397$-$1312}
As mentioned above, IRAS~00397$-$1312 is one of the most striking objects in the local Universe for many reasons. Owing to its
huge luminosity and composite nature, it is an ideal source for a case study of the AGN/SB connection in extreme environments. This
ULIRG does not show clear morphological indications of a recent interaction, and likely represents an old merger stage (Veilleux, Kim \&
Sanders 2002). Despite its optical classification as H~\textsc{ii} region, the mid-IR spectral shape reveals all the unambiguous
signatures of a buried AGN (Imanishi 2009), first and foremost the outstanding CO absorption profile around $\sim$4.65~$\mu$m.
The X-ray spectrum of IRAS~00397$-$1312 is shown in Fig.~\ref{f2}: the only detected X-ray emission is well reproduced by means
of an absorbed power law, resulting in a $C$-stat of 45.0 for 52 degrees of freedom (d.o.f.). Besides the Galactic column density along the
line of sight towards the target, which has been taken into account in all our fits, an additional X-ray absorber with
$N_\rmn{H} \simeq 4.5 \times 10^{21}$~cm$^{-2}$ local to the source at $z=0.262$ is required. The unabsorbed 0.5--2 and 2--10~keV
fluxes are, respectively, $\sim$9 and $7 \times 10^{-15}$~erg s$^{-1}$ cm$^{-2}$, and are both in good agreement with the SB
X-ray flux predicted from our mid-IR analysis and the Ranalli et al. (2003) relations for star-forming galaxies (Table~\ref{t2}). The steepness
of the power law, whose photon index is $\Gamma \simeq 2.3^{+1.1}_{-0.6}$, is itself consistent with pure SB emission with no AGN contribution;
for the latter, an upper limit has been evaluated by adding a \textsc{pexrav} component (Magdziarz \& Zdziarski 1995) to account for possible
AGN reflection. All the geometrical and abundance parameters have been frozen to their nominal \textsc{pexrav} values, as well as the
photon index of the illuminating AGN power law (assumed to be 2.0). The upper limit to the observed 2--10~keV AGN flux is
$\sim 7.6 \times 10^{-15}$~erg s$^{-1}$ cm$^{-2}$, implying a reflection efficiency $< 0.3$ per cent (Table~\ref{t2}), far below the usual values
of $\sim$1--5 per cent. We will discuss this point further in the following. We note that the inclusion of a \textsc{mekal} component (i.e. emission
from hot diffuse gas) improves the fit at soft energies, but makes the hard X-ray parameters badly constrained. Given the low quality of our
data, we favour a simpler model even if the statistical significance is slightly lower.
\begin{figure}
\includegraphics[width=8.5cm]{f2.eps}
\caption{\textit{Chandra} spectrum of IRAS~00397$-$1312 reproduced with a \textsc{tbabs*zwabs*powerlaw} model. In these figure
and the following ones, the X-ray data are rebinned for plotting purposes.}
\label{f2}
\end{figure}
\subsection{IRAS~01003$-$2238}
This source is usually classified as H~\textsc{ii} region in the optical (Veilleux et al. 1999), yet there are also some conflicting claims
about its possible Seyfert~2 nature (Allen et al. 1991; Yuan, Kewley \& Sanders 2010). Although detected, absorption features
in the mid-IR are not dramatic in both depth and shape, and PAH emission is clearly distinguished on top of a very steep
continuum. The X-ray spectrum is described with a power law plus a \textsc{mekal} component, yielding a $C$-stat/d.o.f.$= 21.3/35$.
The spectrum is very soft (Fig.~\ref{f3}), and is characterized by a \textsc{mekal} temperature of $kT = 0.66^{+0.31}_{-0.21}$~keV and
a power law photon index of $2.9^{+0.4}_{-0.3}$, suggesting again a sole SB origin. The observed flux is broadly consistent with such
interpretation ($\sim 2 \times 10^{-14}$ and $5 \times 10^{-15}$~erg s$^{-1}$ cm$^{-2}$ in the 0.5--2 and 2--10~keV energy range).
An upper limit to the 2--10~keV AGN contribution has been computed as above, resulting in $\sim 1.3 \times 10^{-14}$~erg s$^{-1}$
cm$^{-2}$. Hence, similar considerations can be drawn in this case: no direct AGN emission is detected below 10~keV, hinting at a
Compton-thick circumnuclear environment. The lack of any AGN signature even in the form of a reflected spectrum implies a complete
covering of the X-ray absorber.
\begin{figure}
\includegraphics[width=8.5cm]{f3.eps}
\caption{\textit{XMM-Newton} spectrum of IRAS~01003$-$2238 and best fit obtained according to a \textsc{tbabs*(mekal+powerlaw)}
model. Only the EPIC pn data have been plotted.}
\label{f3}
\end{figure}
\subsection{IRAS~01298$-$0744}
This is undoubtedly the most puzzling object of the present selection, since the outcome of our X-ray observation is at
first a striking non detection, as also confirmed with the \textsc{ciao} `wavdetect' tool. A serendipitous unidentified
faint source is found $\sim$15\arcsec ~away from the ULIRG
position, but any physical relation between the two can be ruled out since this angular distance is equivalent to over
30~kpc at the redshift of IRAS~01298$-$0744, which is actually an advanced merger with only residual tidal tails. By
considering the background count rate, we can provide an upper limit to the observed 2--10~keV source excess
of $\sim 2 \times 10^{-15}$~erg s$^{-1}$ cm$^{-2}$, corresponding to a marginal detection with 2$\sigma$ confidence level.
Such a limit is significantly lower than the related entries of Table~\ref{t2}, and yet reconcilable with a SB component with
a far-IR luminosity of $\sim 2 \times 10^{11} L_{\sun}$, that would account for $\sim$10 per cent of the overall IR emission.
Ascribing instead the upper limit to an AGN with an intrinsic X-ray flux as that predicted, the resulting column density has to
be larger than $\sim 8 \times 10^{24}$~cm$^{-2}$. As discussed in the following, this case is not unique in its class.
\subsection{IRAS~12127$-$1412}
IRAS~12127$-$1412 is one of the few objects in our large IR sample whose 5--8~$\mu$m spectral shape cannot be
reproduced by means of the standard AGN and SB templates. In fact, the observed emission is characterized by a
broad formless absorption trough affecting most of the range of interest, and the putative continuum is slightly flatter than
the assumed AGN hot dust component. This prevents us from measuring $\tau_6$ from the reddening, and yet the AGN
obscuration is likely to be substantial: first, because a stepwise correlation between the continuum reddening and the
presence of individual absorption features is reasonably expected (e.g. Sani et al. 2008; Nardini et al. 2009); second,
because the steeper trend seen shortwards of 5~$\mu$m (Fig.~\ref{f1}) hints at a change in the slope of the extinction law
(e.g. Nishiyama et al. 2009) rather than at vanishing obscuration. Based on these considerations, the intrinsic X-ray flux
quoted in Section~2 for the AGN component of IRAS~12127$-$1412 can even be underestimated. It is therefore remarkable
that the only clear AGN detection among our new observations unfolds in this source (Fig.~\ref{f4}). Indeed, the X-ray
spectrum is quite complex: a steep power-law ($\Gamma \simeq 3.0 \pm 1.5$) or thermal component is needed at soft
energies, while the hard emission can be modelled with a prominent iron feature at $\sim 6.77^{+0.50}_{-0.33}$~keV on
top of an absorbed ($N_\rmn{H} \simeq 9^{+17}_{-5} \times 10^{22}$~cm$^{-2}$) continuum. All the parameters cannot be
adequately constrained, but the characteristic flatness of a cold reflection spectrum is not observed, since the photon index
of the AGN component is $\sim 2.0^{+1.1}_{-0.4}$. Also, the line energy is more consistent with He-like iron than
with cold neutral material, pointing to ionized reflection. Unfortunately, we are not able to discuss this uncommon but intriguing
scenario in detail. We simply note that such a possibility has already been suggested for IRAS~00182$-$7112, which is
included in our X-ray sample as well and is remarkably similar to IRAS~12127$-$1412 both in the mid-IR and in the X-rays
(Spoon et al. 2004; Nandra \& Iwasawa 2007), as described below. Returning to the source at issue, we estimate an
equivalent width for the K-shell feature of $\sim$1.3~keV, with a lower limit that can be roughly placed at 0.4~keV.\footnote{A
rigorous calculation of the line parameters is precluded by the vanishing continuum at higher energies.} The observed flux in
the 0.5--2 and 2--10~keV range is $\sim 2 \times 10^{-15}$ and $4 \times 10^{-14}$~erg s$^{-1}$ cm$^{-2}$, and adds
further evidence to the presence of an obscured AGN with a reflection efficiency of $\sim$1 per cent.
\begin{figure}
\includegraphics[width=8.5cm]{f4.eps}
\caption{The complex X-ray emission of IRAS~12127$-$1412, modelled as \textsc{tbabs*[powerlaw+zwabs*(zgauss+powerlaw)]}
with $C$-stat/d.o.f.$= 48.0/47$. The observed 2--10~keV flux and the detection of a prominent iron line clearly point to AGN reflection.}
\label{f4}
\end{figure}
\subsection{IRAS~01166$-$0844 and IRAS~07251$-$0248}
Two other objects within our selection of ULIRGs that are missed by the optical diagnostics as Seyfert galaxies but likely harbour
an AGN with quasar-like luminosity, although observed in the X-rays, have never been published to date. We have therefore reduced
as outlined above and then analysed the $\sim$15~ks long observations of IRAS~01166$-$0844 and IRAS~07251$-$0248
retrieved from the \textit{Chandra} archives. The former source is a close equivalent (although much fainter) to IRAS~01298$-$0744
at 5--8~$\mu$m; strikingly, it turns out to be completely undetected as well. The latter, instead, which is a very bright source in the \textit{IRAS}
Revised Bright Galaxy Sample ($z \simeq 0.088$, $f_{60} \simeq 6.5$~Jy; Sanders et al. 2003) delivers a 3.5$\sigma$ detection, and its
weak emission can be fitted with a steep power law of photon index $\simeq$3.6($\pm1.6$). In both cases, the upper limit (or estimated)
2--10~keV flux is of the order of $\sim$10$^{-15}$~erg s$^{-1}$ cm$^{-2}$. Assuming again the far-IR to X-ray conversion of Ranalli et al.
(2003), such a tiny value implies not only that the AGN is thoroughly absorbed, but also that the SB provides very little contribution to the
bolometric luminosity of the host galaxy.
\subsection{The whole X-ray sample}
The remaining four sources in our sample similarly display strong evidence for buried nuclear activity at mid-IR wavelengths, with spectral
trends that are well represented by the ones shown in Fig.~\ref{f1}. At high energies, IRAS~00091$-$0738 and IRAS~11095$-$0238 have
been classified as \textit{weak} X-ray ULIRGs by Teng \& Veilleux (2010): the small count rate does not allow traditional spectral fitting, and
the only possible analysis is based on hardness ratio methods. We have reviewed these $\sim$15~ks \textit{Chandra} observations to obtain
a measure of the received 2--10~keV flux, and found that it is consistent with a SB origin of only less than $\sim$15 and 30~per cent, respectively,
of the far-IR emission (see Table~\ref{t2}), assuming that no X-ray contribution from the AGN is detected. \\
Meaningful results have been obtained, instead, for the other two sources, which are definitely among the most extreme ULIRGs.
IRAS~00182$-$7112, as already mentioned, is shown to harbour a Compton-thick AGN, with flat continuum and prominent Fe~\textsc{xxv}
K$\alpha$ emission; the reflection component itself is likely to be transmitted through a screen with $N_\rmn{H} \sim 10^{23}$~cm$^{-2}$
(Nandra \& Iwasawa 2007). IRAS~08572+3915, instead, has been investigated with all the current X-ray observatories: it exclusively reveals
a very marginal detection with \textit{Chandra}, but no high-energy emission is seen with \textit{Suzaku} (Teng et al. 2009). CO measurements
imply for this source a gas column up to $\sim 3 \times 10^{25}$~cm$^{-2}$ (Evans et al. 2002). As a consequence, both these galaxies can
be considered as a template type~2 quasar in the local Universe. \\
Summarizing, the low number of counts collected from the elusive AGN in our present X-ray ULIRG sample is compatible with the Compton-thickness
of the putative X-ray absorber, as well as the detection of iron lines with huge equivalent width, resembling the other notable case of Arp~220
(Iwasawa et al. 2005; Downes \& Eckart 2007). In other words, virtually all ULIRGs with unambiguous mid-IR signatures of powerful, buried
AGN activity are extinguished in the X-rays, with high incidence of Compton-thick objects and a wide gamut of reflection efficiencies, ranging from
values expected for a standard type 2 obscuration geometry to much lower ones indicative of full and/or complex covering.
\section{Discussion}
The X-ray observations analysed in this work provide further insight into the physical and geometrical properties of the gaseous
material surrounding the AGN and responsible for its obscuration in a ULIRG environment. Assuming the validity of the Compton-thick
scenario rather than invoking extremely unusual IR to X-ray corrections and/or X-ray reflection efficiencies, the missing AGN detections
below 10~keV have two main implications which are tightly linked with each other, concerning the covering factor of the X-ray
absorber and the coupling of the gas and dust components in the outskirts of the compact active source. Also, we discuss our results
in the context of the AGN/SB connection, searching for possible relations between the excess of AGN obscuration and the strenght of
the simultaneous burst of star formation.
\subsection{The nature of the X-ray absorber}
As pointed out earlier, the upper limits that can be placed on the possible AGN reflected emission imply a complete covering of the nuclear
region. In the X-ray obscured AGN that are classified in the optical as type 2s, the direct optical to soft X-ray radiation blocked by the claimed
axisymmetric dusty absorber can be scattered into the line of sight after the interactions with material located above the equatorial plane. This
process strongly affects the shape of the emerging spectrum in the X-rays, and the usual efficiency in terms of the received flux is of the order
of a few per cent, depending on geometrical effects. Except for the two AGN detections, in all the other sources of our sample we assess a
reflection efficiency lower than $\sim 2 \times 10^{-3}$, assuming an intrinsic 2--10~keV flux consistent with the results of the mid-IR analysis.
Thus, regardless of the large and manifold uncertainties involved in the single cases, the X-ray follow-up observations of buried AGN presented
or reviewed here generally hint at a cocoon-like geometry of the absorber, which is not surprising in extreme and chaotic systems such as
ULIRGs. \\
Indeed, the detection of so prominent CO gas absorption in objects like IRAS~00182$-$7112 (Spoon et al. 2004) and IRAS~00397$-$1312
could be related with full covering rather than toroidal obscuration, since this feature is not usually found among type 2 Seyferts (e.g. Lutz et al.
2004b). As a consequence, the most interesting aspect to investigate concerns the location of the absorber and the properties of its gas and
dust content. By adopting a typical Galactic extinction curve (e.g. Draine 2003; Nishiyama 2008, 2009) and the standard Galactic dust-to-gas
ratio (Bohlin, Savage \& Drake 1978), one can easily obtain the relation between the optical depth at 6~$\mu$m and the X-ray column density,
that is $N_\rmn{H} \simeq 8$--11~$\tau_6 \times 10^{22}$~cm$^{-2}$. Then, even questioning the accuracy of our measure, any reasonable value
of the \textit{real} $\tau_6$ would not be consistent with a Compton-thick environment with dust and gas components similar to the Galactic diffuse
interstellar medium. The ensuing point is whether the X-ray column density is physically related to the dust affecting the optical and mid-IR diagnostics.
There is plenty of observational evidence indicating \textit{anomalous} dust properties in AGN for which different explanations have been proposed.
For instance, X-ray absorption may occur in the very inner regions, inside the dust sublimation radius (e.g. Granato, Danese \& Franceschini 1997);
conversely, the formation of large dust grains favoured by high density environments yields a flat extinction curve misleading about the actual dust-to-gas
ratio (Maiolino, Marconi \& Oliva 2001). \\
In ULIRGs the situation is even more complex: a remarkable example is represented by Mrk~231, a bright ULIRG which is optically classified as a
\textit{Broad Absorption Line} quasar, and whose fairly unabsorbed \textit{Spitzer}-IRS spectrum suggests a relative AGN contribution to the bolometric
luminosity of the galaxy of $\sim$30 per cent at least. The 2--10~keV emission of Mrk~231 reveals a reflection component only (Gallagher et al. 2002),
while the direct AGN continuum shows up at 15--60~keV in the \textit{BeppoSAX} data (Braito et al. 2004). This mismatch between the optical and
X-ray classification based on the obscuration degree in each band is not infrequent, and suggests a scarce gas/dust coupling around the AGN.
Moreover, the case of Mrk~231 is indicative of the incidence of outflows: massive gas inflows are needed to sustain nuclear accretion, yet the possible
link of the X-ray absorbing gas with AGN-driven winds and outflows has far-reaching implications in terms of the AGN feedback and the ULIRG/quasar
evolutionary pattern. According to this scenario, ULIRGs (or at least a significant fraction of them) are a transitory phase of galactic lifetime preliminary
to the optically-bright quasar stage, during which the merger-driven inflows provide the gas reservoir for starburst and buried nuclear activity, until the
AGN feedback disrupts the obscuring shell of dust and gas, quenching the circumnuclear star formation and sweeping the line of sight to the optical
quasar (Sanders et al. 1988; Hopkins et al. 2008). \\
AGN- or SB-driven galactic winds are known to play a key role in the evolution of galaxies and are found in most ULIRGs (Veilleux, Cecil \&
Bland-Hawthorn 2005, and references therein). Unambiguous evidence for gas outflows has been recently discovered in the mid-IR also in buried
AGN candidates. High-resolution \textit{Spitzer}-IRS spectra reveal the presence of high-velocity gas in IRAS~00182$-$7112, giving rise to a broad
component in the [Ne~\textsc{ii}] and [Ne~\textsc{iii}] lines at 12.81 and 15.56~$\mu$m (Spoon et al. 2009). In both cases the line profile is asymmetric
due to a blue wing. A similar behaviour is not observed in optical or near-IR forbidden lines, as a result of the large extinction. This blueshift is interpreted
as the signature of the ongoing breaching of the obscuring medium surrounding the active nucleus. Only two other sources exhibit such an asymmetry
in neon profiles: our target IRAS~12127$-$1412, and IRAS~13451+1232. The latter is also detected in [Ne~\textsc{v}] at 14.32~$\mu$m, which is a
straightforward tracer of the AGN radiation field due to the high ionization energy required (97.1~eV). This strongly blueshifted detection indicates a more
advanced stage of this source in the process of cleaning its line of sight, which is consistent with its type 2 Seyfert classification in the optical. The larger
part of ULIRGs with [Ne~\textsc{v}] emission show a blueshift, and the higher the neon ionization state the larger the blueshift, proving that the speed
of the outflow decreases with the distance from the ionizing core (Spoon \& Holt 2009). In some cases the gas kinematics traced by neon lines can be
associated with the expansion of radio jets. Anyway, out of 25 ULIRGs with either [Ne~\textsc{iii}] or [Ne~\textsc{v}] blueshifted lines, 21 are
optically-identified AGN.\footnote{Most of these sources are also included in our IR sample.} IRAS~00182$-$7112, IRAS~12127$-$1412 and
IRAS~01003$-$2238 (neglecting its controversial classification) are among the few exceptions. In this picture, mid-IR buried and X-ray Compton-thick
AGN are still fully enshrouded, and their feedback is not yet in action.
\subsection{AGN feedback and SB intensity}
On the wake of the considerations above, it is worth investigating the possible relation between the AGN mid-IR obscuration and the SB intensity.
Fierce star formation activity is ubiquitous in AGN both in the local Universe (Schweitzer et al. 2006) and at high redshift (Hatziminaoglou et al. 2010).
Moreover, far-IR observations suggest that X-ray absorbed AGN suffer a larger contamination from the concurrent star formation than
unabsorbed ones at the same luminosity and redshift (e.g. Stevens et al. 2005). In order to address this issue, we have plotted in Fig.~\ref{f5} the
luminosities of the AGN and SB components derived from our mid-IR spectral decomposition, with a code emphasizing the degree of AGN extinction
at 6~$\mu$m. The relative contribution is indicated by the diagonal lines, which follow the AGN \textit{weight} classification defined in Nardini et al. (2010).
Before drawing some qualitative considerations, we stress that the exact location of each ULIRG in this plot is affected by several sources of systematic
uncertainty. In fact, $\alpha_\rmn{bol}$ [here simply given by $(1+L_\rmn{SB}/L_\rmn{AGN})^{-1}$] is an absorption-corrected quantity
and was calculated as a function of the intrinsic AGN and SB bolometric corrections averaged over the whole sample, but their \textit{actual}
values for the single objects can be quite different.\footnote{The 1$\sigma$ dispersion inherent in our method is $\sim$0.18~dex (see Nardini et
al. 2009, 2010 for a review and a comprehensive discussion of our IR analysis).} Also, any deviation from the
assumed templates and extinction law will modify both $\tau_6$ and $\alpha_\rmn{bol}$, and the difference in the merger stage does not allow
a uniform interpretation. This notwithstanding, at the higher AGN luminosities (i.e. above $L_\rmn{AGN} \approx 10^{12} L_{\sun}$),
where all the entries of our X-ray sample lie and the mentioned
\textit{confusion} effects should be less important, the intensity of the concurrent SB activity seems to be loosely correlated to the AGN mid-IR
obscuration. Similarly, in terms of growing $\alpha_\rmn{bol}$, when the AGN is still obscured the SB can be very powerful, accounting for roughly
half of the total luminosity, but the production of stars subsides when the line of sight is clean. \\
At a more speculative level, we focus on the three ULIRGs that have an \textit{anomalous} 5--8~$\mu$m spectral shape with flat but deeply
absorbed continuum, and are clearly AGN-dominated in the mid-IR. These are the magenta stars in the top left-hand corner of Fig.~\ref{f5}, i.e.
IRAS~00182$-$7112, IRAS~12127$-$1412 and IRAS~12514+1027. It turns out that all of them are detected as reflected AGN emission in the
hard X-rays (see Wilman et al. 2003 for the latter object, not included in the present sample). This intriguing case hints at the possibility of
distinguishing through mid-IR analysis only between a Compton-thickness associated to a toroidal X-ray absorber and a Compton-thickness
due to a complete AGN covering, coeval with the peak of SB activity. In the former scenario, one can argue that the X-ray reflection signatures
arise from the inner edge of a slightly tilted torus with little opening angle; this would also imply direct access to a certain fraction of hot dust, possibly
resulting in a flattening of the continuum in agreement with the spectral trends of quasars around $\sim$5~$\mu$m (see Netzer et al. 2007, and our
discussion on the AGN template in Nardini et al. 2009). However, further observations are needed to test this conjecture and the chance, if any, to
infer the X-ray column density and covering factor exclusively from IR diagnostics; this would be relevant to the study of IR galaxies at high redshift
and to the measure of the contribution to the X-ray background from type 2 quasars. \\
Incidentally, we finally note that by extrapolating the fraction of absorbed AGN in Fig.~\ref{f5} below $\alpha_\rmn{bol}=0.05$, it looks possible that
almost every ULIRG in the local Universe is the seat of nuclear activity, and that also some of the pure SB/ULIRGs
(mostly falling outside the plotted region and representing $\sim$30 per cent of the total ULIRG population) harbour an AGN which is both
intrinsically faint and heavily obscured.\footnote{Such AGN components would be totally missed even with our method, because their contribution
to the observed mid-IR spectra is too small: in our scheme, these objects are classified as pure SB/ULIRGs.} In this perspective, the clear cut in
the presence of absorbed sources at $\alpha_\rmn{bol} < 0.05$ is simply due to selection effects: first, when the SB dominates at even
5--8~$\mu$m, the AGN optical depth can be understimated, and some of the green dots would turn into red circles with just a tiny displacement;
second, some of the non-detections would be shifted upwards, entering the plotted portion of the $\alpha_\rmn{bol} < 0.05$ region as either
red circles or blue crosses.
\begin{figure*}
\includegraphics[width=15cm]{f5.eps}
\caption{AGN versus SB luminosity from our mid-IR analysis. Different colours and symbols are used to distinguish the different
estimates of the AGN optical depth at 6~$\mu$m. The magenta stars represent those ULIRGs with flat but heavily absorbed
5--8~$\mu$m spectra, for which the continuum reddening cannot be used to measure the AGN optical depth. Interestingly, the
largest X-ray reflection efficiencies for buried AGN are found among these objects. The cyan curve describes the boundary
of our ULIRG sample, $L_\rmn{IR} = L_\rmn{AGN} + L_\rmn{SB} > 10^{12} L_{\sun}$ (all the sources meet this
condition within the error bars, not shown), while the dashed diagonal lines trace the four regions of the $\alpha_\rmn{bol}$ space
defined in Nardini et al. (2010). Most of the pure SB/ULIRGs lie outside the plot due to their low upper limit to
$L_\rmn{AGN}$. Conversely, when the AGN is dominant, the degree of AGN obscuration in the mid-IR seems to be loosely
related to the intensity of the concurrent star formation. Anyway, since each entry can be somewhat displaced due to the various
sources of uncertainty (see the text), this plot is only intended for qualitative considerations (consequently, purely statistical error
bars are omitted for clarity). The encircled objects are the ten ULIRGs presented or reviewed here.}
\label{f5}
\end{figure*}
\section{Conclusions}
Our recent mid-IR diagnostic campaign on local ULIRGs has revealed the presence in a sizable fraction of these sources of powerful
AGN components, providing a significant contribution to the total energy output but missed at optical wavelengths because of their heavy
obscuration. The mid-IR spectra of these intriguing sources provide strong evidence of buried nuclear activity, in the shape of a steep and
intense continuum, dramatic suppression of PAH features and various absorption. In order to pursue a more exhaustive study of the extreme
environment surrounding the AGN component inside this subclass of ULIRGs, we have presented the X-ray analysis of a representative
sample consisting of ten objects of this kind, and including our new follow-up X-ray observations obtained with \textit{Chandra} and
\textit{XMM-Newton} of IRAS~00397$-$1312, IRAS~01003$-$2238, IRAS~01298$-$0744 and IRAS~12127$-$1412. Only in
IRAS~12127$-$1412 the presence of the AGN is clearly established in the X-rays, while IRAS~01298$-$0744 is not detected at all. The
faint X-ray emission of IRAS~00397$-$1312 and IRAS~01003$-$2238 is consistent with a pure SB contribution, in terms of both spectral
shape and observed flux. The other six sources in our X-ray sample, two of which have not been published to date, fully validate this general
trend, with only another safe AGN detection in IRAS~00182$-$7112. \\
The main implications of this work are the following: 1) Assuming a typical dust extinction law, the dust-to-gas ratio is required to be much lower
than the Galactic standard in order to support the Compton-thick interpretation for the unsuccessful AGN detections. Of course, this is not due to
a large-scale dust deficiency, since ULIRGs are, by definition, dust-rich systems. Anyway, absorption features such as that of CO at 4.65~$\mu$m
observed in IRAS~00397$-$1312 hint at large amounts of gas concentrated nearby a compact active core. A possible imbalance between the dust
and gas components involving only the nuclear regions (i.e. within a region comparable in size with the dust sublimation radius) may be related to
massive gaseous inflows/outflows during the most active phases of AGN accretion. 2) Whenever the AGN is not detected but the amplitude of a
possible reflection component can be constrained, its upper limit is $>500$ times lower than the expected intrinsic 2--10~keV AGN flux. This is an
order of magnitude lower than the usual reflection efficiency observed in type 2 active galaxies, suggesting an almost complete AGN covering in these
ULIRGs. It is unlikely that the standard toroidal geometry of the dust absorber predicted for isolated and morphologically quiet AGN will hold
systematically also for much more complex systems like ULIRGs, whose geometrical structure can be poorly retraced. Again, a complete covering is
compatible with a spherical shell of gas infalling on to the accreting SMBH or, more conceivable in advanced merger stages, with the dust-free base
of the winds invoked in the feedback scenario to clean the line of sight to the AGN and, possibly, to quench star formation. \\
The elusive AGN identified through their mid-IR spectral properties, the most notable of which have been discussed in this work, are likely going
through the ultimate phase of buried activity, and are therefore the ideal sources to search for detectable signatures of the impending transition from
ULIRGs to optically bright quasars. Even if the quality of the single X-ray spectra is very modest, the availability of X-ray constraints is crucial to derive
a comprehensive multiwavelength picture. Extending the number of intrinsically luminous but heavily obscured AGN for which it is possible to infer the
amount and distribution of gas in the circumnuclear environment and to compare such information with the optical to mid-IR dust extinction and absorption
features and the intensity of the concurrent star formation may shed light on the ULIRG/quasar evolutionary sequence and the AGN feedback mechanisms
affecting the properties of the host galaxy. Moreover, if the tentative correlation between mid-IR spectral properties and X-ray column density and covering
factor is confirmed and clearly established, a simple \textit{qualitative} inspection of the mid-IR spectra of the high-redshift counterparts of local ULIRGs will
unveil Compton-thick AGN candidates and enable a measure of their contribution to the X-ray background.
\section*{Acknowledgments}
This research has been partially supported by NASA grants NNX09AT10G and GO0-11017X.
We thank the anonymous referee for his/her constructive comments which improved our work.
|
\section{Introduction and Results}
Quantization of Chern-Simons gauge field theory has garnered much interest in the last twenty years. There are several different ways that the theory can successfully be quantized, and they relate diverse areas in mathematics and physics. The path-integral approach~\cite{Witten:1988hf,Moore:1989yh} shed light on the relation between Chern-Simons theory, knot invariants and conformal field theory. Geometric quantization~\cite{Axelrod:1989xt} gave a general three-dimensional quantization. Canonical quantization can be performed using either a real polarization~\cite{Elitzur:1989nr}, or a complex polarization and coherent states~\cite{Bos:1989kn,Gukov:2004id}, and a general theory has been developed using quantum groups~\cite{Alekseev:1994pa,Alekseev:1995,Meusburger:2003wk}.
Chern-Simons is a topological field theory which does not depend on details of the geometry of the space-time on which it is defined, rather only the topological features. In particular, since its action is an integral of 3-form, it is invariant under diffeomorphisms of the underlying space-time. Like gauge transformations, diffeomorphisms split into two types, small diffeomorphisms which can be continually deformed to the identity and large diffeomorphisms which cannot. The quotient group of all the diffeomorphisms by small diffeomorphisms is called mapping class group (MCG). In Chern-Simons theory, small diffeomorphisms are equivalent with small gauge transformations on shell, and can be fixed at the classical level. On the other hand, MCG, as well as the large gauge transformations (LGT) group, are discrete symmetries of the theory and there is no requirement that quantum states should be invariant. Instead, they should carry representations of these groups. It is generally an interesting question to ask what representations are allowed. In this Paper, we shall address this question in the context of $\mathrm{U}(1)$ Chern-Simons theory.
The example of Chern-Simons theory with a $\mathrm{U}(1)$ gauge group is particularly tractable in that quantization can be done very explicitly. When the spacetime manifold is the product $M=\mathbb{R}\times\Sigma$, with $\mathbb{R}$ the time and $\Sigma$ a two-dimensional orientable surface, explicit wave-functions can be obtained~\cite{Bos:1989wa,Manoliu:1996fx} and shown to have an interesting relationship to rational conformal field theory.
In this paper, we shall revisit the quantization of $\mathrm{U}(1)$ Chern-Simons theory on closed orientable surfaces. Our goal is to examine role of the MCG of the surface in quantizing the theory. We find the following results. We shall demonstrate that the MCG is quantizable and we shall find its representation explicitly. If we seek quantization with a finite dimensional Hilbert space, we find that it is possible only when the parameter $k$ (in Eq.(\ref{ChernSimonsAction})) is a rational number. This result is known from previous works \cite{Polychronakos:1990xq,Dunne:1998qy}. In these works and in Ref.~\cite{Bos:1989wa}, when $k=p/q$ with $p$ and $q$ coprime, it was stated that $p$ (or $k$ in~\cite{Bos:1989wa}) must be an even integer. Our results differ qualitatively from these quantizations and depend on the genus. When the genus is one ($\Sigma$ is the 2-torus), $k$ can be any rational number. For higher genus, one of $p$ or $q$ must be even. Moreover, by incorporating MCG, we find that for a given $k$, the representations of the holonomy group and LGT group become unique, and the representation for MCG is also unique, apart from an arbitrary unitary sub-representation which acts on the holonomy group and LGT group trivially.
Generally, at the classical level, when $\Sigma$ has genus $g$, the group of large gauge transformations is $\mathbb{Z}^{2g}$. We find that, commensurate with results of Ref.~\cite{Bos:1989wa}, this discrete group is realized undeformed at the quantum level only when $k$ is an integer. However, even in that case, we find that, augmenting the quantization with the requirement that the Hilbert space carries a unitary representation of the MCG, restricts the representation of the large gauge transformations to those where states are strictly invariant (theta angles associated with large gauge transforms vanish). Furthermore, we shall show that, when $k$ is rational but not integer-valued, the discrete group of large gauge transformations, which was abelian at the classical level, obtains a 2-cocycle and becomes a clock algebra~\cite{Polychronakos:1990xq,Dunne:1998qy}. We find an interesting $k\leftrightarrow 1/k$ duality of the representations of the homology group and the large gauge group which, with the restrictions on $k$ stated above, is compatible with our quantization of the MCG.
The reader might wonder why, in a topological field theory, where the action does not depend on the metric, the quantization of the MCG could be an issue at all. In order to do canonical quantization, we must choose a set canonical variables and, to quantize, we must further choose a polarization. It is the latter which is not generally covariant. Then, covariance needs to be restored by quantizing the MCG. As we shall show, the quantization of the MCG is non-trivial and, as we have discussed above, it can only be carried out with some restrictions on $k$ and even then it poses restrictions on certain parameters which arise naturally in the quantization of the theory. See also the discussions in~\cite{Peldan:1995rv,Matschull:1999he}.
This Paper is organized as follows. In Section 2 we will review the properties of classical Chern-Simons theory. Because MCGs for $\Sigma_1$, $\Sigma_2$ and $\Sigma_g, g\ge3$ have different explicit presentations, in Section 3, 4 and 5 we will quantize the phase space and then the MCG for those three cases respectively.
\section{General Formalism}
Chern-Simons theory with gauge group $\mathrm{U}(1)$ has the action
\se{
I_{\textrm{CS}}[A]=\frac{k}{4\pi}\int_{M}A\wedge dA,
\label{ChernSimonsAction}
}
which has gauge symmetry $A\rightarrow A+g^{-1}dg$, where $g$ is a $\mathrm{U}(1)$-valued function. We shall
consider the Hamiltonian approach on a 3-manifold $M=\mathbb{R}\times\Sigma$ where $\mathbb{R}$ is time and $\Sigma$ is a closed oriented 2-manifold. The 1-form field $A$ defined on $M$ can be decomposed as $A=A_t+A_\Sigma$, with $A_t$ the temporal component and $A_\Sigma$ the components on $\Sigma^2$. The Chern-Simons action is decomposed as
\me{
I_{\textrm{CS}}[A]=\frac{k}{4\pi}\int_{M} A\wedge dA=&\frac{k}{4\pi}\int_{M}(A_t\wedge dA_t+A_\Sigma\wedge dA_t+A_t\wedge dA_\Sigma+A_\Sigma\wedge dA_\Sigma)\\
=&\frac{k}{4\pi}\int_{M}(A_\Sigma\wedge d_t A_\Sigma+2A_t\wedge d_\Sigma A_\Sigma)
\label{ChernSimons}
}
where $d_t$ and $d_\Sigma$ are the exterior differentiation operators on $\mathbb{R}$ and $\Sigma$ respectively. In the above action, $A_t$ acts as a Lagrange multiplier and enforces the constraint that the connection on $\Sigma$ is flat, $F_\Sigma=d_\Sigma A_\Sigma=0$. In the meantime, $A_t$ is integrated over. By an abuse of notation, we use $A$ to denote $A_\Sigma$ from now on.
By Hodge decomposition, any 1-form field $A$ can be written as
\sen{
A=dU+\bar dV+h
}
where $\bar d$ is the adjoint to $d$, $U$ is a 0-form, $V$ is a 2-form, $h$ is a harmonic 1-form, and $U,V,h$ are all $\mathrm{u}(1)$-valued. By the equation of motion $dA=0$, we get $V=0$. Because any $\mathrm{u}(1)$-valued function can be continuously deformed to zero, the term $dU$ can be eliminated by a small gauge transformation. There are still large gauge transformations to consider, which have gauge function with nonzero winding number around some non-trivial loop. To be explicit, for $\Sigma_g$, we can take a set of generators of the fundamental group $\bar\alpha_n,\bar\beta_n,n=1,\ldots,g$, such that $\#(\bar\alpha_n,\bar\alpha_m)=0,\#(\bar\beta_n,\bar\beta_m)=0,\#(\bar\alpha_n,\bar\beta_m)=\delta_{n-m}$, where $\#(,)$ is the algebraic intersection number between two loops. Then there is complete basis of harmonic 1-forms $\omega^{\alpha n}, \omega^{\beta n}$, such that $\oint_{\bar\alpha_n}\omega^{\alpha m}=\oint_{\bar\beta_n}\omega^{\beta m}=\delta_{n-m},\oint_{\bar\alpha_n}\omega^{\beta m}=\oint_{\bar\beta_n}\omega^{\alpha m}=0$, which implies $\int_{\Sigma^2}\omega^{\alpha n}\wedge\omega^{\beta m}=\delta_{n-m}$, and $\int_{\Sigma^2}\omega^{\alpha n}\wedge\omega^{\alpha m}=\int_{\Sigma^2}\omega^{\beta n}\wedge\omega^{\beta m}=0$. Since the field $A$ only has the harmonic part,
\se{
A=\sum_{n=1}^g\left(a_n\omega^{\alpha n}+b_n\omega^{\beta n}\right).
\label{ADecomp}
}
For $N_{\alpha n},N_{\beta n}\in\mathbb{Z}$, gauge transformations of gauge function
\sen{
g(x)=\exp\left[\sum_{n=1}^g\i2\pi\left(N_{\alpha n}\int_{x_0}^x\omega^{\alpha n}+N_{\beta n}\int_{x_0}^x\omega^{\beta n}\right)\right]
}
commute with the small gauge transformations, and are called large gauge transformations. Effectively they translate the variables $a_n,b_n$ by multiples of $\i2\pi$. They form the abelian group $\mathbb{Z}^{2g}$.
Substitute (\ref{ADecomp}) into (\ref{ChernSimons}), the action reduces to
\sen{
I_{\textrm{CS}}=\frac{k}{2\pi}\int dt\sum_{n=1}^ga_n\partial_tb_n.
\label{Defab}
}
According to the canonical quantization recipe, at this point, there is some freedom in choosing the canonical coordinate and momentum. We will use a real polarization here. Namely, $b_i$ and $\frac{k}{2\pi}a_i$ are taken as the canonical variables, with the commutation relation $[a_n,b_n]=\frac{-\i 2\pi}{k}$, and any other pairs commute.
We shall require the quantum states to transform under LGTs covariantly, and try to preserve classical properties of LGT as much as possible. Generators of LGTs can be written as $\rho_n=\exp(ka_n)$ and $\sigma_n=\exp(kb_n)$, and they label different Fourier modes of the wavefunction on the lattice. From the commutator of $a_n$ and $b_n$, we can find they satisfy the relation
\se{
\rho_n\sigma_n=\sigma_n\rho_n\exp(k^2[a_n,b_n])=\sigma_n\rho_n\exp\left(-\i2\pi k\right).
\label{ClockAlgebra}
}
This is called the clock algebra. When $k$ is not integer-valued, this is a deformed version of the classical commutation relation between $\rho_n$ and $\sigma_n$. It seems natural to interpret this deformation of algebra as a quantum effect. If one requires the quantum states to form a representation of the original undeformed classical algebra, then $k$ is quantized to be integer-valued. We will not make such requirement in this paper, and our results apply to the case of integer-valued $k$ straightforwardly.
Aside from carrying a representation of the above clock algebra, a quantum state also stores information that is invariant under LGT. The invariant subspace of the phase space is a $2g$-dimensional torus, parameterized by generators of the holonomy group $\alpha_n=\exp\left(\oint_{\bar\alpha_n}A\right)=\exp(a_n)$ and $\beta_n=\exp\left(\oint_{\bar\beta_n}A\right)=\exp(b_n)$. The non-trivial relation is another clock algebra
\se{
\alpha_n\beta_n=\beta_n\alpha_n\exp\left(-\frac{\i2\pi}{k}\right).
\label{DualClockAlgebra}
}
These two clock algebras (\ref{ClockAlgebra}) and (\ref{DualClockAlgebra}) can be regarded as being dual to each other, with duality transformation $k\leftrightarrow1/k$. Note that these two sets of operators commute. For example, $\alpha_n\sigma_n=\sigma_n\alpha_n\exp(-\i2\pi)$. Thus they realize the statement that on the classical level holonomies are invariant under LGTs.
Since both the LGT group and the holonomy group need to be represented in quantization, and there exists the above interesting duality transformation between them, we like to treat these two groups on an equal footing. This is another reason why $k$ is not restrained to be integer-valued. There is no obvious reason to insist that the holonomy group is undeformed, so by (\ref{DualClockAlgebra}), $1/k$ does not need to integer-valued. To exploit the duality, it is better to allow non-integer-valued $k$ as well.
To quantize the theory, we shall look for representations of the algebras (\ref{ClockAlgebra}) and (\ref{DualClockAlgebra}), as well as of the MCG with appropriate induced action on (\ref{ClockAlgebra}) and (\ref{DualClockAlgebra}), and then the quantum states form the left modules of the representations. Because the algebras (\ref{ClockAlgebra}) and (\ref{DualClockAlgebra}) commute with each other, we can look for representations for them separately, and the complete representation is a tensor product.
\section{Quantization on $\Sigma_1=T^2$}
\label{sec:g1}
Now we proceed to quantize the theory on a torus. We will need the following topological facts. For a torus, the fundamental group $\pi_1(T^2)$ is abelian and generated by two loops $\bar\alpha,\bar\beta$, with $\bar\alpha\bar\beta=\bar\beta\bar\alpha$. The MCG of the torus, $\mathrm{MCG}(T^2)$ is generated by a pair of Dehn twists, $\A,\B$, which can be presented in Fig.\ref{fgg1} as the loops $A, B$. (In the following, we always pick MCG generators in such a way that they can be presented as simple loops, and for a loop $C$, we denote the corresponding MCG generator by $\C$.) To derive how they operate on loops, we take this convention: after performing a Dehn twist, a loop turns left when it hits the representative loop of the Dehn twist, and goes along the Dehn twist loop, until it comes back to the turning point and continues its original path. With this convention, the MCG generators act on the the fundamental group generators as
\begin{subequations}
\begin{align}
\A(\bar\alpha,\bar\beta)=&(\bar\alpha,\bar\beta\bar\alpha),\\
\B(\bar\alpha,\bar\beta)=&(\bar\beta^{-1}\bar\alpha,\bar\beta).
\end{align}
\label{MCGG1Classical}
\end{subequations}
This group is actually $\mathrm{SL}(2,\mathbb{Z})$, so the relations are
\meg{
&\A\B\A=\B\A\B,\\
&(\B\A\B)^4=1.
}
\begin{figure}
\centering
\includegraphics[width=0.35\textwidth]{FigGenus1}
\caption{Fundamental group generators and MCG generators for $\Sigma_1$. The fundamental group generators are denoted by oriented loops with single thin lines, and labeled by Greek letters; the MCG generators are the unoriented loops with double lines, and labeled by Roman letters.}
\label{fgg1}
\end{figure}
Let us consider the representation of holonomy group first. This group is generated by $\alpha$ and $\beta$ with $\alpha\beta=\beta\alpha\exp\left(-\frac{\i2\pi}{k}\right)$. In the representation where $\beta$ is diagonal, $\alpha,\beta$ has the block-diagonal form: (From here on, index of every series starts from 0.)
\se{
\beta=\mathrm{diag}\{\tilde\beta(k,\theta_0^\beta),\ldots,\tilde\beta(k,\theta_{r-1}^\beta)\},\quad\alpha=\mathrm{diag}\{\tilde\alpha(k,\theta_0^\alpha),\ldots,\tilde\alpha(k,\theta_{r-1}^\alpha)\},
}
where each block forms irreducible representation and is determined by two parameters
\se{
\tilde\beta(k,\theta^\beta)=\pmx{1\\&\omega\\&&\ddots\\&&&\omega^{p-1}}e^{\i\theta^\beta},\quad
\tilde\alpha(k,\theta^\alpha)=\pmx{&&&1\\1\\&\ddots\\&&1}e^{\i\theta^\alpha},
\label{alphabeta}
}
where $\omega=\exp\left(\frac{\i2\pi}{k}\right)$, $k=p/q$ with $p,q$ coprime, and each block is $p\times p$ dimensional. See appendix A for a proof that this is the most general irreducible representation of the clock algebra. Note that we have not used up all the freedom of unitary transformation on these matrices. While keeping the above form, we can do a cyclic permutation of basis, which changes $\theta^\beta$ by a multiple of $\frac{2\pi}{p}$, and we can shift the phases of the set of basis, which changes $\theta^\alpha$ by a multiple of $\frac{2\pi}{p}$. Thus the phase parameters can be restricted as $\theta^\beta,\theta^\alpha\in[0,\frac{2\pi}{p})$.
Then we need to find representation for the MCG generators, whose operation on the holonomies is derived from their classical operation (\ref{MCGG1Classical}). The relations that $\A$ should satisfy are
\se{
\A^\dag\alpha\A=\alpha,\quad \A^\dag\beta\A=\exp(b+a)=\beta\alpha\omega^{-1/2}.
\label{AQuantumEqn}
}
To solve these equations, we decompose $\A$ into $r\times r$ blocks of $p\times p$ elements as the holonomies. It turns out each block is given by
\se{
\A_{mn}=u_{mn}^\A\tilde\A(k,\theta_n^\alpha,\theta^\A),
}
where $u_{mn}^\A$ is a complex number, and components of $p\times p$ matrix $\tilde\A$ is given by
\se{
\tilde\A(k,\theta_n^\alpha,\theta^\A)_{ij}=\frac{1}{\sqrt p}\omega^{-(i-j)^2/2}e^{\i(i-j)\theta_n^\alpha}e^{\i\theta^\A}.
\label{ABlock}
}
The parameter $\theta^\A$, which does not depend on the block indices $m,n$, is redundant, because it can be absorbed into $u_{mn}^\A$. The reason for this redundancy will be clear later. In addition, (\ref{ABlock}) solves the equations (\ref{AQuantumEqn}) provided it is periodic with respect to the two indices. Changing $i$ from $0$ to $p$ in $\tilde\A_{ij}$ gives the condition $\omega^{-p^2/2}e^{\i p\theta_n^\alpha}=\exp[\i2\pi(-pq/2+p\theta_n^\alpha/2\pi)]=1$. This means if $p$ or $q$ is even, then $\theta_n^\alpha$ are multiples of $\frac{2\pi}{p}$; if $p,q$ are both odd, then $\theta_n^\alpha$ are multiples of $\frac{2\pi}{p}$ plus $\frac{\pi}{p}$. On the other hand, we knew $0\le\theta_n^\alpha<\frac{2\pi}{p}$, so $\theta_n^\alpha$'s are uniquely determined by $k$,
\sen{
\theta_n^\alpha=\frac{\pi}{p}\cdot\Delta, \quad \Delta=\left\{\begin{array}{ll}0, & \textrm{$p$ or $q$ even,}\\
1, & \textrm{$p$ and $q$ odd.}\end{array}\right.
}
In other words, the representations must have the form $\alpha=I\otimes\tilde\alpha(k,\frac{\pi\Delta}{p}),\beta=I\otimes\tilde\beta(k,\frac{\pi\Delta}{p}),\A=U^\A\otimes\tilde\A(k,\frac{\pi\Delta}{p},\theta^\A)$, where $U^\A$ is any unitary matrix. For $\B$, we need
\se{
\B^\dag\alpha\B=\beta^{-1}\alpha\omega^{1/2},\quad \B^\dag\beta \B=\beta.
}
The solution is
\se{
\B_{mn}=u_{mn}^\B\tilde\B(k,\theta_n^\beta,\theta^\B),
}
where
\sen{
\tilde\B(k,\theta_n^\beta,\theta^\B)_{ij}=\omega^{i^2/2}e^{\i i\theta_n^\beta}\delta_{i-j}e^{\i\theta^\B}.
}
By the same periodicity argument, we have
\se{
\theta_n^\beta=\theta_n^\alpha=\frac{\pi}{p}\cdot\Delta, \quad \Delta=\left\{\begin{array}{ll}0, & \textrm{$p$ or $q$ even,}\\
1, & \textrm{$p$ and $q$ odd.}\end{array}\right.
\label{DefinitionDelta}
}
So $\B$ also have the form $\B=U^\B\otimes\tilde\B(k,\frac{\pi\Delta}{p},\theta^\B)$.
Since MCG of $T^2$ is $\mathrm{SL}(2, \mathbb{Z})$, its generators $\A,\B$ should satisfy the relations $\A\B\A=\B\A\B, (\B\A\B)^4=1$. Because of the redundant parameters $\theta^\A$ and $\theta^\B$, we can require that the two parts of direct production satisfy the relations separately. So $U^\A$ and $U^\B$ are generators of an arbitrary unitary representation of the MCG. $\tilde\A$ and $\tilde\B$ are specified by some parameters, and we need to check how the relations constrain those parameters. For the first relation, the right hand side is
\sen{
(\tilde\B\tilde\A\tilde\B)_{ij}=\frac{1}{\sqrt{p}}\omega^{i^2/2+j^2/2-(i-j)^2/2}e^{\i(i+j)\pi\Delta/p+\i(i-j)\pi\Delta/p}e^{\i(2\theta^\B+\theta^\A)}=\frac{1}{\sqrt{p}}\omega^{ij}e^{\i2i\pi\Delta/p}e^{\i(2\theta^\B+\theta^\A)},
}
and the left hand side is
\men{
(\tilde\A\tilde\B\tilde\A)_{ij}=&\frac{1}{p}\sum_{l=0}^{p-1}\omega^{-(i+j-l)^2/2}\omega^{ij}e^{\i(i-j+l)\pi\Delta/p}e^{\i(2\theta^\A+\theta^\B)}\\
=&\frac{1}{p}\sum_{l=0}^{p-1}\omega^{-l^2/2}\omega^{ij}e^{\i(l+2i)\pi\Delta/p}e^{\i(2\theta^\A+\theta^\B)}\\
=&\frac{1}{p}\sum_{l=0}^{p-1}\omega^{-l^2/2+\Delta l/2q}\omega^{ij}e^{\i2i\pi\Delta/p}e^{\i(2\theta^\A+\theta^\B)}.
}
Comparing the two sides, we find a condition for $\theta^\A, \theta^\B$
\se{
e^{\i(\theta^\B-\theta^\A)}=\frac{1}{\sqrt{p}}\sum_{l=0}^{p-1}\exp\left(\frac{\i\pi(-ql^2+\Delta l)}{p}\right).
\label{RelationOne}
}
The right hand side is a quadratic Gauss sum, which is analytically expressed in term of some number-theoretical function. The exact expression will not be presented here. The important fact is, this equation is satisfiable for all possible $p$ and $q$'s ($p,q$ odd, or $p$ even $q$ odd, or $p$ odd $q$ even). One can check that with the help of the term with $\Delta$, the two sides always have the same abstract values, and the phases can be matched by the yet-unknown $\theta_\A$ and $\theta_\B$. For the second relation $(\B\A\B)^4=1$,
\men{
(\tilde\B\tilde\A\tilde\B)_{ij}=&\frac{1}{\sqrt{p}}\omega^{ij}e^{\i2i\pi\Delta/p}e^{\i(2\theta^\B+\theta^\A)},\\
((\tilde\B\tilde\A\tilde\B)^2)_{ij}=&\frac{1}{p}\sum_{l=0}^{p-1}\omega^{il}\omega^{lj}e^{\i(2i+2l)\pi\Delta/p}e^{\i2(2\theta^\B+\theta^\A)}\\
=&\delta_{(q(i+j)+\Delta)\m p}e^{\i2i\pi\Delta/p}e^{\i2(2\theta^\B+\theta^\A)},\\
((\tilde\B\tilde\A\tilde\B)^4)_{ij}=&\delta_{(i-j)}e^{-\i2\pi\Delta q'/p}e^{\i4(2\theta^\B+\theta^\A)},
}
where $q'$ satisfies $qq'\equiv1\mod{p}$. So we get the another condition on $\theta^\A, \theta^\B$
\se{
e^{-\i2\pi\Delta q'/p}e^{\i4(\theta^\A+2\theta^\B)}=1.
\label{RelationTwo}
}
The two conditions (\ref{RelationOne}) and (\ref{RelationTwo}) will determine $\theta^\A$ and $\theta^\B$ up to 12 choices, but different choices are equivalent. Actually, if $\theta^\A$ and $\theta^\B$ are a set of solution, then $\theta^\A+n\frac{2\pi}{12}$ and $\theta^\B+n\frac{2\pi}{12}$ also solve (\ref{RelationOne}) and (\ref{RelationTwo}). The difference between these two set of solutions is nothing but an abelian representation of the MCG, so it can be absorbed into $U^\A$ and $U^\B$.
We still need to find representation for the dual clock algebra (\ref{ClockAlgebra}) for LGTs and representation for the action of MCG to the LGTs. But these are exactly same algebras as those solved above, except $k$ is replaced by $1/k$, or equivalently, $p,q$ are exchanged. Thus irreducible unitary representation for the dual clock algebra is $q$-dimensional, phase parameters $\theta_n^\rho$ and $\theta_n^\sigma$ are uniquely determined to be $\frac{\pi}{q}\cdot\Delta$, and counterparts of (\ref{RelationOne}) and (\ref{RelationTwo}) give some solution to the phase parameters $\hat\theta^\A$ and $\hat\theta^\B$. The full representation for the algebras is
\men{
&\alpha=I_r\otimes\tilde\alpha\left(k,\pi\Delta/p\right)\otimes I_q,\quad\beta=I_r\otimes\tilde\beta\left(k,\pi\Delta/p\right)\otimes I_q,\\
&\rho=I_r\otimes I_p\otimes\tilde\alpha\left(1/k,\pi\Delta/q\right),\quad\sigma=I_r\otimes I_p\otimes\tilde\beta\left(1/k,\pi\Delta/q\right)\\
&\A=U^\A\otimes\tilde\A\left(k,\pi\Delta/p,\theta^\A\right)\otimes\tilde\A\left(1/k,\pi\Delta/q,\hat\theta^\A\right),\quad\B=U^\B\otimes\tilde\B\left(k,\pi\Delta/p,\theta^\B\right)\otimes\tilde\B\left(1/k,\pi\Delta/q,\hat\theta^\B\right).
}
Unlike Ref.\cite{Bos:1989wa}, value of $k$ is not restricted after imposing MCG. However, representing MCG still affect the representation of the holonomy group and the LGT group, because in this process, the phase parameters $\theta_n^\alpha$ and $\theta_n^\beta$ are determined as in (\ref{DefinitionDelta}), which are otherwise free to change within $[0,\frac{2\pi}{p})$, and $\theta_n^\rho, \theta_n^\sigma$ are restricted similarly.
\section{Quantization on $\Sigma_2$}
The relevant topological properties of $\Sigma_2$ are the following. Fundamental group of $\Sigma_2$ has 4 generators $\bar\alpha_1,\bar\beta_1,\bar
\alpha_2,\bar\beta_2$ with one relation $\bar\alpha_1^{-1}\bar\beta_1\bar\alpha_1\bar\beta_1^{-1}\bar\alpha_2^{-1}\bar\beta_2\bar\alpha_2\bar\beta_2^{-1}=1$. The MCG of $\Sigma_2$, $\mathrm{MCG}(\Sigma_2)$, is generated by five Dehn twists, $\A_1,\B_1,\A_2,\B_2,\S$, as are drawn in Fig.\ref{fgg2}. Their operation on the loops are derived to be
\begin{subequations}
\begin{align}
\A_1(\bar\alpha_1,\bar\beta_1,\bar\alpha_2,\bar\beta_2)&=(\bar\alpha_1,\bar\beta_1\bar\alpha_1,\bar\alpha_2,\bar\beta_2),\label{op1}\\
\B_1(\bar\alpha_1,\bar\beta_1,\bar\alpha_2,\bar\beta_2)&=(\bar\beta_1^{-1}\bar\alpha_1,\bar\beta_1,\bar\alpha_2,\bar\beta_2),\\
\A_2(\bar\alpha_1,\bar\beta_1,\bar\alpha_2,\bar\beta_2)&=(\bar\alpha_1,\bar\beta_1,\bar\alpha_2,\bar\beta_2\bar\alpha_2),\\
\B_2(\bar\alpha_1,\bar\beta_1,\bar\alpha_2,\bar\beta_2)&=(\bar\alpha_1,\bar\beta_1,\bar\beta_2\bar\alpha_2^{-1},\bar\beta_2),\\
\S(\bar\alpha_1,\bar\beta_1,\bar\alpha_2,\bar\beta_2)&=(\bar\alpha_1\bar\beta_1^{-1}\bar\beta_2,\bar\beta_1,\bar\beta_2^{-1}\bar\beta_1\bar\alpha_2,\bar\beta_2),\label{op5}
\end{align}
\end{subequations}
and they satisfy the following relations \cite{Birman:1971}
\begin{subequations}
\begin{align}
&[\A_1,\A_2]=[\A_1,\B_2]=[\B_1,\A_2]=[\B_1,\B_2]=[\B_1,\S]=[\B_2,\S]=1,\label{relation1}\\
&\A_1\B_1\A_1=\B_1\A_1\B_1,\quad \A_2\B_2\A_2=\B_2\A_2\B_2,\quad \A_1\S\A_1=\S\A_1\S,\quad \A_2\S\A_2=\S\A_2\S,\label{relation2}\\
&(\B_1\A_1\S)^4=\B_2^2,\label{relation3}\\
&[\B_2\A_2\S\A_1\B_1\B_1\A_1\S\A_2\B_2,\B_1]=1,\label{relation4}\\
&(\B_2\A_2\S\A_1\B_1\B_1\A_1\S\A_2\B_2)^2=1.\label{relation5}
\end{align}
\end{subequations}
\begin{figure}
\centering
\includegraphics[width=0.65\textwidth]{FigGenus2}
\caption{Fundamental group generators and MCG generators for $\Sigma_2$. The fundamental group generators are denoted by oriented loops with single thin lines, and labeled by Greek letters; the MCG generators are the unoriented loops with double lines, and labeled by Roman letters.}
\label{fgg2}
\end{figure}
By the convention of Fig.\ref{fgg2}, with $a_i=\int_{\alpha_i}A, b_i=\int_{\beta_i}A$, the Chern-Simons action is
\sen{
I_{\textrm{CS}}=\frac{k}{2\pi}\int dt(a_1\partial_tb_1+a_2\partial_tb_2)
}
which results in the commutators
\sen{
[a_1,b_1]=[a_2,b_2]=\frac{-\i2\pi}{k}
}
and any other pair commutes.
Like the previous example, let us only consider representation of the holonomy group for now. From the above commutation relations, the holonomy group generators satisfy the clock algebra,
\se{
\alpha_1\beta_1=\beta_1\alpha_1\omega^{-1}, \quad \alpha_2\beta_2=\beta_2\alpha_2\omega^{-1},
}
and any other pair commutes. Same as in the previous section, we have the following equations for the MCG generators,
\sen{
\B_1^\dag\alpha_1\B_1=\beta_1^{-1}\alpha_1\omega^{1/2},\quad \B_2^\dag\alpha_2\B_2=\beta_2\alpha_2\omega^{1/2},\quad \A_1^\dag\beta_1\A_1=\beta_1\alpha_1\omega^{-1/2},\quad \A_2^\dag\beta_2\A_2=\beta_2\alpha_2^{-1}\omega^{-1/2},
}
and they commute with the rest of the holonomy group generators.
The equations we have listed so far are just two copies of their counterparts in section \ref{sec:g1}, so the solution of the generators is just a direct product of the solutions in section \ref{sec:g1},
\me{
&\beta_1=I_r\otimes\tilde\beta_1,\quad\beta_2=I_r\otimes\tilde\beta_2,\quad \alpha_1=I_r\otimes\tilde\alpha_1,\quad\alpha_2=I_r\otimes\tilde\alpha_2,\\
&\B_1=U_1^\B\otimes\tilde\B_1,\quad\B_2=U_2^\B\otimes \tilde\B_2,\quad \A_1=U_1^\A\otimes\tilde\A_1,\quad\A_2=U_2^\A\otimes \tilde\A_2,
}
where
\men{
&\tilde\beta_1=\tilde\beta\left(k,\pi\Delta/p\right)\otimes I_p,\quad \tilde\beta_2=I_p\otimes\tilde\beta\left(k,\pi\Delta/p\right),\\ &\tilde\alpha_1=\tilde\alpha\left(k,\pi\Delta/p\right)\otimes I_p,\quad \tilde\alpha_2=I_p\otimes\tilde\alpha\left(k,\pi\Delta/p\right),\\
&\tilde\B_1=\tilde\B\left(k,\pi\Delta/p,\theta^\B_1\right)\otimes I_p,\quad \tilde\B_2=I_p\otimes\tilde\B\left(k,\pi\Delta/p,\theta^\B_2\right),\\ &\tilde\A_1=\tilde\A\left(k,\pi\Delta/p,\theta^\A_1\right)\otimes I_p,\quad \tilde\A_2=I_p\otimes\tilde\A\left(k,\pi\Delta/p,\theta^\A_2\right).
}
The new generator $\S$ are determined by the equations
\sen{
\S^\dag\alpha_1\S=\alpha_1\beta_1^{-1}\beta_2\omega^{-1/2},\quad \S^\dag\alpha_2\S=\beta_2^{-1}\beta_1\alpha_2\omega^{1/2},\quad \S^\dag\beta_1\S=\beta_1,\quad \S^\dag\beta_2\S=\beta_2
}
and the solution is
\sen{
\S=U^\S\otimes\tilde\S(k,\theta^\S),
}
where
\sen{
\tilde\S(k,\theta^\S)_{i_1j_1,i_2j_2}=\omega^{(i_2-i_1)^2/2}\delta_{i_1-j_1}\delta_{i_2-j_2}e^{\i\theta^\S}.
}
This matrix is periodic only if one of $p,q$ are even, that is, $\Delta=0$. This is a non-trivial ``quantization condition'' for $k$ on $\Sigma_2$. One can see that if we restrict $k$ to be integer-valued, then this condition agree with that in \cite{Bos:1989wa}, which is $k$ must be an even integer.
Now we check the relations (\ref{relation1})-(\ref{relation5}). (\ref{relation1}) is automatically satisfied. (\ref{relation2}) gives the same equation as (\ref{RelationOne}) except now $\Delta=0$,
\se{
e^{\i(\theta^\B_1-\theta^\A_1)}=e^{\i(\theta^\B_2-\theta^\A_2)}=e^{\i(\theta^\S-\theta^\A_1)}=e^{\i(\theta^\S-\theta^\A_2)}=\frac{1}{\sqrt{p}}\sum_{l=0}^{p-1}\exp\left(\frac{-\i\pi ql^2}{p}\right)\label{result1}
}
which means $\theta^\A_1=\theta^\A_2, \theta^\B_1=\theta^\B_2=\theta^\S$, and these two angles are related by this equation. So there is really only one free angle parameter left. Consider the relation (\ref{relation3}),
\men{
((\tilde\B_1\tilde\A_1\tilde\S)^2)_{i_1j_1,i_2j_2}=&\delta_{i_1+j_1-j_2}\delta_{i_2-j_2}\omega^{-j_1j_2+j_2^2}e^{\i2(\theta^\B_1+\theta^\A_1+\theta^\S)}\\
((\tilde\B_1\tilde\A_1\tilde\S)^4)_{i_1j_1,i_2j_2}=&\delta_{i_1-j_1}\delta_{i_2-j_2}\omega^{j_2^2}e^{\i4(\theta^\B_1+\theta^\A_1+\theta^\S)}
}
so (\ref{relation3}) gives
\se{
e^{\i4(\theta^\B_1+\theta^\A_1+\theta^\S)-\i2\theta^\B_2}=1\label{result2}
}
and this will fix all the angles up to 10 choices. Next consider the relation (\ref{relation4}),
\men{
(\tilde\B_2\tilde\A_2\tilde\S\tilde\A_1\tilde\B_1)_{i_1j_1,i_2j_2}=&\frac{1}{p}\omega^{i_1j_1-i_1j_2+i_2j_2}e^{\i(\theta^\A_1+\theta^\A_2+\theta^\B_1+\theta^\B_2+\theta^\S)}\\
(\tilde\B_2\tilde\A_2\tilde\S\tilde\A_1\tilde\B_1\tilde\B_1\tilde\A_1\tilde\S\tilde\A_2\tilde\B_2)_{i_1j_1,i_2j_2}=&\delta_{i_1+j_1}\delta_{i_2+j_2}e^{\i2(\theta^\A_1+\theta^\A_2+\theta^\B_1+\theta^\B_2+\theta^\S)}
}
It can be checked now that this matrix commutes with $\tilde\B_1$. Multiplying $\tilde\B_1$ from left or from right just adds the factor $\omega^{i_1^2/2}$ or $\omega^{j_1^2/2}$ respectively, and they are the same because of the first delta function. So the relation (\ref{relation4}) poses no condition on the angles. The last relation (\ref{relation5}) is
\sen{
((\tilde\B_2\tilde\A_2\tilde\S\tilde\A_1\tilde\B_1\tilde\B_1\tilde\A_1\tilde\S\tilde\A_2\tilde\B_2)^2)_{i_1j_1,i_2j_2}=\delta_{i_1-j_1}\delta_{i_2-j_2}e^{\i4(\theta^\A_1+\theta^\A_2+\theta^\B_1+\theta^\B_2+\theta^\S)}
}
so
\se{
e^{\i4(\theta^\A_1+\theta^\A_2+\theta^\B_1+\theta^\B_2+\theta^\S)}=1\label{result3}
}
Given (\ref{result1}) and (\ref{result2}), this condition is redundant. The nontrivial conditions are (\ref{result1}) and (\ref{result2}), and they have $10$ distinct solutions, which are equivalent because the difference between the solutions are merely an abelian representation of MCG.
By repeating the above calculation with $k$ replaced by $1/k$, we can get the representation for the dual algebra satisfied by LGT generators. In particular, the quantization condition for $k$, which is one of $p,q$ must be even, is symmetric with respect to $p$ and $q$, so the dual algebra will not give additional restriction on $k$.
\section{Quantization on $\Sigma_g$, $g\ge3$}
Fundamental group and MCG of $\Sigma_g$ with $g\ge3$ have the following presentation. The fundamental group $\pi_1(\Sigma_g)$ is generated by $2g$ loops $\bar\alpha_1,\ldots,\bar\alpha_g,\bar\beta_1,\ldots,\bar\beta_g$, with one relation $\bar\alpha_1^{-1}\bar\beta_1\bar\alpha_1\bar\beta_1^{-1}\cdots\bar\alpha_g^{-1}\bar\beta_g\bar\alpha_g\bar\beta_g^{-1}=1$. For $g\ge3$, MCG have $2g+2$ generators, which are $\A_n,n=1,\ldots,g$, $\S_n,n=1,\ldots,g$, and $\B_n,n=1,2$, as shown in Fig.\ref{fggg}. An explicit presentation of MCG is given in \cite{Wajnryb:1983}. It takes the following form in our convention
\meg{
&[\C,\C']=1,\quad\textrm{when $\#(C,C')=0$},\label{g3r1}\\
&\C\C'\C=\C'\C\C',\quad\textrm{when $\#(C,C')=\pm1$},\quad&\textrm{(braid relation)},\label{g3r2}\\
&(\S_1\A_1\B_1)^4=\E_0\B_2,\quad&\textrm{(3-chain relation)},\label{g3r3}\\
&\E_2\E_1\B_2=\E_3\S_2\S_1\B_1,\quad&\quad\textrm{(lantern relation)},\label{g3r4}\\
&[\A_g\S_{g-1}\A_{g-1}\cdots \S_1\A_1\B_1\B_1\A_1\S_1\cdots \A_{g-1}\S_{g-1}\A_g,\B_g]=1,\quad&\textrm{(hyperelliptic relation)},\label{g3r5}
}
where
\men{
\E_0=&(\A_2\S_1\A_1\B_1\B_1\A_1\S_1\A_2)\B_2(\A_2\S_1\A_1\B_1\B_1\A_1\S_1\A_2)^{-1},\\
\E_1=&(\A_2\S_2\S_1\A_2)^{-1}\B_2(\A_2\S_2\S_1\A_2),\\
\E_2=&(\A_1\S_1\B_1\A_1)^{-1}\E_1(\A_1\S_1\B_1\A_1),\\
\E_3=&(\A_3\S_2\A_2\S_1\A_1\U\B_1^{-1}\A_1^{-1}\S_1^{-1}\A_2^{-1})\B_2(\A_3\S_2\A_2\S_1\A_1\U\B_1^{-1}\A_1^{-1}\S_1^{-1}\A_2^{-1})^{-1},\\
\U=&(\A_3\S_2)^{-1}\E_1^{-1}(\A_3\S_2),
}
and $\B_{n+2}$ is computed from $\B_n, \B_{n+1}$ and the generators by induction
\se{
\B_{n+2}=\W_n\B_n\W_n^{-1},\label{induction}
}
where
\sen{
\W_n=(\A_n\S_n\A_{n+1}\B_{n+1})(\S_{n+1}\A_{n+2}\A_{n+1}\S_{n+1})(\S_n\A_{n+1}\A_n\S_n)(\B_{n+1}\A_{n+1}\S_{n+1}\A_{n+2}).
}
\begin{figure}
\centering
\includegraphics[width=1.0\textwidth]{FigGenusg}
\caption{MCG generators for $\Sigma_g, g\ge3$.}
\label{fggg}
\end{figure}
The holonomies are $g$ copies of quantum torus, and they satisfy clock algebra in pairs. Representation of the $2g+2$ MCG generators can also be derived from the their actions on the holonomies
\men{
&\beta_n=I_r\otimes\tilde\beta_n,\quad \alpha_n=I_r\otimes\tilde\alpha_n,\\
&\A_n=U^\A_n\otimes\tilde\A_n,\quad \S_n=U^\S_n\otimes\tilde\S_n,\quad \B_n=U^\B_n\otimes\tilde\B_n,
}
where
\meg{
&\tilde\beta_n=I_p^{\otimes n-1}\otimes\tilde\beta(k,0)\otimes I_p^{\otimes g-n},\\
&\tilde\alpha_n=I_p^{\otimes n-1}\otimes\tilde\alpha(k,0)\otimes I_p^{\otimes g-n},\\
&\tilde\A_n=I_p^{\otimes n-1}\otimes\tilde\A(k,0,\theta^\A_n)\otimes I_p^{\otimes g-n},\\
&\tilde\S_n=I_p^{\otimes n-1}\otimes\tilde\S(k,0,\theta^\S_n)\otimes I_p^{\otimes g-n-1},\\
&\tilde\B_n=I_p^{\otimes n-1}\otimes\tilde\B(k,0,\theta^\B_n)\otimes I_p^{\otimes g-n}.
\label{Bn}
}
In this process, we get the same quantization condition on $k$, which is one of $p,q$ must be even.
We still need to make sure the relations (\ref{g3r1})-(\ref{g3r5}) are satisfied. (\ref{g3r1}) is trivially satisfied. Braid relations (\ref{g3r2}) give as before
\se{
\theta^\A_n=\theta^\A,\ \theta^\B_n=\theta^\S_n=\theta^\B,\ e^{\i(\theta^\B-\theta^\A)}=\frac{1}{\sqrt{p}}\sum_{l=0}^{p-1}\exp\left(\frac{-\i\pi ql^2}{p}\right)
}
There leaves only one undetermined phase in the MCG generators. The 3-chain relation (\ref{g3r3}) is
\sen{
(\S_1\A_1\B_1)^4=(\A_2\S_1\A_1\B_1\B_1\A_1\S_1\A_2)\B_2(\A_2\S_1\A_1\B_1\B_1\A_1\S_1\A_2)^{-1}\B_2
}
Note that every element in this relation have the same form as in the previous section, so the results can be used. From (\ref{relation4}), we know $[\A_2\S_1\A_1\B_1\B_1\A_1\S_1\A_2,\B_2]=1$, by conjugation and commutation, this relation reduce to the previously proven relation (\ref{relation3}), $(\B_1\A_1\S_1)^4=\B_2^2$. About the lantern relation (\ref{g3r4})
\sen{
\E_2\E_1\B_2=\E_3\S_2\S_1\B_1
}
After some calculation, we find
\men{
(\tilde\E_2\tilde\E_1\tilde\B_2)_{i_1j_1,\ldots,i_gj_g}=&\omega^{j_1^2+j_2^2+j_3^2-j_1j_2-j_2j_3}e^{\i3\theta^\B}\delta_{i_1-j_1}\delta_{i_2-j_2}\delta_{i_3-j_3}\cdots\\
(\tilde\E_3\tilde\S_2\tilde\S_1\tilde\B_1)_{i_1j_1,\ldots,i_gj_g}=&\omega^{j_1^2+j_2^2+j_3^2-j_1j_2-j_2j_3}e^{\i4\theta^\B}\delta_{i_1-j_1}\delta_{i_2-j_2}\delta_{i_3-j_3}\cdots
}
so $\theta^\B=0$. This fix the remaining freedom in the MCG generators. To check the hyperelliptic relation (\ref{g3r5}),
\sen{
[\A_g\S_{g-1}\A_{g-1}\cdots \S_1\A_1\B_1\B_1\A_1\S_1\cdots \A_{g-1}\S_{g-1}\A_g,\B_g]=1
}
by (\ref{induction}), we find $\tilde\B_n$ takes the form of (\ref{Bn}) for all $n$, and
\sen{
(\tilde\A_g\tilde\S_{g-1}\tilde\A_{g-1}\cdots\tilde\S_1\tilde\A_1\tilde\B_1\tilde\B_1\tilde\A_1\tilde\S_1\cdots\tilde\A_{g-1}\tilde\S_{g-1}\tilde\A_g)_{i_1j_1,\ldots,i_gj_g}=\delta_{i_1+j_1}\cdots\delta_{i_g+j_g}e^{\i g(\theta^\A+\theta^\B)}
}
Obviously this commutes with $\tilde\B_g$.
Like in the previous section, representation of the dual clock algebra give no additional quantization condition on $k$.
\section{Discussion}
To summarize the result, we find that the $\mathrm{U}(1)$ Chern-Simons theory defined on $\mathbb{R}\times\Sigma_g$ is quantizable, if we require the quantum states to form representations for the deformed holonomy group, the deformed LGT group, and the MCG. Explicit, finite dimensional representation of these groups are found. The parameter $k$ is quantized in an interesting manner. For $\Sigma_1$, $k$ can take any nonzero rational value, while for $\Sigma_g$ with $g\ge2$, $k$ must be a rational number with either its numerator or denominator being even. The representations are unique in the sense that, apart from a choice of arbitrary unitary representation of MCG, the representations of the discrete groups (holonomy group + LGT group + MCG) are completely fixed.
Uniqueness of the representation of these discrete groups is an interesting result. In general, when one considers gravity with some non-trivial space-time topology, it is expected that some theta angle parameters, or more complicated non-abelian parameters, will arise. But in our toy model, which can be regarded as three dimensional Chern-Simons gravity with gauge group $\mathrm{U}(1)$ instead of some non-abelian gauge group \cite{Witten:1988hc}, although some theta angle parameters appear in representation of the clock algebra, they disappear after representing the MCG.
The representations of the discrete groups on $\Sigma_g$ are found to be $r(pq)^g$ dimensional, where $r$ is the dimensionality of an arbitrary unitary representation of the MCG, and $k=p/q$ with $p,q$ coprime. This can be compared with the result of the path-integral quantization in Ref.\cite{Bos:1989wa}. If we take $r=1$ and $k$ integer-valued, it is easy to check that, the $k^g$ dimensional dimensional representation of the holonomy group from this work is exactly the same representation formed by states in the $k^g$ dimensional Hilbert space found in Ref.\cite{Bos:1989wa}(see Eqn. (23) and (24) therein). As a consequence, the representations of MCG in this work and that in Ref.\cite{Bos:1989wa} are also the same.
The $r(pq)^g$ dimensional Hilbert space that we find can be viewed as a direct product of $g$ $(pq)$-dimensional subspaces, and one $r$-dimensional subspace. Each $(pq)$-dimensional subspace is associated with one specific handle on $\Sigma_g$, and the $r$-dimensional subspace forms an extra representation of the MCG. Due to this decomposed structure, we can consider pinching of the handles, in which some handles are shrank to marked points. Because the quantization condition on $k$ is stronger for $g\ge2$ than for $g=1$, an allowed $k$ value on a higher genus surface will not pose any problem on a lower genus surface. Without loss of generality, assume the first handle is pinched. The holonomy around the remaining marked points is $\alpha_1\beta_1\alpha_1^{-1}\beta_1^{-1}$, which is $\exp\left(-\frac{\i2\pi}{k}\right)$ by Eqn.(\ref{DualClockAlgebra}). This means in a quantum state, after the pinching, all information about holonomies of the first handle is lost. The same is true for information about LGTs, with the same reason. Thus the $(pq)^g$-dimensional subspace becomes $(pq)^{g-1}$ dimensional after pinching. For the $r$-dimensional subspace, the situation is more complicated. In general, representations of MCG of higher genus surface does not reduce to representations of MCG of lower genus surface with marked points, so if pinching is allowed, there may be some extra conditions on the $r$-dimensional representation of MCG.
As shown in previous sections, to impose the large gauge symmetry and large diffeomorphisms, it is impossible to simply reduce the original classical phase space to some invariant subspace of these symmetries. However, it is straightforward, at least classically, to find the invariant phase space under one of the two symmetries. In fact, we chose to represent the large gauge symmetry first, and as a result the $b_1(\Sigma_g)$-dimensional quantum plane, where $b_1$ is the first Betti number, reduces to a $b_1(\Sigma_g)$-dimensional torus times a $\mathbb{Z}^{b_1(\Sigma_g)}$ lattice, both of which are deformed by the canonical commutator to non-commutative spaces. Wave functions take values on these two parts of phase space separately. Then the implementation of MCG gives some non-trivial and rather technical restrictions on the parameters of the theory, as were listed above.
In principle, the other way around, i.e., to impose MCG first should be equally practicable. Indeed, the invariant subspace is the moduli space of Teichm\"uller space of $\Sigma_g$, and wave functions are sections based upon this moduli space. But it is not clear how the wave functions can carry a representation of the large gauge transformations.
|
\section{\bf Introduction}
Let us consider the data $(\ck, \cs, \ct, p)$, where $p$ is an odd prime, $\ck$ is the cyclotomic $\zp$--extension of a number field $K$ (i.e. $\ck$ is a $\zp$--field, in the terminology of \cite{Iwasawa-RH}), and $\cs$ and
$\ct$ are two finite sets of finite primes in $\ck$. We assume that $\ct\cap(\cs\cup\cs_p)=\emptyset$, where $\cs_p$ is the set of $p$--adic primes in $\ck$. Also, let us assume that Iwasawa's $\mu$--invariant
conjecture holds for $K$ and $p$, i.e. we have $\mu_{K,p}=0$, where $\mu_{K,p}$ is the classical Iwasawa $\mu$--invariant associated to $K$ and $p$.
The main goals of this paper are threefold:
{\it Firstly,} for any data $(\ck,\cs, \ct, p)$ as above, we
construct a new class of Iwasawa modules $T_p(\mk)$ which are
$\zp$--free of finite rank (see \S3 below.) We study the $\zp[[\cg]]$--module structure of these modules, under the assumption that
$\ck$ is of CM-type, $\ck$ is a Galois extension
of a totally real number field $k$ of Galois group $\cg:={\rm Gal}(\ck/k)$, the sets $\cs$ and $\ct$ are
$\cg$--invariant, $\ct\ne\emptyset$ and $\cs$ contains the finite primes which ramify in $\ck/k$. In this context, we prove that
\begin{equation}\label{pd-intro}{\rm pd}_{\zp[[\cg]]^-}T_p(\mk)^-=1,\end{equation}
whenever $T_p(\mk)^-\ne 0$. (See Theorem \ref{projective} below.) If $\cg$ is abelian, then the equality above implies that the first Fitting ideal
${\rm Fit}_{\zp[[\cg]]^-} T_p(\mk)^-$ is principal (see Proposition \ref{fitting-calculation}(1) below.)
This brings us naturally to our next goal, namely
the construction of a canonical generator of this ideal.
{\it Secondly,} working under the additional assumption that $\cg$ is abelian, we construct an equivariant $p$--adic $L$--function
$\Theta_{S,T}^{(\infty)}\in\zp[[\cg]]^-$ (see \S\ref{equivariant-L-functions} below) and prove the following Equivariant Main Conjecture - type equality.
\begin{equation}\label{emc-intro} {\rm Fit}_{\zp[[\cg]]^-} T_p(\mk)^-=\Theta_{S,T}^{(\infty)}\cdot\zp[[\cg]]^-.\end{equation}
(See Theorem \ref{emc} below.) This refines the classical Main Conjecture proved by Wiles in \cite{Wiles} in the following precise sense.
For simplicity,
let us assume that $\ck$ contains the group $\bmu_{p^\infty}$ of $p$--power roots of unity. Then, under our working assumption that $\mu_{K,p}=0$, Wiles's
main theorem in \cite{Wiles} can be restated as follows.
\begin{equation}\label{Wiles-intro}{\rm Fit}_{\qp(\zp[[\cg]]^-)}(\qp\otimes_{\zp} \cxs^+(-1)^\ast)= \mathfrak G_S^{(\infty)}\cdot\qp(\zp[[\cg]]^-),\end{equation}
where $\qp(\zp[[\cg]]^-):=\qp\otimes_{\zp}\zp[[\cg]]^-$, $\cxs$ is the classical Iwasawa module given by the Galois group of the maximal pro-$p$ abelian extension of $\ck$, unramified away from $\cs\cup\cs_p$,
and $\mathfrak G_S^{(\infty)}$ is a canonical element\footnote{The reader can easily check that $\mathfrak G_S^{(\infty)}=(\iota\circ t_1)(G_S)$, with notations as in \S\ref{equivariant-L-functions}.} in $\qp(\zp[[\cg]]^-)$ constructed out of the $S$--imprimitive $p$--adic $L$--functions
associated to $\ck/k$.
A na\"ive integral refinement of Wiles's result would aim for a calculation of ${\rm Fit}_{\zp[[\cg]]^-} (\cxs^+(-1)^\ast)$ in terms of $p$--adic $L$--functions.
Unfortunately, the $\zp[[\cg]]^-$--module $\cxs^+(-1)^\ast$ is only very rarely of finite projective dimension. Consequently, its Fitting ideal
is not principal and difficult to compute, in general.
However, we prove that there is an exact sequence of $\zp[[\cg]]^-$--modules
\begin{equation}\label{classical-intro}\xymatrix{0\ar[r] &T_p(\Delta_{\ck, \ct})^-/\zp(1)\ar[r] &T_p(\mk)^-\ar[r] &\cxs^+(-1)^\ast
\ar[r] &0}.\end{equation}
(See \S\ref{classical} below for the proof and notations.) So, the new modules $T_p(\mk)^-$ are certain extensions of the classical modules $\cxs^+(-1)^\ast$ by $T_p(\Delta_{\ck, \ct})^-/\zp(1)$. These modules
happen to be of projective dimension $1$ over $\zp[[\cg]]^-$. Also, a consequence of the definition of the equivariant $p$--adic $L$--function $\Theta_{S,T}^{(\infty)}$ is that
\begin{equation}\label{theta-intro}\Theta_{S,T}^{(\infty)}=\mathfrak G_S^{(\infty)}\cdot\mathfrak D_T^{(\infty)},\end{equation}
where
$\mathfrak D_T^{(\infty)}$ is a non zero--divisor distinguished generator\,\footnote{The reader can easily check that $\mathfrak D_T^{(\infty)}=\delta_T^{(\infty)}/(\iota\circ t_1)(H_S)$, with notations as in \S\ref{equivariant-L-functions}.}
of the Fitting ideal ${\rm Fit}_{\qp(\zp[[\cg]]^-)}(T_p(\Delta_{\ck, \ct})^-/\zp(1)\otimes_{\zp}\qp)$. Since Fitting ideals over $\qp(\zp[[\cg]]^-)$ are multiplicative in short exact sequences, by way of \eqref{classical-intro}--\eqref{theta-intro} our main theorem \eqref{emc-intro} ``tensored with $\qp$'' (i.e. base-changed from the integral group ring $\zp[[\cg]]^-$ to the rational one
$\qp(\zp[[\cg]]^-)$)
is equivalent to Wiles's main theorem \eqref{Wiles-intro}.
{\it Finally}, we use our main theorems (1) and (2) and Iwasawa co-descent to prove refinements of the (imprimitive) Brumer-Stark Conjecture (see Theorem \ref{refined-BS} below)
and the Coates-Sinnott Conjecture (see Theorem \ref{cs-theorem} and Corollary \ref{cs-corollary} below), away from their $2$--primary component, in the most general number field setting.
The main ideas and motivation behind the construction of the new
Iwasawa modules $T_p(\mk)$ are rooted in Deligne's theory of
$1$--motives (see \cite{Deligne-HodgeIII}) and in our previous
work \cite{GP} on the Galois module structure of $p$--adic
realizations of Picard $1$--motives. Also, we were strongly
motivated and inspired by the Deligne-Tate proof of the
Brumer-Stark Conjecture in function fields via $p$--adic
realizations of Picard $1$--motives. (See Chapter V of
\cite{Tate-Stark}.)
In \cite{GP}, we consider the Picard $1$--motive $\cm_{\cs, \ct}^X$ associated to a smooth, projective curve $X$ defined
over an arbitrary algebraically closed field $\kappa$ and two finite, disjoint sets of closed points $\cs$ and $\ct$ on $X$. We study the Galois module structure of its $p$--adic realizations $T_p(\cm_{\cs, \ct}^X)$ in finite $G$--Galois
covers $X\to Y$ of smooth, projective curves over $\kappa$, in the case where $\cs$ contains the ramification locus of the cover and $\cs$ and $\ct$ are $G$--invariant. In particular, if $\kappa=\overline{\Bbb F_q}$, for some finite field $\Bbb F_q$, and the data $(X\to Y, \cs, \ct)$ is defined over $\Bbb F_q$, with $Y=Y_0\times_{\Bbb F_q}\overline{\Bbb F_q}$, for a smooth, projective curve $Y_0$ over $\Bbb F_q$, then we prove that
\begin{equation}\label{ff-intro}{\rm pd}_{\zp[[\cg]]} T_p(\cm_{\cs,\ct}^X)=1, \qquad {\rm Fit}_{\zp[[\cg]]} T_p(\cm_{\cs,\ct}^X)=\zp[[\cg]]\cdot\Theta_{S,T}(\gamma^{-1}),\end{equation}
for all primes $p$, where $\cg:={\rm Gal}(\overline{\Bbb F_q}(X)/\Bbb F_q(Y_0))$, $\gamma$ is the $q$--power arithmetic Frobenius (viewed in $\cg$) and $\Theta_{S,T}(\gamma^{-1})$ is an equivariant $L$--function which is the exact function field analogue of $\Theta_{S,T}^{(\infty)}$. Theorem \ref{projective} (equality \eqref{pd-intro} above) is the number field analogue of the first equality in \eqref{ff-intro}, while Theorem \ref{emc} (equality \eqref{emc-intro} above) is the number field
analogue of the second equality in \eqref{ff-intro}.
Wiles's Theorem (equality \eqref{Wiles-intro} above) is equivalent to the number field analogue of Th\'eor\`eme 2.5 in Ch.V of \cite{Tate-Stark}. The second equality in \eqref{ff-intro} is a natural
integral refinement of Th\'eor\`eme 2.5 in loc.cit. Further, in \cite{GP}, we use the second equality in \eqref{ff-intro} and Galois co-descent with respect to ${\rm Gal}(\overline{\Bbb F_q}/\Bbb F_q)$ to prove refinements
of the Brumer-Stark and Coates-Sinnott Conjectures in function fields. In the number field case, these tasks are achieved with almost identical techniques, in Theorems \ref{refined-BS} and \ref{cs-theorem} below.
In \S\ref{abstract-one-motives} below, we define a category of purely algebraic objects which we call ``abstract $1$--motives''. This category is endowed with covariant functors $T_p$ to the category of free $\zp$--modules
of finite rank, called $p$--adic realizations, for all primes $p$. Two particular examples of abstract $1$--motives are Deligne's Picard $1$--motives $\cm_{\cs, \ct}^X$ and their number field analogues
$\mk$, constructed in \S\ref{the-abstract-one-motive}. The Iwasawa module $T_p(\mk)$ of interest in this paper is the $p$--adic realization of $\mk$, and it is the number field analogue of $T_p(\cm_{\cs, \ct}^X)$.
The paper is organized as follows. In \S\ref{algebra-section}, we define the category of abstract $1$--motives and lay down the number theoretic groundwork necessary for constructing the abstract $1$--motive $\mk$.
In \S\ref{Iwasawa-modules-section}, we construct the abstract $1$--motive $\mk$, show that its $p$--adic realizations $T_p(\mk)$ have a natural Iwasawa module structure and link these to the classical Iwasawa modules $\cxs^+(-1)^\ast$, under the appropriate hypotheses. In \S\ref{ct-section}, we prove the projective dimension result, \eqref{pd-intro} above. In \S\ref{EMC}, we prove the Equivariant Main Conjecture, \eqref{emc-intro} above.
In \S\ref{applications-section}, we use the results obtained in the previous sections to prove refined version of the (imprimitive) Brumer-Stark Conjecture and the Coates-Sinnott Conjecture, away from their $2$--primary components, in the general number field setting. The paper has an Appendix (\S7), where we gather various homological algebra results needed throughout.
In upcoming work, we will give further applications of the results
obtained in this paper and its function field companion \cite{GP}
in the following directions: {\bf I.} An explicit construction of
$\ell$--adic models for Tate sequences in the general case of
global fields; {\bf II.} A proof of the Equivariant Tamagawa
Number Conjecture for Dirichlet motives, in full generality for
function fields, and on the ``minus side'' and away from its
$2$--primary part for CM--extensions of totally real number
fields. A proof of the Gross-Rubin-Stark Conjecture (a vast
refined generalization of the Brumer-Stark Conjecture, see
Conjecture 5.3.4 in \cite{Popescu-PCMI}) will ensue in the cases
listed above; {\bf III.} Finally, we will consider non-abelian
versions of the Equivariant Main Conjectures proved in this paper
and \cite{GP}.
\noindent {\bf Acknowledgement.} The authors would like to thank
their home universities for making mutual visits possible. These visits were funded by the DFG, the NSF, and U. C. San Diego.
\section{\bf Algebraic and number theoretic preliminaries}\label{algebra-section}
\subsection{Abstract $1$--motives}\label{abstract-one-motives}
Assume that $J$ is an arbitrary abelian group, $m$ is a strictly positive integer and $p$ is a prime.
We denote by $J[m]$ the maximal $m$--torsion subgroup of $J$ and let $J[p^\infty]:=\cup_mJ[p^m]$. As usual, $T_p(J)$ will denote the $p$--adic Tate module of $J$. By definition, $T_p(J)$ is the $\Bbb Z_p$--module given by
$$T_p(J):=\underset{n}{\underset{\longleftarrow}\lim}\, J[p^n]\,,$$
where the projective limit is taken with respect to the multiplication--by--$p$ maps. There is an obvious canonical isomorphism
of $\Bbb Z_p$--modules
$$T_p(J)\simeq{\rm Hom}_{\Bbb Z_p}(\Bbb Q_p/\Bbb Z_p, J)={\rm Hom}_{\Bbb Z_p}(\Bbb Q_p/\Bbb Z_p, J[p^\infty])\,.$$
Now, let us assume that $J$ is an abelian, divisible group. $J$ is said to be {\it of finite local corank} if there exists a positive integer $r_p(J)$ and a $\Bbb Z_p$--module isomorphism
$$J[p^\infty]\simeq\left(\Bbb Q_p/\Bbb Z_p\right)^{r_p(J)},$$
for any prime $p$.
The integer $r_p(J)$ is called the $p$--corank of $J$. Obviously, in this case, we have $T_p(J)\simeq\Bbb Z_p^{r_p(J)}$, for all
primes $p$. Further, one can use the ``Hom''--description of $T_p(J)$ given above combined with the injectivity of the $\Bbb Z_p$--module $J[p^\infty]$ to establish
canonical $\Bbb Z_p$--module isomorphisms
$$T_p(J)/p^n T_p(J)\simeq J[p^n],$$
for all primes $p$ and all $n\in\Bbb Z_{\geq 1}$. (See Remark 3.8 in \cite{GP} for these isomorphisms.)
\begin{definition}\label{define-abstract-one-motive} An abstract $1$--motive $\mathcal M:=[L\overset{\delta}\longrightarrow J]$ consists of the following.
\begin{itemize}
\item A free $\Bbb Z$--module $L$ of finite rank (also called a lattice in what follows);
\item An abelian, divisible group $J$ of finite local corank;
\item A group morphism $\delta: L\longrightarrow J$.
\end{itemize}
\end{definition}
In many ways, an abstract $1$--motive $\mathcal M$ enjoys the same properties as an abelian, divisible group of finite local corank.
Namely, for any $n\in\Bbb Z_{\geq 1}$ and any prime $p$, one can construct an $n$--torsion group $\mathcal M[n]$,
a $p$--divisible $p$--torsion group of finite corank $\mathcal M[p^\infty]$, and a $p$--adic Tate module (or $p$--adic realization)
$T_p(\mathcal M)$, which is $\zp$--free of finite rank. This is done as follows.
For every $n\in\mathbb Z_{\geq 1}$, we take the fiber-product of groups
$J\times^n_J L$, with respect to the map $L\overset{\delta}\longrightarrow J$ and the
multiplication by $n$ map $J\overset
n\longrightarrow J.$ An element in this fiber
product consists of a pair $(j, \lambda)\in J\times L,$
such that $nj=\delta(\lambda)$.
Since $J$ is a divisible group, the map $J\overset
n\longrightarrow J$ is surjective. Consequently, we have a
commutative diagram (in the category of abelian groups) whose rows
are exact.
$$\xymatrix {
0\ar[r] & J[n]\ar[r]\ar[d]^{=} &J\times^n_{J}L\ar[r]\ar[d] & L\ar[r]\ar[d]^{\delta} &0\\
0\ar[r] & J[n]\ar[r] & J\ar[r]^n & J\ar[r] &0}
$$
\begin{definition}\label{define-n-torsion} The group $\mathcal M[n]$ of $n$--torsion
points of $\mathcal M$ is defined by
$$\mathcal M[n]:=(J\times^n_{J}L)\otimes\mathbb Z/n\mathbb Z\,.$$
\end{definition}
\noindent Since $L$ is a free $\mathbb
Z$--module, the rows of the above diagram stay exact when tensored with $\mathbb Z/n\mathbb Z$ and $\z/m\z$, respectively. Consequently,
we have commutative diagrams with exact rows
\begin{equation}\label{motive-n-torsion}\nonumber \xymatrix {
0\ar[r] & J[m]\ar[r]\ar@{>>}[d]^{m/n} &\mathcal M[m]\ar[r]\ar@{>>}[d] &L\otimes\mathbb Z/m\mathbb Z\ar[r]\ar@{>>}[d] &0\\
0\ar[r] & J[n]\ar[r] &\mathcal M[n]\ar[r] &L \otimes\mathbb Z/n\mathbb Z\ar[r] &0\,,
}\end{equation}
for all $n, m\in\mathbb N$, with $n\mid m$, where the left vertical map is multiplication by $m/n$, the right vertical map is the canonical surjection and the middle vertical
map is the unique morphism which makes the diagram commute. This middle vertical morphism maps $(j, \lambda)\otimes\widehat 1$ to $(m/n\cdot j, \lambda)\otimes\widehat 1$ and, in what follows, we
refer to it as the multiplication--by--$m/n$ map $\mathcal M[m]\overset{m/n}\longrightarrow\mathcal M[n]$.
\begin{definition} For a prime $p$, the $p$--adic Tate module $T_p(\mathcal M)$ of $\mathcal M$
is given by
$$T_p(\mathcal M)=\underset{\underset n\longleftarrow}\lim\, \mathcal M[p^n]\,,$$
where the projective limit is taken with respect to the surjective multiplication--by--$p$ maps described above.
\end{definition}
\noindent This way, for every prime $p$, we obtain exact sequences of free $\mathbb Z_p$--modules
$$0\longrightarrow T_{p}(J)\longrightarrow T_{p}(\mathcal M)\longrightarrow L\otimes\mathbb Z_p\longrightarrow 0\,.$$
Clearly, we have an isomorphism of $\Bbb Z_p$--modules $T_p(\mathcal M)\simeq \mathbb Z_p^{(r_p(J)+{\rm rk}_{\mathbb Z}L)}$, where
$r_p(J)$ is the $p$--corank of $J$ and ${\rm rk}_{\mathbb Z}L$ is the $\mathbb Z$--rank of $L$.
\medskip
For $n, m\in\Bbb Z_{\geq 1}$ with $n\mid m$, one also has commutative diagrams with exact rows
\begin{equation}\nonumber \xymatrix {
0\ar[r] & J[m]\ar[r] &\mathcal M[m]\ar[r] &L\otimes\mathbb Z/m\mathbb Z\ar[r] &0\\
0\ar[r] & J[n]\ar[r]\ar@{_{(}->}[u] &\mathcal M[n]\ar[r]\ar@{_{(}->}[u] &L \otimes\mathbb Z/n\mathbb Z\ar@{_{(}->}[u]\ar[r] &0\,,
}\end{equation}
with injective vertical morphisms described as follows: the left--most morphism is the usual inclusion; the right-most morphism
sends $\lambda\otimes\widehat x$ to $\lambda\otimes\widehat{m/n\cdot x}$; the middle morphism sends $(j, \lambda)\otimes\widehat 1$
to $(j, m/n\cdot\lambda)\otimes\widehat 1$. Now, if $p$ is a prime number, we define
$$\mathcal M[p^\infty]:=\underset{m}{\underset{\longrightarrow}\lim}\,\mathcal M[p^m]\,,$$
where the injective limit is taken with respect to the injective maps described above. Obviously, $\mathcal M[p^\infty]$ is a
torsion, $p$--divisible group which sits in the middle of an exact sequence
$$0\longrightarrow J[p^\infty]\longrightarrow\mathcal M[p^\infty]\longrightarrow L\otimes\mathbb Q_p/\mathbb Z_p\longrightarrow 0\,,$$
and whose $p$--corank equals $(r_p(J)+{\rm rk}_{\mathbb Z}L)$.
It is not difficult to see that there are canonical $\zp$--module isomorphisms
\begin{equation}\label{Tate-torsion} T_p(\mathcal M)\simeq T_p(\mathcal M[p^\infty]),\qquad T_p(\mathcal M)/p^n T_p(\mathcal M)\simeq \mathcal M[p^n],
\end{equation}
for all $n\in\z_{\geq 1}$.
\begin{remark}\label{abstract-p-adic-1-motives} For a given prime $p$, one can define an {\rm abstract $p$--adic $1$--motive}
$\cm:=[L\overset\delta\longrightarrow J]$ to consist of a $\zp$--module morphism $\delta$ between a free $\zp$--module of finite rank $L$ and a $p$--divisible
$\zp$--module of finite corank $J$. Obviously, the constructions above lead to a $p$--divisible group $\cm[p^\infty]$ and a $p$--adic Tate module
$T_p(\cm)\simeq T_p(\cm[p^\infty])$, for any abstract $p$--adic $1$--motive $\cm$.
Any abstract $1$--motive $\cm:=[L\overset\delta\longrightarrow J]$ gives an abstract $p$--adic $1$--motive
$$\cm_p:=[L\otimes\zp\overset{\delta\otimes\mathbf 1}\longrightarrow J\otimes\zp],$$
for every prime $p$. It is easy to show that there are canonical isomorphisms
$$\cm[p^n]\simeq\cm_p[p^n],\qquad T_p(\cm)\simeq T_p(\cm_p),$$
for all primes $p$ and all $n\in\z_{\geq 1}$.
\end{remark}
\begin{definition}\label{morphism} A morphism $f: \cm'\to\cm$ between two abstract (respectively, abstract $p$--adic) $1$--motives
$\cm':=[L'\overset{\delta'}\longrightarrow J']$ and $\cm:=[L\overset{\delta}\longrightarrow J]$ is a pair
$f:=(\lambda, \iota)$ of group (respectively, $\zp$--module) morphisms
$\lambda: L'\to L$ and $\iota: J'\to J$, such that $\iota\circ\delta'=\delta\circ\lambda$.
\end{definition}
\noindent Now, it is clear what it is meant by composition of morphisms and the category of abstract (respectively, abstract $p$--adic) $1$--motives.
It is easy to see that any morphism $f=(\lambda, \iota)$ as in the definition above induces
canonical $\z/m\z$--module and $\zp$--module morphisms
$$f[m]: \cm'[m]\to \cm[m], \qquad T_p(f): T_p(\cm')\to T_p(\cm)\,,$$
respectively, for all the appropriate $m$ and $p$. Moreover, the associations $f\to f[m]$ and $f\to T_p(f)$
commute with composition of morphisms, so we obtain functors from the category of abstract ($p$--adic) $1$-motives to that of $\z/m\z$--modules
and $\zp$--modules, respectively, for all $m$ and $p$ as above. A consequence is the following.
\begin{remark}\label{equivariance}If $R$ is a ring, $L$ and $J$ are (say, left) $R$--modules and the map $\delta$ is
$R$--linear, then the groups $\cm[m]$, $T_p(\cm)$, $\cm[p^\infty]$ associated to the abstract ($p$--adic) $1$--motive $\mathcal M:=[L\overset{\delta}\longrightarrow J]$ are naturally endowed with $R$--module
structures. Also, all the above exact sequences can be viewed as exact sequences in the category of $R$--modules. This observation will be of particular interest to us
in the case where $R$ is either a group algebra or a profinite group algebra with coefficients in $\z$ or $\zp$.
\end{remark}
\begin{example}\label{Picard} In the case where $J:=\mathcal A(\kappa)$ is the group of $\kappa$--rational points of a semi-abelian variety $\mathcal A$ defined over an algebraically
closed field $\kappa$ and $L$ is an arbitrary lattice then, for any group morphism $\delta: L\to \ca(\kappa)$, the abstract $1$--motive $\mathcal M:=[L\overset{\delta}\longrightarrow J]$
is precisely what Deligne calls a $1$--motive in \cite{Deligne-HodgeIII}.
In particular, let $X$ be a smooth, connected, projective curve defined over an algebraically closed field $\kappa$, and let $\cs$ and $\ct$ be two finite, disjoint
sets of closed points on $X$. Let $J_{\ct}$ be the generalized Jacobian associated to $(X, \ct)$. We remind the reader that $J_{\ct}$ is a semi-abelian variety (an extension of the Jacobian $J_X$ of $X$
by a torus $\tau_\ct$) defined over $\kappa$,
whose group of $\kappa$--rational points is given by
$$J_{\ct}(\kappa)=\frac{{\rm Div}^0(X\setminus\mathcal T)}{\{{\rm div}(f)\mid
f\in\kappa(X)^\times,\quad f(v)=1, \forall\, v\in \ct\}}.$$
Here, ${\rm Div}^0(X\setminus\ct)$ denotes the set of $X$--divisors of degree $0$ supported away from $\ct$ and ${\rm div}(f)$ is the divisor associated to
any non--zero element $f$ in the field $\kappa(X)$ of $\kappa$--rational functions on $X$. The group $J_{\ct}(\kappa)$ is divisible, as an extension of the divisible
groups $J_X(\kappa)$ and $\tau_{\ct}(\kappa)$. Its local coranks satisfy
$$\text{\rm $p$--corank}\,(J_{\ct}(\kappa))=\left\{
\begin{array}{ll}
2g_X+\mid\ct\mid -1, & \hbox{$p\ne{\rm char}(\kappa)$;} \\
\gamma_X, & \hbox{$p={\rm char}(\kappa)$,}
\end{array}
\right.$$
where $g_X$ and $\gamma_X$ are the genus and Hasse--Witt invariant of $X$, respectively. (See \S3 in \cite{GP} for the definitions and the above equality.)
If ${\rm Div}^0(\cs)$ is the lattice of divisors of degree $0$ on the curve $X$
supported on $\cs$ and $\delta: {\rm Div}^0(\cs)\to J_{\ct}(\kappa)$ is the usual divisor--class map, then the abstract $1$--motive
$$\mathcal M^X_{\cs, \ct}:=[{\rm Div}^0(\cs)\overset\delta\longrightarrow J_{\ct}(\kappa)]$$
is called the Picard $1$--motive associated to $(X, \cs, \ct, \kappa)$. Its $\ell$--adic realizations $T_{\ell}(\mathcal M^X_{\cs, \ct})$, for all primes $\ell$,
were extensively studied in \cite{GP}.
\end{example}
In this paper, we will construct a class of abstract $1$--motives which arises naturally in the context of Iwasawa theory of number fields. These should be
viewed as the number field analogues of the Picard $1$--motives $\mathcal M^X_{\cs, \ct}$ described in the example above. As we will see in the next few sections, their $\ell$--adic realizations
satisfy similar properties to those proved to be satisfied by $T_{\ell}(\mathcal M^X_{\cs, \ct})$ in \cite{GP}.
\subsection{Generalized ideal--class groups.}\label{ideal-class-groups} Let $p$ be a prime number. In what follows, we borrow Iwasawa's terminology (see \cite{Iwasawa-RH})
and call a field $\ck$ a {\it $\zp$--field} if it is the cyclotomic $\zp$--extension of a number field. If a $\zp$--field $\ck$ contains the group
${\bmu}_{p^\infty}$ of $p$--power roots of unity, then it is called a {\it cyclotomic $\zp$--field.} A $\zp$--field $\ck$ is said to be of CM--type if it is
the cyclotomic $\zp$--extension of a CM--number field $K$. This is equivalent to the existence of a (necessarily unique) involution $j$ of $\ck$ (the so--called complex conjugation automorphism of $\ck$),
such that its fixed subfield $\ck^+:=\ck^{j=\mathbf 1}$ is the cyclotomic $\zp$--extension of a totally real number field. The uniqueness of $j$ with these properties makes it commute with any field automorphism
of $\ck$.
\smallskip
For the moment, $\mathcal K$ will denote either a number field or a $\zp$--field, for some fixed prime $p$. As usual, a (finite) prime in $\ck$ is an equivalence class of valuations
on $\ck$. If $\ck$ is a number field, all these valuations are discrete of rank one (with value group isomorphic to $(\,\z,\, +)$ with the usual order.) If $\ck$ is a $\zp$--field, then its valuations are
either discrete of rank one, if they extend an $\ell$--adic valuation of $\q$, for some prime $\ell\ne p$, or they have value group isomorphic to $(\,\z[1/p],\, +)$ with the usual order, if they
extend the $p$--adic valuation on $\q$. For every $\ck$ and $v$ as above, we pick a valuation ${\rm ord}_v$ in the class $v$ as follows.
$\bullet$ If $\ck$ is a number field, then ${\rm ord}_v$ is the unique valuation in the class $v$ whose strict value group $\Gamma_v:={\rm ord}_v(\ck^\times)$ satisfies $\Gamma_v=\z$. Note that
if $\ck/\ck'$ is an extension of number fields and $v$ and $v'$ are finite primes in $\ck$ and $\ck'$, with $v$ dividing $v'$, then we have a commutative diagram
\begin{equation}\label{extending-valuations}\xymatrix{\ck^\times\ar[r]^{{\rm ord}_v} &\Gamma_{v}\\
{\ck'}^\times\ar[r]^{\quad{\rm ord}_{v'}}\ar[u]^{\subseteq} &\Gamma_{v'}\ar[u]_{\times\, e(v/v')},}\end{equation}
where $e(v/v'):=[\Gamma_v\,:\,{\rm ord}_v(\ck'^\times)]$ is the usual ramification index.
$\bullet$ If $\ck$ is a $\zp$--field, then we denote by $v_K$ the prime sitting below $v$ in any number field $K$ contained in $\ck$. We let $\Gamma_v:= {\underrightarrow{\lim}}_K \Gamma_{v_K}$,
where the limit is viewed in the category of ordered groups and is taken with respect to all number fields $K$ contained in $\ck$ and the (ordered group morphisms given by) multiplication
by the ramification index maps in the diagram above. Then, since $\ck^\times={\underrightarrow{\lim}}_K\, K^\times$, where the limit is taken with respect to the inclusion morphisms, we can
define
$${\rm ord}_v: \ck^\times \longrightarrow \Gamma_v, \qquad {\rm ord}_v:={\underrightarrow{\lim}}_K {\rm ord}_{{v_K}}\,.$$
It is easy to check that ${\rm ord}_v$ is a valuation on $\ck$ in the class $v$ and that $\Gamma_v={\rm ord}_v(\ck^\times)$. As a consequence of the existence of a number
field $K$, with $K\subseteq\ck$, such that $\ck/K$ is a $\zp$--extension totally ramified at all $p$--adic primes $v$ in $K$ and unramified everywhere else, there are non-canonical ordered group isomorphisms
$$\Gamma_v\simeq\z[1/p],\qquad \Gamma_v\simeq\z,$$
respectively, depending on whether $v$ sits above the $p$--adic prime in $\q$ or not. Also, for any (necessarily finite) extension $\ck/\ck'$ of $\zp$--fields, we have a commutative diagram
identical to (\ref{extending-valuations}) above.
\smallskip
For any field $\ck$ as above, its (additive) divisor group is given by
$${\mathcal Div}_{\ck}:=\bigoplus_v \Gamma_v\cdot v\,,$$
where the direct sum is taken with respect to all the finite primes in $\ck$. Let $\ct$ be a finite (possibly empty) set of finite primes of $\ck$, which is disjoint from the set $\cs_p$ consisting
of all the primes in $\ck$ extending the $p$--adic valuation of $\q$. We let
$${\mathcal Div}_{\ck, \ct}:=\bigoplus_{v\not\in\ct} \Gamma_v\cdot v\,,\qquad \ck_{\ct}^\times:=\{x\in\ck^\times\mid {\rm ord}_v(x-1)>0\,, \forall\, v\in\ct\}\,.$$
The usual divisor map associated to $\ck$ induces a group morphism
$${div}_{\ck}: \ck_{\ct}^\times\longrightarrow {\mathcal Div}_{\ck, \ct}, \quad {div}_{\ck}(x)=\sum_{v}{\rm ord}_v(x)\cdot v=\sum_{v\not\in\ct}{\rm ord}_v(x)\cdot v\,,$$
whose kernel is the following subgroup of the group $U_{\ck}$ of units in $\ck$.
$$U_{\ck, \ct}:=\{u\in U_{\ck}\mid {\rm ord}_v(u-1)>0\,, \text{ for all } v\in\ct \}.$$
To $\ck$ and $\ct$ as above, we associate the following generalized
ideal--class group
$$C_{\ck, \ct}:= \frac{{\mathcal Div}_{\ck, \ct}}{{div}_{\ck}(\ck_{\ct}^\times)}\,.$$
If $\ck/\ck'$ is a finite extension of fields as above and $\ct'$ is the set of primes in $\ck'$ sitting below primes in $\ct$, then we have an injective group morphism
$${\mathcal Div}_{\ck', \ct'}\longrightarrow {\mathcal Div}_{\ck, \ct},\qquad v'\to \sum_{v\mid v'}e(v/v')\cdot v,$$
where the sum is taken over all the primes $v$ of $\ck$ sitting above the prime $v'$ in $\ck'$. This morphism is compatible with the divisor maps (see diagram (\ref{extending-valuations}))
and it induces a (not necessarily injective)
group morphism $C_{\ck', \ct'}\to C_{\ck, \ct}$ at the level of generalized ideal class--groups. It is easy to check that
$$\underset K{\underrightarrow{\lim}}\,{\mathcal Div}_{K, T_K}\simeq {\mathcal Div}_{\ck, \ct},\qquad \underset K{\underrightarrow{\lim}}\,C_{K, T_K}\simeq C_{\ck, T_K}\,,$$
where the limits are taken with respect to all number fields $K\subseteq\ck$ and $T_K$ is the set of primes in $K$ sitting below primes in $\ct$.
In particular, let us assume that $\ck$ is the $\zp$--cyclotomic extension of a number field $K$ and write it as a union of number fields $\ck=\cup_n K_n$, where
$K_n$ is the unique intermediate field $K\subseteq K_n\subseteq \ck$ with $[K_n:K]=p^n$. Then, since the set $\{K_n\mid n\geq 1\}$ is cofinal (with respect to inclusion) in the set of all number fields contained in $\ck$, we have group isomorphisms
\begin{equation}\label{limits}\underset n{\underset \longrightarrow{\lim}}\,{\mathcal Div}_{K_n, T_n}\simeq {\mathcal Div}_{\ck, \ct},\qquad \underset n{\underset \longrightarrow{\lim}}\,C_{K_n, T_n}\simeq C_{\ck, \ct}\,,
\end{equation}
where $T_n$ is the set of primes in $K_n$ sitting below those in $\ct$ and the injective limits are taken with respect to the morphisms described above.
\medskip
If $\ct=\emptyset$, then we drop it from the notation arriving this way at the classical group of units $U_{\ck}$ and ideal--class group $C_{\ck}$ associated to $\ck$. Let
$$\Delta_{\ck, \ct}:=\bigoplus_{v\in\ct}\kappa(v)^\times\,,$$
where $\kappa(v)$ denotes the residue field associated to the prime $v$. It is easy to see that for any $\ct$ as above we have an exact sequence of groups
\begin{equation}\label{t-sequence}\xymatrix{
0\ar[r] &U_{\ck}/U_{\ck, \ct}\ar[r] &\Delta_{\ck, \ct}\ar[r] &C_{\ck, \ct}\ar[r] &C_{\ck}\ar[r] &0.}
\end{equation}
If $\ck$ is a number field, the exact sequence above is described in detail in \cite{Rubin-Stark} or \cite{Popescu-PCMI}, for example. If $\ck$ is the cyclotomic $\zp$--extension
of a number field $K$, then the exact sequence above is obtained by taking the injective limit of the corresponding sequences at the
finite levels $K_n$ with respect to the obvious transition maps. If $\ck$ is a number field, (\ref{t-sequence}) shows that $C_{\ck, \ct}$ is finite.
In that case, the Artin reciprocity map associated to $\ck$ establishes an isomorphism between $C_{\ck, \ct}$ and the Galois group of the maximal abelian extension
of $\ck$ which is unramified away from $\ct$ and at most tamely ramified at primes in $\ct$ (see \cite{Rubin-Stark}.) In the case of a $\zp$--field $\ck$, the second isomorphism
in (\ref{limits}) shows that $C_{\ck, \ct}$ is a torsion group.
\smallskip
Throughout the rest of this section, $\ck$ is a $\zp$--field, for some prime $p$. We let
$$\ca_{\ck}:=C_{\ck}\otimes\zp,\qquad \ca_{\ck, \ct}:=C_{\ck, \ct}\otimes\zp\,,$$
for any $\ct$ as above.
A classical theorem of Iwasawa (see \cite{Iwasawa-RH}, pp. 272-273 and the references therein) shows that there is an isomorphism of
groups
$$\ca_{\ck}\simeq(\qp/\zp)^{\lambda_{\ck}}\oplus\ca',$$
where $\lambda_{\ck}$ is a positive integer which depends only on
$\ck$, called the $\lambda$--invariant of $\ck$ and $\ca'$ is a
torsion $\zp$--module of finite exponent. Iwasawa conjectured that
$\ca'$ is trivial (see loc. cit.). This conjecture is equivalent
to the vanishing of the classical Iwasawa $\mu$--invariants
$\mu_{\ck/K}:=\mu_{K,p}$ associated to $p$ and all number fields
$K$ such that $\ck$ is the cyclotomic $\zp$--extension of $K$.
For a given $\zp$--field $\ck$, although
$\mu_{\ck/K}$ depends on the chosen $K$ with the above
properties, its vanishing is independent of that choice.
The vanishing is known to hold if $\ck$ is the cyclotomic
$\zp$--extension of an abelian number field (see
\cite{Ferrero-Washington}.) In what follows, if $\ck$ is a
$\zp$--field, we write $\mu_{\ck}=0$ to mean that $\mu_{\ck/K}=0$,
for all (one) $K$ as above.
\smallskip
Now, assume that $\ck$ is of CM--type and $j$ is its complex conjugation automorphism. Assume that the prime $p$ is odd. For any $\z$--module $M$ endowed with a $j$--action
(called a $j$--module in what follows) and on which multiplication by $2$ is invertible, we let
$M^\pm:=\frac{1}{2}(1\pm j)\cdot M$ denote the $\pm$--eigenspaces of $j$ on $M$. We have a direct sum decomposition
$M =M^-\oplus M^+\,$ and
the two functors $M\to M^\pm$ are obviously exact.
In particular, if we take $M:=\ca_\ck$, we obtain a direct sum decomposition and $\zp$--module isomorphisms
$$\ca_\ck=\ca_{\ck}^-\oplus\ca_{\ck}^+, \quad \ca_\ck^-\simeq(\qp/\zp)^{\lambda_\ck^-}\oplus\ca^{'-}, \quad \ca_\ck^+\simeq(\qp/\zp)^{\lambda_\ck^+}\oplus\ca^{'+},$$
where $\lambda_\ck^{\pm}$ are positive integers such that $\lambda_\ck=\lambda_\ck^-+\lambda_{\ck}^+$, and $\ca^\pm$ are torsion $\zp$--modules of finite exponent, such
that $\ca'=\ca^{'+}\oplus\ca^{'-}$. Also, if the set of primes $\ct$ above is $j$--invariant, then (\ref{t-sequence}) is as an exact sequence in the category
of $j$--modules. Consequently, we obtain two exact sequences of $\zp$--modules:
\begin{equation}\label{t-sequence-pm}\xymatrix{
0\ar[r] &(U_{\ck}/U_{\ck, \ct}\otimes\zp)^\pm\ar[r] &(\Delta_{\ck, \ct}\otimes\zp)^\pm\ar[r] &\ca_{\ck, \ct}^\pm\ar[r] &\ca^\pm_\ck\ar[r] &0.}
\end{equation}
\begin{lemma}\label{class-group-coranks} Let $\ck$ be a $\zp$--field and $\ct$ a finite, non--empty set of finite primes in $\ck$, disjoint from the set of $p$--adic primes $\cs_p$. Then the following hold.
\begin{enumerate}\item The module $\Delta_{\ck, \ct}\otimes\zp$ is $p$--torsion, divisible, of finite corank denoted $\delta_{\ck, \ct}$.
\item If $\mu_{\ck}=0$, then the module $\ca_{\ck, \ct}$ is $p$--torsion, divisible, of corank at most $\lambda_{\ck}+\delta_{\ck, \ct}$.
\item If $\ck$ is of CM--type, $p$ is odd, $\ct$ is $j$--invariant, and $\mu_{\ck}=0$, then the module $\ca_{\ck, \ct}^{-}$ is $p$-torsion, divisible, of corank
$$\lambda_{\ck}^-+\delta_{\ck, \ct}^--\delta_{\ck},$$
where $\delta_{\ck, \ct}^\pm:={\rm corank}\, (\Delta_{\ck, \ct}\otimes\zp)^\pm$ and $\delta_{\ck}=1$, if $\ck$ is a cyclotomic $\zp$--field and $\delta_\ck=0$, otherwise.
\end{enumerate}
\end{lemma}
\begin{proof} (1) Assume that $\ck$ is the cyclotomic $\zp$--extension of a number field $K$, let $v\in\ct$ and let $v_0$ be the prime in $K$ sitting below $v$.
Since $v\not\in\cs_p$, the prime $v_0$ splits completely up to a finite level $K_n$ in $\ck/K$ and remains inert in $\ck/K_n$. As a consequence, the residue field $\kappa(v)$ associated to $v$ is
a $\zp$--extension of the finite field $\kappa(v_0)$. Consequently, the $p$--primary part $\kappa(v)^\times\otimes\zp$ of the torsion group $\kappa(v)^\times$ either equals the group $\bmu_{p^\infty}$ of all the $p$--power roots of unity or it is trivial, according to whether $\kappa(v_0)^\times$ contains $\bmu_p$ or not. Consequently,
$$\Delta_{\ck, \ct}\otimes\zp\simeq (\qp/\zp)^{\delta_{\ck, \ct}},$$
where $\delta_{\ck, \ct}$ is the number of primes $v\in\ct$, such that $\bmu_p\subseteq\kappa(v)^\times.$ Of course, if $\ck$ is a cyclotomic $\zp$--field, then $\delta_{\ck, \ct}=\mid\ct\mid$. This is because no non--trivial $p$--power root of unity
is congruent to $1$ modulo a non $p$--adic prime $v$, so the reduction mod $v$ maps $\bmu_{p^\infty}\to \kappa(v)$ are injective, for all $v\not\in\cs_p$, in particular for $v\in\ct$.
Part (2) is a consequence of part (1), exact sequence (\ref{t-sequence}), and the fact that the category of $p$--torsion, divisible,
groups of finite corank is closed under quotients and extensions.
Under the hypotheses of part (3), we obviously have
\begin{equation}\label{minus-units}
(U_\ck\otimes\zp)^{-}=\bmu_{p^\infty}\cap \ck^\times\simeq(\qp/\zp)^{\delta_{\ck}}.
\end{equation}
Since $\ct\cap\cs_p=\emptyset$, this implies that $(U_{\ck, \ct}\otimes\zp)^-$ is trivial.
In light of these facts, part (3) is a direct consequence of exact sequence (\ref{t-sequence-pm}).
\end{proof}
Now, assume that $\ck$ is the cyclotomic $\zp$--extension of the number field $K$ and that $\ck$ is of CM--type. Then,
every intermediate field $K_n$ is a CM--number field whose complex conjugation automorphism is the restriction of the complex conjugation $j$
of $\ck$ to $K_n$ and will be denoted by $j$ as well in what follows. For a finite set $\ct$ of finite primes in $\ck$, disjoint from $\cs_p$, we denote by
$T_n$ the set of primes in $K_n$ sitting below primes in $\ct$, for all $n\geq 0$. We let
$$A_{K_n}:=C_{K_n}\otimes \zp, \qquad A_{K_n, T_n}:=C_{K_n, T_n}\otimes\zp\,.$$
If $p$ is odd and the set $\ct$ is $j$--invariant, then we can split these $\zp$--modules
$$A_{K_n}=A_{K_n}^+\oplus A_{K_n}^-, \qquad A_{K_n, T_n}:=A_{K_n, T_n}^+\oplus A_{K_n, T_n}^- ,$$
into their $\pm$--eigenspaces with respect to the $j$--action.
\begin{lemma}\label{no-capitulation} Under the above hypotheses, the natural maps
$$A_{K_n, T_n}^-\longrightarrow A_{K_{n+1}, T_{n+1}}^-$$
are injective, for all $n\geq 0$.
\end{lemma}
\begin{proof} If $\ct=\emptyset$, this is Prop. 13.26 in \cite{Washington}. The general case follows from the case $\ct=\emptyset$ by a snake lemma argument applied to the ``minus''
exact sequence (\ref{t-sequence-pm}).
\end{proof}
\medskip
\section{The relevant Iwasawa modules}\label{Iwasawa-modules-section}
\subsection{An Iwasawa theoretic abstract $1$--motive}\label{the-abstract-one-motive} In this section, we fix a prime $p$, a $\zp$--field $\ck$ and two finite sets
$\cs$ and $\ct$ of finite primes in $\ck$, with the property that $\ct\cap(\cs\cup\cs_p)=\emptyset.$ {\it From this point on, we assume that $\mu_{\ck}=0$.}
To the data $(\ck, \cs, \ct)$ we associate the abstract $1$--motive
$$\cm_{\cs, \ct}^\ck:=[\,{\mathcal Div}_\ck(\cs\setminus\cs_p)\overset{\delta}\longrightarrow\ca_{\ck, \ct}\,]\,,$$
where ${\mathcal Div}_\ck(\cs\setminus \cs_p)$ is the group of divisors of $\ck$ supported on $\cs\setminus\cs_p$ and $\delta$ is the usual divisor--class map
sending the divisor $D$ into $\widehat D\otimes 1$ in $\ca_{\ck, \ct}=C_{\ck, \ct}\otimes\zp$, where $\widehat D$ denotes the class of $D$ in $C_{\ck, \ct}$. Note that, under our current hypotheses, $\rm{Div}_\ck(\cs\setminus\cs_p)$
is a free $\z$--module of rank $d_{\ck, \cs}:=\mid\cs\setminus\cs_p\mid$ and $\ca_{\ck, \ct}$ is torsion, divisible, of finite local corank. This local corank equals $0$ at primes $\ell\ne p$ and
equals at most $(\lambda_{\ck}+\delta_{\ck, \ct})$ at $p$, as Lemma \ref{class-group-coranks}(2) shows. Consequently, all the requirements in Definition \ref{define-abstract-one-motive} are met.
The $p$--adic realization $T_p(\cm_{\cs, \ct}^\ck)$ of $\cm_{\cs, \ct}^\ck$ is a free $\zp$--module of
rank at most $(\lambda_\ck + \delta_{\ck, \ct}+ d_{\ck, \cs})$, sitting in an exact sequence
$$\xymatrix{ 0\ar[r] &T_p(\ca_{\ck, \ct})\ar[r] &T_p(\cm_{\cs, \ct}^\ck)\ar[r] & {\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\zp\ar[r] &0\,. }$$
Next, we give a new interpretation of the $p^n$--torsion groups $\cm_{\cs, \ct}^\ck[p^n]$ of the abstract $1$--motive above, for all $n\geq 1$. For that purpose, we need the following.
\begin{definition} For every $p$, $\ck$, $\cs$, $\ct$ as above and every $m:=p^n$, where $n\in\Bbb Z_{\geq 1}$,
we define the following subgroup of $\ck_{\ct}^\times$.
$$\ck_{\cs, \ct}^{(m)}:=\left\{f\in \ck_{\ct}^\times\mid {div}_{\ck}(f)=mD+y\right\},$$
where $D\in{\mathcal Div}_{\ck, \ct}$ and $y\in{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$. (Note that since the group ${\mathcal Div}_{\ck}(\cs_p)$ is $p$--divisible, we could
write $y\in{\mathcal Div}_{\ck}(\cs)$ or $y\in{\mathcal Div}_{\ck}(\cs\cup\cs_p)$ instead of $y\in{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$ and arrive at an equivalent definition for $\kstm$.)
\end{definition}
\noindent Note that we have an inclusion $\ck_\ct^{\times m}\cdot U_{\ck, \ct}\subseteq \ck_{\cs, \ct}^{(m)}$ of subgroups of $\ck_{\ct}^\times$.
\begin{proposition}\label{reinterpret} For every $p$, $\ck$, $\cs$, $\ct$ and $m$ as in the definition above, we have a canonical group isomorphism
$$\cm_{\cs, \ct}^\ck[m]\simeq \ck_{\cs, \ct}^{(m)}/(\ck_\ct^{\times m}\cdot U_{\ck, \ct})\,.$$
\end{proposition}
\begin{proof} In what follows, in order to simplify notation, we view $\ca_{\ck, \ct}$ as the $p$--Sylow subgroup of the torsion group $C_{\ck, \ct}$. The general element in the fiber--product
$\ca_{\ck, \ct}\times^m_{\ca_{\ck, \ct}}{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$ (see \S\ref{abstract-one-motives} for the definition) consists of a pair
$(\widehat D, x)$, where $D\in{\mathcal Div}_{\ck, \ct}$, such that its class $\widehat D$ in $C_{\ck, \ct}$ lies in $\ca_{\ck, \ct}$, and $x\in{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$
with the property that $mD-x={div}_{\ck}(f)$, for some $f\in\ck_{\ct}^\times$. Note that, by definition, any such $f$ lies in $\ck_{\cs, \ct}^{(m)}$.
We define a map
$$\ca_{\ck, \ct}\times^m_{\ca_{\ck, \ct}}{\mathcal Div}_{\ck}(\cs\setminus\cs_p)\overset{\phi}\longrightarrow \ck_{\cs, \ct}^{(m)}/(\ck_\ct^{\times m}\cdot U_{\ck, \ct}), \qquad \phi(\widehat D, x)=\widehat f, $$
where $D$, $x$ and $f$ are as above, and $\widehat f$ is the class of $f$ in the quotient to the right. We claim that $\phi$ is a well-defined, surjective group morphism. We check this next.
\smallskip
{\bf Step 1. $\phi$ is a well defined group morphism.} Assume that $D, D'\in{\mathcal Div}_{\ck, \ct}$ and $x\in{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$, such that $(\widehat D, x)$ and $(\widehat D', x)$ belong to and are equal in
the fiber product above. This means that
$${div}_{\ck}(f)=mD-x,\qquad {div}_{\ck}(f')=mD'-x, \qquad D-D'={div}_{\ck}(g),$$
for some $f, f', g\in \ck_{\ct}^\times$. Obviously, this implies that ${div}_{\ck, \ct}(f)={div}_{\ck, \ct}(f'g^m)$. Consequently, there exists $u\in U_{\ck, \ct}$, such that
$f=f'\cdot g^mu$. This implies that $\phi(\widehat D, x)=\widehat f=\widehat f'=\phi(\widehat D', x)$, showing that $\phi$ is well-defined as a function.
The fact that $\phi$ is a group morphism is obvious.
\smallskip
{\bf Step 2. $\phi$ is surjective.} Let $f\in\ck_{\cs, \ct}^{(m)}$ and $D\in{\mathcal Div}_{\ck, \ct}$ and $x\in{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$, such that ${div}_{\ck}(f)=mD-x$. Since $C_{\ck, \ct}$ is a torsion group, there exists
a natural number $a$, with $\gcd(a, m)=1$, such that $a\widehat D\in\ca_{\ck, \ct}$. This shows that $(a\widehat D, -ax)\in \ca_{\ck, \ct}\times^m_{\ca_{\ck, \ct}}{\mathcal Div}(\cs\setminus\cs_p)$ and $\phi(a\widehat D, -ax)={\widehat f}^{\,\,a}$. Consequently, ${\widehat f}^{\,\,a}\in{\rm Im}(\phi)$. However, since ${\rm Im}(\phi)$ is a group of exponent dividing $m$ (as a subgroup of $\ck_{\cs, \ct}^{(m)}/\ck_\ct^{\times m}\cdot U_{\ck, \ct}$), it is
uniquely $a$--divisible. This shows that $\widehat f\in {\rm Im}(\phi)$, which concludes the proof of the surjectivity of $\phi$.
\smallskip
{\bf Step 3. The kernel of $\phi$.} Next, we prove that
$$\ker(\phi)=m(\ca_{\ck, \ct}\times^m_{\ca_{\ck, \ct}}{\mathcal Div}_{\ck}(\cs\setminus\cs_p)).$$
Let $(\widehat D, x)\in \ker(\phi)$. This means that there exists $g\in\ck_{\ct}^\times$, such that
$$mD-x={div}_{\ck}(g^m)=m\cdot {div}_{\ck}(g)\,.$$
This implies that $x=mx'$, for some $x'\in{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$. However, since $\ca_{\ck, \ct}$ is $m$--divisible, there exists
$D'\in{\mathcal Div}_{\ck, \ct}$, such that $\widehat D'\in\ca_{\ck, \ct}$ and $\widehat D=m\widehat D'$. This means that
$D-mD'={div}_{\ck}(f')$, for some $f'\in\ck_{\ct}^\times$.
This shows that
$$(\widehat D, x)=m(\widehat D', x'),\qquad mD'-x'={div}_{\ck}(gf')\,.$$
Therefore, $\ker(\phi)\subseteq m(\ca_{\ck, \ct}\times^m_{\ca_{\ck, \ct}}{\mathcal Div}(\cs\setminus\cs_p))$. The opposite inclusion is obvious.
\smallskip
Now, we combine Steps 1--3 above with the definition of $\cm_{\cs, \ct}^\ck[m]$ (see \S\ref{abstract-one-motives}) to conclude that the map $\phi$ factors
through a group isomorphism $\widetilde\phi$
$$\xymatrix{\ca_{\ck, \ct}\times^m_{\ca_{\ck, \ct}}{\mathcal Div}_{\ck}(\cs\setminus\cs_p)\ar@{>>}[r]^{\quad\phi}\ar@{>>}[d] &\ck_{\cs, \ct}^{(m)}/(\ck_\ct^{\times m}\cdot U_{\ck, \ct})\\
\cm_{\cs, \ct}^\ck[m]\ar[ur]^{\widetilde\phi}_{\sim} &
}$$
This concludes the proof of the proposition.
\end{proof}
\begin{remark} If $G$ is a group of field automorphisms of $\ck$ and the sets $\cs$ and $\ct$ are $G$--equivariant, then $T_p(\cm_{\cs, \ct}^\ck)$ and $\cm_{\cs, \ct}^\ck[p^n]$ are endowed with natural
$\zp[G]$--module structures (see Remark \ref{equivariance}.) In this case, it is easily seen that
the canonical isomorphism in Proposition \ref{reinterpret} is $\zp[G]$--linear.
If, in addition, $\ck$ is of CM--type, $p$ is odd, and $\cs$ and $\ct$ are $j$--invariant, where $j$ is the complex conjugation automorphism of $\ck$,
then we can talk about the eigenspaces $T_p(\cm_{\cs, \ct}^\ck)^\pm$ and $\cm_{\cs, \ct}^\ck[p^n]^\pm$, for all $n\geq 1$. Since the actions of $G$ and $j$ commute, these eigenspaces have natural $\zp[G]$--module structures. Moreover, the isomorphism in Proposition \ref{reinterpret} induces two $\zp[G]$--linear isomorphisms
$$\cm_{\cs, \ct}^\ck[m]^\pm\simeq (\ck_{\cs, \ct}^{(m)}/\ck_\ct^{\times m}\cdot U_{\ck, \ct})^\pm\,,$$
for all $m$ which are powers of $p$. This follows immediately from the fact that the action of $j$ and $G$ on $\ck$ commute, for any $G$ as above.
\end{remark}
\begin{corollary}\label{minus-reinterpret} If $\ck$ is of CM--type, $p$ is odd and $\cs$ and $\ct$ are $j$--invariant, then the isomorphism in Proposition \ref{reinterpret} induces isomorphisms
of $\Bbb Z_p$--modules
$$\cm_{\cs, \ct}^\ck[m]^-\simeq (\ck_{\cs, \ct}^{(m)}/\ck_\ct^{\times m}\cdot U_{\ck, \ct})^-\simeq\left(\ck_{\cs, \ct}^{(m)}/\ck_\ct^{\times m}\right)^-\,,$$
for all $m:=p^n$, where $n\geq 1$. If, in addition, $\cs$ and $\ct$ are $G$--invariant, for a group of automorphisms $G$ of $\ck$, then the above
isomorphisms are $\zp[G]$--linear.
\end{corollary}
\begin{proof} Let $m$ be as above. There is an obvious exact sequence of $\zp[\langle j\rangle ]$--modules
$$\xymatrix{
U_{\ck, \ct}/U_{\ck, \ct}^m\ar[r] & \ck_{\cs, \ct}^{(m)}/\ck_\ct^{\times m}\ar[r] & \ck_{\cs, \ct}^{(m)}/(\ck_{\ct}^{\times m}\cdot U_{\ck, \ct})\ar[r] &0.
}$$
Now, (\ref{minus-units}) above shows that $(U_{\ck, \ct}\otimes\zp)^-=\bmu_{p^\infty}\cap U_{\ck, \ct}$. If $\ct\ne\emptyset$, this intersection is trivial, whereas if $\ct=\emptyset$, this intersection
is either equal to $\mu_{p^\infty}$ or trivial, depending on whether $\ck$ is a cyclotomic $\zp$--field or not. In all these cases, $(U_{\ck, \ct}\otimes\zp)^-$ is $p$--divisible.
Consequently, $(U_{\ck, \ct}/U_{\ck, \ct}^m)^-\simeq (U_{\ck, \ct}\otimes\zp)^-\otimes\z/m\z=0$. Consequently, the exact sequence above leads to a group isomorphism
$$(\ck_{\cs, \ct}^{(m)}/\ck_\ct^{\times m})^-\simeq (\ck_{\cs, \ct}^{(m)}/\ck_{\ct}^{\times m}\cdot U_{\ck, \ct})^-.$$
Now, the first part of the corollary is a direct consequence of Proposition \ref{reinterpret}. The $\zp[G]$--linearity is a consequence of the previous Remark combined with
the obvious $\zp[G]$--linearity of the last displayed isomorphism.
\end{proof}
\begin{remark}\label{reinterpret-extensions} Assume that $\ck/\ck'$ is an extension of $\zp$--fields, and $\cs$ and $\ct$ are sets of primes in $\ck$ as above. Let $\cs'$, $\ct'$ and $\cs_p'$ denote
the sets of primes in $\ck'$ sitting below primes in $\cs$, $\ct$ and $\cs_p$, respectively. Then, the inclusion $\ck'\to\ck$ induces natural group morphisms ${\mathcal Div}_{\ck'}(\cs'\setminus\cs_p')\to
{\mathcal Div}_{\ck}(\cs\setminus\cs_p)$ and $\ca_{\ck', \ct'}\to\ca_{\ck, \ct}$. These lead to a morphism of abstract $1$--motives $\mkp\to\mk$.
Since the isomorphisms constructed in Proposition \ref{reinterpret} are functorial, we get commutative diagrams
$$\xymatrix{\mk[m]\ar[r]^{\sim\qquad } &\kstm/(\ktm\cdot U_{\ck, \ct})\\
\mkp[m]\ar[u]\ar[r]^{\sim\qquad } &\kpstm/(\kptm\cdot U_{\ck', \ct'}).\ar[u]
}
$$
\end{remark}
\medskip
\subsection{The ensuing Iwasawa modules} Let us assume that $\ck$ is the cyclotomic $\zp$--extension of a number field $K$. As usual,
we let $\Gamma:={\rm Gal}(\ck/K)$. We pick two finite sets $\cs$ and $\ct$ of finite primes in $\ck$, such that $\ct\cap(\cs\cup\cs_p)=\emptyset$. We assume that
$\cs$ and $\ct$ are $\Gamma$--invariant.
For all $n\in\z_{\geq 0}$, we let $\Gamma_n:={\rm Gal}(\ck/K_n)$, where $K_n$ is the unique intermediate field $K\subseteq K_n\subseteq\ck$,
such that $[K_n:K]=p^n$. Also, we let
$$\Lambda=\zp[[\Gamma]]:=\underset{n}{\underleftarrow{\lim}}\,\zp[\Gamma/\Gamma_n]$$
denote the $p$--adic profinite group ring associated to $\Gamma$, where the projective limit is taken with
respect to the surjections $\Gamma/\Gamma_{n+1}\twoheadrightarrow\Gamma/\Gamma_n$ induced by Galois restriction.
As in \S\ref{ideal-class-groups}, we denote by $T_n$, $S_n$ and $S_{p,n}$ the sets of primes in $K_n$ which sit below primes in $\cs$, $\ct$, and $\cs_p$,
respectively, for all $n\geq 0$. Obviously, $S_n$, $T_n$ and $S_{p, n}$ are $\Gamma/\Gamma_n$--invariant. Consequently, $A_{K_n, T_n}$ and ${\mathcal Div}_{K_n}(S_n\setminus S_{p, n})\otimes\zp$ have canonical
$\zp[\Gamma/\Gamma_n]$--module structures. Via the ring morphisms $\Lambda\twoheadrightarrow\zp[\Gamma/\Gamma_n]$, they are endowed with canonical $\Lambda$--module structures.
The natural morphisms
$$A_{K_n, T_n}\longrightarrow A_{K_{n+1}, T_{n+1}}, \qquad {\mathcal Div}_{K_n}(S_n\setminus S_{p,n})\otimes\zp\longrightarrow {\mathcal Div}_{K_{n+1}}(S_{n+1}\setminus S_{p,n+1})\otimes\zp$$
are $\Lambda$--linear. This leads to canonical $\Lambda$--module structures for
$$\ca_{\ck, \ct}\simeq \underset{n}{\underrightarrow\lim}\, A_{K_n, T_n}\quad \text{ and }\quad {\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\zp\simeq
\underset{n}{\underrightarrow\lim}\,({\mathcal Div}_{K_n}(S_n\setminus S_{p,n})\otimes\zp),$$
as direct limits of $\Lambda$--modules with respect to $\Lambda$--linear transition maps. This way, $T_p(\ca_{\ck, \ct})$ inherits
a canonical $\Lambda$--module structure as well.
\begin{lemma}\label{lambda-module-lemma} With notations as above, the $\zp$--module $T_p(\mk)$ can be endowed with a natural $\Lambda$--module structure which makes the sequence
$$\xymatrix{ 0\ar[r] &T_p(\ca_{\ck, \ct})\ar[r] &T_p(\cm_{\cs, \ct}^\ck)\ar[r] & {\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\zp\ar[r] &0}$$
exact in the category of $\Lambda$--modules.
\end{lemma}
\begin{proof}
Remark \ref{abstract-p-adic-1-motives} shows that in defining the torsion groups $\mk[m]$, with $m$ a power of $p$, and the $p$--adic Tate module $T_p(\mk)$, we may replace the
abstract $1$--motive $\mk$ with the associated abstract $p$--adic $1$--motive
$$(\mk)_p:=[{\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\zp\overset{\delta\otimes 1}\longrightarrow\ca_{\ck, \ct}],$$
where $\delta\otimes 1$ is the extension of $\delta$ by
$\zp$--linearity to ${\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\zp$. (Note that $\ca_{\ck, \ct}\simeq \ca_{\ck, \ct}\otimes\zp$.) Now, we obviously have
$$[{\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\zp\overset{\delta\otimes 1}\longrightarrow\ca_{\ck, \ct}]=\underset{n}{\underrightarrow\lim}\,[{\mathcal Div}_{K_n}(S_n\setminus S_{p,n})\otimes\zp\overset{\delta_n\otimes 1}\longrightarrow A_{K_n, T_n}]\,$$
where $\delta_n$ is the usual divisor--class map
sending the divisor $D\in {\mathcal Div}_{K_n}(S_n\setminus S_{p,n})$ to $\widehat D\otimes 1$ in $C_{K_n, T_n}\otimes\zp=A_{K_n, T_n}$, with $\widehat D$ denoting the class of $D$ in $C_{K_n, T_n}$. Since the maps
$\delta_n\otimes 1$ are $\zp[\Gamma/\Gamma_n]$--linear, they are $\Lambda$--linear and therefore $\delta\otimes 1$ is $\Lambda$--linear. Remark \ref{equivariance} applied to ${(\mk)_p}$ and $R:=\Lambda$,
implies that $\mk[m]\simeq (\mk)_p[m]$ can be endowed with a natural
$\Lambda$--module structure which makes the sequence
$$\xymatrix{0\ar[r] &\ca_{\ck, \ct}[m]\ar[r] &\mk[m]\ar[r] &{\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\z/m\z\ar[r] & 0
}$$
exact in the category of $\Lambda$--modules, for all $m$ which are powers of $p$. The statement in the Lemma is now obtained by taking the projective limit
of the exact sequences above for $m=p^a$, with $a\in\z_{\geq 1}$, with respect to the multiplication-by-$p$ maps, which are clearly $\Lambda$--linear.
\end{proof}
\begin{remark}\label{lambdag} Assume that $K/k$ is a Galois extension of number fields, of Galois group $G$. Let $\ck$ and be the cyclotomic $\zp$--extensions of $K$,
for some prime $p$. Then, the extension $\ck/k$ is Galois. Let $\cg:={\rm Gal}(\ck/k)$.
Now, let us assume that the sets $\cs$ and $\ct$ in $\ck$ considered above are $\cg$--invariant. Then, just as above, ${\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\zp$ and $\ca_{\ck, \ct}$ are endowed
with obvious canonical $\zp[[\cg]]$--module structures and the map $\delta\otimes 1$ above is $\zp[[\cg]]$--linear. Consequently, $\mk[p^a]$, for $a\geq 1$, and $T_p(\mk)$ are endowed with a canonical $\zp[[\cg]]$--module structure
and the sequence in Lemma \ref{lambda-module-lemma} is exact in the category of $\zp[[\cg]]$--modules.
If, in addition, $K$ is a CM--number field (which makes $\ck$ of CM--type), $\cs$ and $\ct$ are $j$--invariant
and $p$ is odd, then $\mk[p^a]^\pm$, for $a\geq 1$, and $T_p(\mk)^\pm$ have natural $\zp[[\cg]]$--module structures.
The exact sequence of $\zp[[\cg]]$--modules in Lemma \ref{lambda-module-lemma} can be split into a direct sum of two exact sequences of $\zp[[\cg]]$--modules
\begin{equation}\label{motive-exact-sequence}
\xymatrix{ 0\ar[r] &T_p(\ca_{\ck, \ct})^\pm\ar[r] &T_p(\cm_{\cs, \ct}^\ck)^\pm\ar[r] & {\mathcal Div}_{\ck}(\cs\setminus\cs_p)^\pm\otimes\zp\ar[r] &0}.\end{equation}
\end{remark}
\begin{remark} For every $n\geq 0$, let $M_{S_n, T_n}^{K_n}:=[{\mathcal Div}_{K_n}(S_n\setminus S_{p,n})\overset{\delta_n}\longrightarrow A_{K_n, T_n}]$, where $\delta_n$ is the divisor--class map defined in the proof of the above Lemma. Of course, $M_{S_n, T_n}^{K_n}$ is not an abstract $1$--motive,
unless the finite group $A_{K_n, T_n}$ happens to be trivial. Nevertheless, one can define the $\zp[\Gamma/\Gamma_n]$--modules
$$M_{S_n, T_n}^{K_n}[m]:=(A_{K_n, T_n}\times^m_{A_{K_n, T_n}}{\mathcal Div}_{K_n}(S_n\setminus S_{p,n}))\otimes\z/m\z,$$
for every $m$ which is a power of $p$. For all $m$ and $n$ as above, we have obvious and canonical $\Lambda$--module morphisms
$$M_{S_n, T_n}^{K_n}[m]\to M_{S_{n+1}, T_{n+1}}^{K_{n+1}}[m]\to \mk[m].$$
It is easily proved that these lead to a $\Lambda$--module isomorphism
$$\underset{n}{\underrightarrow\lim}\, M^{K_n}_{S_n, T_n}[m] \simeq \mk[m].$$
\end{remark}
\subsection{Linking $T_p(\cm_{\cs, \ct})^-$ to the classical Iwasawa modules}\label{classical} In this section, we assume that $K/k$ is a Galois extension of number fields,
where $K$ is CM and $k$ is totally real. We fix a prime $p>2$ and
let $\ck$ denote the cyclotomic $\zp$--extension of $K$. We let $G:={\rm Gal}(K/k)$, $\cg:={\rm Gal}(\ck/k)$, $\Gamma:={\rm Gal}(\ck/K)$ and $\Lambda:=\zp[[\Gamma]]$.
Let $\cs$ be a $\cg$--equivariant finite set of finite primes in $\ck$. As in classical Iwasawa theory, we denote by $\cxs$ the Galois group of the maximal abelian pro--$p$
extension of $\ck$ which is unramified away from $\cs\cup\cs_p$. Then, $\cxs$ is endowed with the usual canonical $\zp[[\cg]]$--module structure (with the $\cg$--action on $\cxs$ given
by lift--and--conjugation.) In particular, $\cxs$ has a canonical $\Lambda$--module structure. Recall Iwasawa's classical theorems stating that
$\cxs$ is a finitely generated $\Lambda$--module of rank $r_2(K)$ (the number of complex infinite primes in $K$) and that $\cxs^+$ is $\Lambda$--torsion and contains no non--trivial
finite $\Lambda$--submodules (see \cite{Iwasawa-zl}.)
Further, let us assume that $\bmu_p\subseteq K$ (i.e. $\bmu_{p^\infty}\subseteq\ck$.) In what follows, we adopt the notations and definitions of \S\ref{appendix-twisting} in the Appendix. In particular, note
that under our current assumptions the Teichm\"uller component $\omega_p$ of the $p$--adic cyclotomic character $c_p:\cg\to\zp^\times$ factors through $G$.
By the definitions of $\cxs$ and $\ck_{\cs, \emptyset}$, Kummer theory leads to perfect $\zp$--bilinear pairings of $\zp[[\cg]]$--modules
$$\langle \cdot\,,\,\cdot\rangle _m\,:\, \cxs/p^m\cxs \, \times\, \ck_{\cs, \emptyset}^{(p^m)}/\ck^{\times p^m}\longrightarrow\, \bmu_{p^m}\,,$$
for all $m\in\z_{\geq 1}$, with the $\cg$--equivariance property
\begin{equation}\label{pairing-equivariance}
\langle {}^gx,\, {}^gy\rangle _m=g(\langle x,\, y\rangle_m),\end{equation}
for all $x\in \cxs/p^m\cxs$,
$y\in \ck_{\cs, \emptyset}^{(p^m)}/\ck^{\times p^m}$ and $g\in \cg$. Consequently, for all $m$ as above,
we obtain perfect $\zp$--bilinear pairings
$$\cxs^+/p^m\cxs^+ \, \times\, (\ck_{\cs, \emptyset}^{(p^m)}/\ck^{\times p^m})^-\longrightarrow\, \bmu_{p^m}\,,$$
with property (\ref{pairing-equivariance}). Now, we use Corollary \ref{minus-reinterpret} and pass to a projective limit with respect to $m$ and the obvious transition maps in the pairings above to obtain a perfect $\zp$--bilinear, continuous pairing of $\zp[[\cg]]$--modules
$$\langle \cdot\,,\,\cdot\rangle \,:\, \cxs^+\, \times\, T_p(\cm^\ck_{\cs, \emptyset})^-\longrightarrow \zp(1),$$
with property (\ref{pairing-equivariance}). Consequently, the following holds.
\begin{lemma}\label{link-classical} The last pairing induces an isomorphism of
$\zp[[\cg]]$--modules
$$T_p(\cm^\ck_{\cs, \emptyset})^-\simeq\, {\rm Hom}_{\zp}(\mathfrak X_S^+,\, \zp(1)), $$
where the right-hand side has the $\zp[[\cg]]$--module structure given by
$$(\lambda\ast f)(x):=f((\iota\circ t_1)(\lambda)\cdot x),$$
for all $f\in{\rm Hom}_{\zp}(\mathfrak X_S^+,\, \zp(1))$,
$x\in \mathfrak X_S^+$ and $\lambda\in \zp[[\cg]]$.
\end{lemma}
\begin{proof}
For all $x\in \mathfrak X_S^+$, $y\in T_p(\cm^\ck_{\cs, \emptyset})^-$ and $g\in \cg$, we have
$$\langle {}^gx,\, y\rangle =g(\langle x,\, {}^{g^{-1}}y\rangle )=c_p(g)\langle x,\, {}^{g^{-1}}y\rangle =\langle x,\,(\iota\circ t_1)(g)\cdot y\rangle .$$
This shows that the isomorphism above is at least $\zp[\cg]$--linear. However, since it is
continuous, it has to be $\zp[[\cg]]$--linear, as desired. \end{proof}
\begin{remark}\label{remark-link-classical} With notations as above, let $\ct$ be a finite, nonempty, $\cg$--invariant set of finite primes
in $\ck$, which is disjoint from $\cs\cup\cs_p$. Then, the group morphism $\ca_{\ck, \ct}\twoheadrightarrow\ca_\ck$ induces
an obvious morphism of abstract $1$--motives $\mk\longrightarrow \cm^\ck_{\cs, \emptyset}$. This leads to a morphism $T_p(\mk)\longrightarrow T_p(\cm^\ck_{\cs, \emptyset})$ of
$\zp[[\cg]]$--modules. As a consequence of (\ref{t-sequence-pm})
this morphism leads to an exact sequence of $\zp[[\cg]]$--modules
\begin{equation}\label{sequence-empty-T}\xymatrix{0\ar[r] &T_p(\Delta_{\ck, \ct})^-/\zp(1)\ar[r] &T_p(\mk)^-\ar[r] &T_p(\cm^\ck_{\cs, \emptyset})^-
\ar[r] &0}\end{equation}
Therefore, Lemma \ref{link-classical} above gives surjective morphisms of $\zp[[\cg]]$--modules
$$T_p(\mk)^-(n-1)\twoheadrightarrow T_p(\cm^\ck_{\cs, \emptyset})^-(n-1))\simeq (\cxs^+)^\ast(n),\qquad \forall\, n\in\z.$$
Above, $(\cxs^+)^\ast:={\rm Hom}_{\zp}(\cxs^+, \zp)$, endowed with the contravariant $\cg$--action given by ${}^gf(x):=f(g^{-1}\cdot x)$, for
all $f\in(\cxs^+)^\ast$, $g\in \cg$ and $x\in \cxs^+$.
\end{remark}
\section{Cohomological triviality}\label{ct-section}
Throughout this section, $\ck/\ck'$ will denote a Galois extension of $\zp$--fields, of Galois group $G$. We fix two finite sets $\cs$ and $\ct$
of finite primes in $\ck$, such that
$\ct\cap(\cs_p\cup\cs)=\emptyset$. From now on, we assume that $\ct\ne\emptyset$ and $\cs$
contains the finite ramification locus $\cs_{\rm ram}^{\rm fin}(\ck/\ck')$ of
$\ck/\ck'$. Also, we assume that $\cs$ and $\ct$ are $G$--invariant and
let $\cs'$ and $\ct'$ denote the sets consisting of all primes on
$\ck'$ sitting below primes in $\cs$ and $\ct$, respectively.
If $\ck$ is of CM--type, $j$ will denote, as usual, the complex conjugation automorphism of $\ck$. Since $j$ commutes with any element in $G$ (as automorphisms of $\ck$), it is easy to check that, in this case,
$\ck'$ is either totally real
or of CM--type, depending on whether $\ck'\subseteq\ck^+:=\ck^{j=1}$ or not. In the case where both $\ck$ and $\ck'$ are of CM--type, then
the complex conjugation automorphism of $\ck'$ is the restriction of $j$ to $\ck'$ and will be denoted by $j$ as well. If $\ck$ is of CM--type, we will assume further that $\cs$ and $\ct$ are $j$--invariant as well.
{\it Throughout this section, we assume the vanishing of the Iwasawa $\mu$--invariant of all $\zp$--fields involved.}
For simplicity, we let $\cm:=\mk$ and $\cm':=\mkp$. The natural abstract $1$--motive morphism $\cm'\to\cm$ (see Remark \ref{reinterpret-extensions}) induces $\zp$--module morphisms $\cm'[p^n]\to\cm[p^n]^G$ and
$T_p(\cm')\to T_p(\cm)^G$. The main goal of this section is to use these morphisms in order to study the $\zp[G]$--module structure of
$T_p(\cm)^-$, in the case where $\ck$ is of CM--type and $p$ is odd.
\medskip
\begin{proposition}\label{invariants} Assume that $m$ is a non--trivial power of
$p$. Then,
\begin{enumerate}
\item The inclusion $\ck'^\times\subseteq\ck^\times$ induces a group isomorphism
$$\kpstm/\kptm \simeq (\kstm/\ktm)^G\,.$$
\item If $\ck$ and $\ck'$ are of CM--type and $p$ is odd, then the morphism $\cm'\to\cm$ of abstract $1$--motives
induces canonical $\zp$--module isomorphisms
$$\cm'[m]^-\simeq(\cm[m]^{-})^G, \qquad T_p(\cm')^- \simeq (T_p(\cm)^-)^G\,.$$
\end{enumerate}
\end{proposition}
\begin{proof} (1) Let us fix an $m$ as above. Since $\ck/\ck'$ is unramified at finite primes outside of $\cs'$, we have inclusions
$$\kpt=({\kt})^G\subseteq\kt, \qquad \kpstm\subseteq({\kstm})^G\subseteq\kstm.$$
Since $\ct$ is non-empty and disjoint from $\cs_p$, the group $\kt$ has no $m$--torsion. Therefore, the $m$--power--map induces
a $G$--invariant group isomorphism $\kt\simeq\ktm$. If one takes $G$--invariants in this isomorphism, one obtains
$$\kptm=(\ktm)^G=\ktm\cap\ck'^\times.$$
When combined with the displayed inclusions above, this leads to an injection
$$\kpstm/\kptm \hookrightarrow (\kstm/\ktm)^G\,.$$
In order to complete the proof of (1), we need to show that this injection is an isomorphism. For this purpose, we write the first four terms in the long $G$--cohomology sequence
associated to the short exact sequence of $\z[G]$--modules
$$\xymatrix{1\ar[r] &\ktm\ar[r]\ar[r] &\kstm\ar[r] &\kstm/\ktm\ar[r] &1}.$$
We obtain an exact sequence of multiplicative groups
$$\xymatrix{1\ar[r] &\kptm\ar[r] &(\kstm)^G\ar[r] &(\kstm/\ktm)^G\ar[r] & H^1(G,\, \ktm)\ar[r]&\cdots}. $$
Therefore, (1) would be a consequence of the following equalities.
\begin{equation}\label{reduction}
H^1(G,\, \ktm)=0,\qquad (\kstm)^G=\kpstm.
\end{equation}
Since $\kt\simeq\ktm$, the first equality above is equivalent to
$H^1(G,\, \kt)=0$. This is proved as follows. We let
$\ck_{(\ct)}^\times:=\{x\in\ck^\times\mid {\rm ord}_w(x)=0,
\forall w\in\ct\}$. We have short exact sequences of
$\z[G]$--modules
$$\xymatrix{1\ar[r]&\kt\ar[r] &\ck_{(\ct)}^\times\ar[r]^{{\rm res}_{\ct}} &\Delta_{\ck, \ct}\ar[r] & 1},$$
$$\xymatrix{1\ar[r]&\ck_{(\ct)}^\times\ar[r] &\ck^\times\ar[r]^{\quad{div}_{\ck, \ct}\qquad} &{\mathcal Div}_{\ck}(\ct)\ar[r] & 0.}$$
Here, we have ${\rm res}_{\ct}(x):=(x\mod w)_{w\in\ct}$, for all
$x\in\ck_{(\ct)}^\times$, and ${div}_{\ck,
\ct}(x):=\sum_{w\in\ct}{\rm ord}_w(x)\cdot w$, for all
$x\in\ck^\times$. The maps ${\rm res}_{\ct}$ and ${div}_{\ck,
\ct}$ are surjective as a consequence of the weak approximation
theorem applied to the independent valuations of $\ck$
corresponding to the primes in $\ct$. Since the extension
$\ck/\ck'$ is unramified at primes in $\ct'$ and
$\ct\cap\cs_p=\emptyset$, we have $\z[G]$--module isomorphisms
\begin{equation}\label{Delta-T}
\Delta_{\ck, \ct}\simeq\bigoplus_{w'\in\ct'}(\kappa(w)^\times\otimes_{\z[G_w]}\z[G]), \qquad {\mathcal Div}_{\ck}(\ct)\simeq \bigoplus_{w'\in\ct'}(\z w\otimes_{\z[G_w]}\z[G])\,,
\end{equation}
where $w$ is a prime in $\ct$ sitting above $w'$ and $G_w$ is the
decomposition group of $w$ in $\ck/\ck'$. Since $G_w\simeq
G(\kappa(w)/\kappa(w'))$ and $\kappa(w')$ is a $\zp$--extension of
a finite field, $G_w$ is cyclic of order coprime to $p$. The first
isomorphism in (\ref{Delta-T}) combined with Shapiro's Lemma,
Hilbert's Theorem 90 and Herbrandt quotient theory (view
$\kappa(w)$ as a union of Galois extensions of Galois group $G_w$
of finite subfields of $\kappa(w')$) give
$$\widehat H^i(G, \Delta_{\ck, \ct})\simeq \bigoplus_{w'\in\ct'}\widehat H^i(G_w,
\kappa(w)^\times)=0\,,\qquad \forall\, i\in\z.$$ Consequently,
Tate cohomology applied to the first exact sequences above
gives
$$H^1(G, \kt)\simeq H^1(G, \ck_{(\ct)}^\times)\,.$$
The second isomorphism in (\ref{Delta-T}) combined with Shapiro's
Lemma gives
$${\mathcal Div}_{\ck'}(\ct')\simeq {\mathcal Div}_{\ck}(\ct)^G, \quad
H^1(G, {\mathcal Div}_{\ck}(\ct))\simeq
\bigoplus_{w'\in\ct'}H^1(G_w, \z)=0\,,$$ where the first
isomorphism above is induced by the canonical injection ${\mathcal
Div}_{\ck'}(\ct')\to {\mathcal Div}_{\ck}(\ct)$. Since
${div}_{\ck, \ct}\mid_{\ck'^\times}={div}_{\ck, \ct}$, Tate
cohomology applied to the second exact sequence above combined
with Hilbert's Theorem 90 gives
$$H^1(G,\ck_{(\ct)}^\times)\simeq H^1(G, \ck^\times)=0\,.$$
Consequently, we have
$$H^1(G, \kt)\simeq H^1(G, \ck_{(\ct)}^\times)\simeq H^1(G, \ck^\times)=0,$$
which concludes the proof of the first equality in (\ref{reduction}).
In order to prove the second equality in (\ref{reduction}), we consider the exact sequence
$$\xymatrix{1\ar[r] &\kstm\ar[r] &\kt\ar[r]^{\overline{div}_{\ck, \cs}\quad\qquad } &{\mathcal Div}_{\ck, \cs}\otimes\z/m\z}$$
in the category of $\z[G]$--modules, where ${\mathcal Div}_{\ck,
\cs}:=\oplus_{v\not\in\cs}\Gamma_v\cdot v$ is the group of
divisors in $\ck$ supported away from $\cs$ and $\overline{div}_{\ck,
\cs}(x):=(\sum_{v\not\in\cs}{\rm ord}_v(x)\cdot v)\otimes\widehat
1$, for all $x\in\kt$. Since $\ck/\ck'$ is unramified away from
$\cs$, the natural injective morphism ${\mathcal Div}_{\ck',\cs'}\to
{\mathcal Div}_{\ck, \cs}$ induces an isomorphism of groups ${\mathcal Div}_{\ck', \cs'}\simeq ({\mathcal Div}_{\ck, \cs})^G$ and the
restriction of $\overline{div}_{\ck, \cs}$ to $\kpt$ equals $\overline{div}_{\ck',
\cs'}$. Consequently, when we take $G$--invariants in the exact
sequence above, we obtain an exact sequence at the $\ck'$--level
$$\xymatrix{1\ar[r] &(\kstm)^G\ar[r] &\kpt\ar[r]^{\overline{div}_{\ck', \cs'}\quad\qquad } &{\mathcal Div}_{\ck', \cs'}\otimes\z/m\z},$$
which shows that $\kpstm=(\kstm)^G$, as desired.
\medskip
(2) The first isomorphism in part (2) of the Proposition is a direct consequence of part (1) combined with Remark \ref{reinterpret-extensions} and Corollary \ref{minus-reinterpret}.
The second isomorphism in part (2) is obtained from the first by taking the obvious projective limit.
\end{proof}
\bigskip
Let us assume that $\ck$ is of CM--type and $p$ is odd. Then, if we let $\fp$ denote the field with $p$ elements, we have an exact sequence of $\fp[G]$--modules
$$\xymatrix{1\ar[r] &\ca_{\ck, \ct}^-[p]\ar[r] &\cm[p]^-\ar[r] &({\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\z/p\z)^-\ar[r] &1. }
$$
According to Lemma \ref{class-group-coranks}, if we let $r_{\cm}^-:=\dim_{\fp}\,\cm[p]^-$, we have
\begin{equation}\label{rm-minus}
r_{\cm}^- =\lambda_{\ck}^- + \delta_{\ck, \ct}^- -\delta_{\ck} + d_{\ck, \cs}^-\,,
\end{equation}
where $\lambda_{\ck}^-$, $\delta_{\ck, \ct}^-$ and $\delta_{\ck}$ are as in Lemma \ref{class-group-coranks} and
$$d_{\ck, \cs}^-:=\dim_{\fp}\,({\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\z/p\z)^-.$$
We need a concrete formula for $d_{\ck, \cs}^-$. For that purpose, we let $J:=G(\ck/\ck^+)$ and we view ${\mathcal Div}_{\ck}(\cs\setminus\cs_p)$ as a $\z[J]$--module.
Obviously, $J$ is cyclic of order $2$, generated by $j$. Let $\cs^+$ and $\cs_p^+$ denote the sets of primes in $\ck^+$ sitting below primes
in $\cs$ and $\cs_p$, respectively. Also, for every $v\in\cs^+\setminus\cs_p^+$, let $J_v$ be the decomposition group of $v$ in $\ck/\ck^+$. Then, we have the following
obvious $\z[J]$--module and $\fp$--vector space isomorphisms, respectively:
$${\mathcal Div}_{\ck}(\cs\setminus\cs_p)\,\,\simeq\bigoplus_{v\in\cs^+\setminus\cs_p^+}\z[J/J_v], \quad ({\mathcal Div}_{\ck}(\cs\setminus\cs_p)\otimes\z/p\z)^-\simeq \bigoplus_{v\in\cs^+\setminus\cs_p^+}\fp[J/J_v]^-.$$
Obviously, we have $\fp[J/J_v]^-\simeq\fp$, if $J_v$ is trivial (i.e. if $v$ splits in $\ck/\ck^+$) and $\fp[J/J_v]^-=0$, otherwise. Consequently, we have the following formula
for $d_{\ck, \cs}^-$:
$$d_{\ck, \cs}^-={\rm card}\, \{v\mid v\in \cs^+\setminus\cs_p^+, \text{ $v$ splits completely in } \ck/\ck^+\}\,.$$
The above formula permits us to reformulate the main result in \cite{Kida} as follows.
\begin{theorem}[Kida]\label{Kida} Assume that $p$ is odd, $G$ is a $p$--group and $\ck$ and $\ck'$ are of CM--type. Then,
we have the following equality.
$$(\lambda_\ck^- - \delta_\ck + d_{\ck, \cs}^-) = |G|\cdot (\lambda_{\ck'}^- - \delta_{\ck'} + d_{\ck', \cs'}^-)\,.$$
\end{theorem}
\noindent The above result is known is ``Kida's formula''. It was first proved by Kida in \cite{Kida} and later and with different methods by Iwasawa (see \cite{Iwasawa-RH}) and
Sinnott (see \cite{Sinnott-Kida}.) This should be viewed as a ``partial'' number field analogue of Hurwitz's genus formula
for finite covers of smooth, projective curves over an algebraically closed field. Here, $\ca_{\ck}^-$ plays the role
of the Jacobian and its corank (equal to $\lambda_{\ck}^-$) plays the role of twice the genus of the underlying curve. Of course, a ``full'' number field
analogue of the Hurwitz genus formula should involve the ideal class--group $\ca_{\ck}$ and its corank $\lambda_{\ck}$. To our knowledge,
no such formula exists at the moment. The following consequence of the above theorem is of interest to us.
\begin{corollary}\label{rm-minus-G} Assume that $p$ is odd, $G$ is a $p$--group and $\ck$ and $\ck'$ are of CM--type. Then,
we have the following equality.
$$r_{\cm}^- = |G|\cdot r_{\cm'}^-\,.$$
\end{corollary}
\begin{proof} The first isomorphism in (\ref{Delta-T}) induces an isomorphism of $\zp[G]$--modules
$$(\Delta_{\ck, \ct}\otimes\zp)^-\simeq (\Delta_{\ck', \ct}\otimes\zp)^-\otimes_{\zp}\zp[G]\,.$$
Above, we took into account that since $G$ is a $p$--group, $G_w$
is trivial, for all $w\in\ct$.
Now, by comparing $p$--coranks on both sides of the above isomorphism, we
get
$$\delta_{\ck, \ct}^-=\mid G\mid\cdot\, \delta_{\ck', \ct'}^-\,.$$
The equality in the statement of the corollary is a direct consequence of the above equality combined
with (\ref{rm-minus}) and Theorem \ref{Kida}.
\end{proof}
\medskip
Now, we are ready to state and prove the main results of this section. The reader will note that in what follows we are using a strategy similar to
that of \S3 in \cite{GP}.
\begin{theorem}\label{free} Assume that $G$ is a $p$--group, $\ck$ and $\ck'$ are of CM--type, and $p$ is odd. Then, the following hold.
\begin{enumerate}
\item The $\fp[G]$--module $\cm[p]^-$ is free of rank
$$r_{\cm'}^-:=\lambda_{\ck'}^-+\delta_{\ck', \ct'}^{-} + d_{\ck', \cs'}^- -\delta_{\ck'}\,.$$
\item The $\zp[G]$--module $T_p(\cm)^-$ is free of rank $r_{\cm'}^-$.
\end{enumerate}
\end{theorem}
\begin{proof} For the proof of (1), we need the following (see \cite{Nakajima}, Proposition 2, \S4.)
\begin{lemma}[Nakajima]
Assume that $k$ is a field of characteristic $\ell\ne 0$, $G$ is a finite $\ell$--group and
$M$ is a finitely generated $k[G]$--module, such that
$$\dim_k M\geq |G|\cdot\dim_k M^G.$$
Then, $M$ is a free $k[G]$--module of rank equal to $\dim_k M^G$.
\end{lemma}
\noindent Now, if we combine the first isomorphism in Proposition \ref{invariants}(2) for $m:=p$, with Corollary \ref{rm-minus-G} above, we obtain the following.
$$\dim_{\fp} \cm[p]^- = \mid G\mid\cdot\dim_{\fp}(\cm[p]^-)^G\,.$$
The above Lemma applied to $\ell:=p$, $k:=\fp$ and $M:=\cm[p]^-$ implies that, indeed, $\cm[p]^-$ is a free $\fp[G]$--module whose rank satisfies
$${\rm rank}_{\fp[G]}\cm[p]^-=\dim_{\fp}(\cm[p]^-)^G=\dim_{\fp}\cm'[p]^-=r_{\cm'}^-\,.$$
This concludes the proof of part (1).
\smallskip
In order to prove part (2), we consider the exact sequence of $\zp[G]$--modules
$$\xymatrix{0\ar[r] &T_p(\cm)^-\ar[r]^{\times p} &T_p(\cm)^-\ar[r] &\cm[p]^-\ar[r] &0}
$$
where the injective morphism is the multiplication--by--$p$ map (see (\ref{Tate-torsion}).) Since $\cm[p]^-$ is a free $\fp[G]$--module, it is $G$--induced and therefore
$G$--cohomologically trivial. Consequently, when we apply Tate $G$--cohomology to the short exact sequence above, we obtain group--isomorphisms
$$\xymatrix{\widehat H^i(G, T_p(\cm)^-)\ar[r]^{\times p}_{\sim} &\widehat H^i(G, T_p(\cm)^-)
}$$
given by the multiplication--by--$p$ map, for all $i\in\z$. Since the cohomology groups above are $p$--groups, this implies that $\widehat H^i(G, T_p(\cm)^-)=0$.
Since $G$ is a $p$--group, this implies that $T_p(\cm)^-$ is cohomologically trivial (see Theorem 8, in \cite{Serre-Local}, Ch. IX, \S5.) Therefore, $T_p(\cm)^-$ is a projective $\zp[G]$--module, as it is
$\zp$--free (see loc.cit.) However, since $G$ is a $p$--group, the ring $\zp[G]$ is local and therefore $T_p(\cm)^-$ is a free $\zp[G]$--module (see \cite{Milnor}, Lemma 1.2.)
The $\fp[G]$--module isomorphism
$$T_p(\cm)^-=\fp[G]\otimes_{\zp[G]}T_p(\cm)^-$$
(a consequence of the short exact sequence above) combined with part (1) leads to
$${\rm rank}_{\zp[G]}T_p(\cm)^-={\rm rank}_{\fp[G]}\cm[p]^-=r_{\cm'}^-.$$
This concludes the proof of the Theorem.
\end{proof}
\begin{theorem}\label{projective} Assume that $\ck$ is of CM--type, $p$ is odd and $G$ is arbitrary. Then, if we let ``{\rm pd}'' denote ``projective dimension'', the following hold.
\begin{enumerate}
\item ${\rm pd}_{\zp[G]^-}T_p(\cm)^-={\rm pd}_{\zp[G]}T_p(\cm)^-=0.$
\item If, in addition, $\ck'$ is the $\zp$--cyclotomic
extension of a number field $k$, such that $\ck/k$ is
Galois of Galois group $\cg$ and $\cs$ and $\ct$ are
$\cg$--invariant, then ${\rm
pd}_{\zp[[\cg]]^-}T_p(\cm)^-={\rm
pd}_{\zp[[\cg]]}T_p(\cm)^-=1$, unless $T_p(\cm)^-=0$.
\end{enumerate}
\end{theorem}
\begin{proof}
(1) Since $T_p(\cm)^-$ is a free $\zp$--module, the statement in part (1) of the Theorem is equivalent to the $G$--cohomological triviality
of $T_p(\cm)^-$. (See \cite{Serre-Local}, Ch. IX, \S5.) In order to prove that, let $H$ denote a $p$--Sylow subgroup of $G$ and let $\ck^H$ be the maximal subfield
of $\ck$ fixed by $H$. Then, $\ck/\ck^H$ is Galois, of Galois group $H$. Also, $\ck^H$ is of CM--type. Indeed, if $\ck^H\subseteq \ck^+$, then
$2=[\ck:\ck^+]$ would divide $\mid H\mid=[\ck:\ck^H]$, which is false, because $p$ is odd. Consequently, we can apply Theorem \ref{free}(2) to the extension
$\ck/\ck^H$ to conclude that
$$\widehat H^i(H, T_p(\cm)^-)=0,\quad \forall\, i\in\z\,.$$
Since this happens for any $p$--Sylow subgroup $H$ of $G$, the module $T_p(\cm)^-$ is $G$--cohomologically trivial and therefore
$\zp[G]$--projective, as desired. (See loc.cit.)
(2) Let $\Gamma_k:=G(\ck'/k)$. Since $\Gamma_k$ is a free abelian topological group,
the usual Galois--restriction exact sequence
\begin{equation}\label{groups-split}
\xymatrix{1\ar[r] &G\ar[r] &\cg\ar[r]^{\rm res_{\ck/\ck'}} &\Gamma_k\ar[r] &1}
\end{equation}
is split. Pick a splitting and let $\Gamma$ be its image in $\cg$. Then, we have
a group isomorphism $\cg\simeq G\rtimes \Gamma$, where $\rtimes$ denotes a semi-direct product. Let $K$ be the normal closure (over $k$) in $\ck$ of
$\ck^\Gamma$ (the maximal subfield of $\ck$ fixed by $\Gamma$.) Then $K/k$ is a Galois extension of number fields
and $\ck=K\cdot\ck'$ (the compositum of $K$ and $\ck'$ inside $\ck$.) Consequently, if we let $G':=G(K/k)$,
Galois restriction induces an isomorphism between $G$ and the normal subgroup $G(K/\ck'\cap K)$ of $G'$. Via this isomorphism,
we identify $G$ with $G(K/\ck'\cap K)$ and $G'/G$ with $G(\ck'\cap K/k)$. We have an exact sequence
of groups induced by the obvious Galois restriction maps
\begin{equation}\label{groups}
\xymatrix{1\ar[r] &\cg\ar[r] &G'\times\Gamma\ar[r] & G'/G\ar[r] &1}.
\end{equation}
Consequently, $\zp[[G'\times\Gamma]]$ viewed as a left or right $\zp[[\cg]]$--module in the obvious way is free
of basis given by any complete set of representatives for the left cosets of $G$ in $G'$. We consider the
left $\zp[[G'\times\Gamma]]$--module
$$\widetilde{T_p(\cm)^-}:= \zp[[G'\times\Gamma]]\otimes_{\zp[[\cg]]}T_p(\cm)^-.$$
As $\zp[[G'\times\Gamma]]$ is a free right $\zp[[\cg]]$--module, (2) in the Theorem is equivalent
to
$${\rm pd}_{\zp[[G'\times\Gamma]]}\widetilde{T_p(\cm)^-}=1.$$
We claim that this follows directly
from part (1) of the Theorem combined with Proposition 2.2 and Lemma 2.3 in \cite{Popescu-CS}. In order to see this, let
us first note that we have an obvious ring isomorphism $\zp[[G'\times\Gamma]]\simeq\Lambda[G']$, where $\Lambda:=\zp[[\Gamma]]$.
Now, we have the following.
\begin{lemma}\label{projective-dimension} A finitely generated $\Lambda[G']$--module
$M$ satisfies ${pd}_{\Lambda[G']}M\leq 1$ if and only if the following conditions are satisfied.
\begin{enumerate}\item [(i)] ${\rm pd}_{\zp[G']}M\leq 1$.
\item[(ii)] $M$ has no non trivial finite $\Lambda$--submodules (i.e. ${\rm pd}_{\Lambda}M\leq 1$.)
\end{enumerate}
\end{lemma}
\begin{proof} Combine Proposition 2.2 and Lemma 2.3 in \cite{Popescu-CS}.\end{proof}
\noindent Since the $\Lambda$--module $\widetilde{T_p(\cm)^-}$ is $\zp$--free, it satisfies (ii) above automatically.
Also, as a consequence of the definitions and exact sequence (\ref{groups}), we have an isomorphism
of left $\zp[G']$--modules
$$\widetilde{T_p(\cm)^-}\simeq \zp[G']\otimes_{\zp[G]}{T_p(\cm)^-}.$$
Consequently, part (1) of the Theorem implies that ${\rm pd}_{\zp[G']}\widetilde{T_p(\cm)^-}=0$.
The above Lemma implies that ${\rm pd}_{\Lambda[G']}\widetilde{T_p(\cm)^-}\leq 1$. However, since
$\widetilde{T_p(\cm)^-}$ is of finite $\zp$--rank, we must have ${\rm pd}_{\Lambda[G']}\widetilde{T_p(\cm)^-}=1$.
This concludes the proof of part (2) in the Theorem.
\end{proof}
\medskip
\begin{corollary}\label{coinvariants} Assume that $\ck$ and $\ck'$ are of CM--type,
$p$ is odd and $G$ is arbitrary. Then, there is a canonical isomorphism of $\zp$--modules
$$T_p(\cm)^-/I_GT_p(\cm)^-\simeq T_p(\cm')^-,$$
where $I_G$ is the usual augmentation ideal in $\zp[G]$.
\end{corollary}
\begin{proof}
Since $T_p(\cm)^-$ is $\zp[G]$--projective, it is $G$--cohomologically trivial. In particular,
$\widehat H^0(G, T_p(\cm)^-)=\widehat H^{-1}(G, T_p(\cm)^-)=0$. As a consequence, if we denote by $N_G$ the usual
norm element in $\zp[G]$, we have
$$(T_p(\cm)^-)^G=N_G\cdot T_p(\cm)^-,\qquad T_p(\cm)^-/I_GT_p(\cm)^-\simeq N_G\cdot T_p(\cm)^-\,,$$
where the isomorphism of $\zp$--modules to the right is induced by multiplication with $N_G$. Now, the corollary is a direct consequence of
Proposition \ref{invariants}, part (2).
\end{proof}
\section{The Equivariant Main Conjecture}\label{EMC}
We begin by recalling two classical theorems on special values of global $L$--functions. Let $K/k$ be an abelian extension
of number fields of Galois group $G$. Throughout, if $\mathcal C$ is an algebraically closed field, we denote by $\widehat G(\mathcal C)$ the group
of irreducible $\mathcal C$--valued characters of $G$. Let $S$ be a finite set of primes in $k$, which contains
the set $S_\infty(k)\cup S_{\rm ram}(K/k)$ of primes which are either infinite or ramify in $K/k$. For every $\chi\in\widehat G(\C)$, we let $L_S(\chi, s)$ denote the $\C$--valued $S$--incomplete Artin $L$--function
associated to $\chi$, of complex variable $s$. These $L$--functions are holomorphic everywhere, except for a simple pole at $s=1$ when $\chi=\mathbf 1_G$. We let
$$\Theta_{S, K/k}:=\sum_{\chi\in\widehat G(\C)}L_S(\chi^{-1}, s)\cdot e_\chi, \qquad \Theta_{S, K/k}: \C\to \C[G]$$
denote the so--called $S$--incomplete $G$--equivariant $L$--function associated to $K/k$, where $e_{\chi}:=1/|G|\sum_{\sigma\in G}\chi(\sigma)\cdot\sigma^{-1}$
is the idempotent associated to $\chi$ in $\C[G]$.
\begin{theorem}[Siegel \cite{Siegel}]\label{siegel-theorem} For all $n\in\z_{\geq 1}$, we have
$$\Theta_{S, K/k}(1-m)\in\q[G].$$
\end{theorem}
\noindent As usual, for every abelian group $M$ and every $m\in\z$, we let $M(m)$ denote the group $M$ endowed with the action by the absolute Galois group
$G_k$ of $k$ given by the $n$--th power of the cyclotomic character $c_k: G_k\to {\rm Aut}(\bmu_\infty)\simeq{\widehat\z}^\times$.
\begin{theorem}[Deligne-Ribet \cite{Deligne-Ribet}, Cassou-Nogu\`es \cite{Cassou-Nogues}]\label{deligne-ribet-theorem} For all $m\in\z_{\geq 1}$, we have
$${\rm Ann}_{\z[G]}(\q/\z(m)^{\, G_K})\cdot\Theta_{S, K/k}(1-m)\subseteq \z[G],$$
where $\q/\z(m)^{\, G_K}$ denotes the $\z[G]$--module of $G_K$--invariants of $\q/\z(m)$.
\end{theorem}
\noindent Now, let $T$ be a finite, non-empty set of finite primes in $k$, disjoint from $S$. We let
$$\delta_{T, K/k}:=\prod_{v\in T}(1-\sigma_v^{-1}\cdot({\bf N}v)^{1-s}), \qquad \delta_{T, K/k}:\C\to\C[G],$$
where $\sigma_v$ denotes the Frobenius morphism associated to $v$ in $G$.
\begin{corollary}\label{Deligne-Ribet-rewrite} Let $p$ be a prime number and $m\in\z_{\geq 1}$. Assume that either $T$ contains at least a prime which does not sit above $p$ or
$p$ does not divide the cardinality of (the finite group) $\q/\z(m)^{\, G_K}$. Then, we have
$$\delta_{T, K/k}(1-m)\cdot\Theta_{S, K/k}(1-m)\in\z_{(p)}[G]\,.$$
\end{corollary}
\begin{proof} Remark that if $v\in T$ does not sit above $p$, then $\sigma_v$ acts on the group $\q/\z(m)^{\, G_K}\otimes\zp\simeq \q_p/\z_p(m)^{\, G_K}$ via
multiplication by $({\bf N}v)^{m}$. This implies that, under the hypotheses of the Corollary, we have
$$\delta_{T, K/k}(1-m)\in {\rm Ann}_{\z_{(p)}[G]}(\qp/\zp(m)^{\, G_K}).$$
Now, the Corollary follows from Theorem \ref{deligne-ribet-theorem}.
\end{proof}
\begin{definition} For $S$ and $T$ as above, the $S$--incomplete $T$--modified $G$--equivariant $L$--function
associated to $K/k$ is defined by
$$\Theta_{S, T, K/k}:=\delta_{T, K/k}(s)\cdot\Theta_{S, K/k}(s), \qquad \Theta_{S, T, K/k}: \C\to\C[G].$$
\end{definition}
\begin{remark}\label{theta-st-remark} Note that $\Theta_{S, T, K/k}$ is holomorphic everywhere on $\C$ and
$$\Theta_{S, T, K/k}(1-m)\in\z_{(p)}[G],$$
whenever the integer $m$, the prime $p$ and the set $T$ satisfy the hypotheses of the Corollary above. In particular, if $T$ contains a prime of
residual characteristics coprime to the cardinality of $(\q/\z(m)^{G_K})$, for some $m\in\z_{\geq 1}$, then
$$\Theta_{S, T, K/k}(1-m)\in\z[G].$$
Also, if $T$ contains two primes of distinct residual characteristics, then
$$\Theta_{S, T, K/k}(1-m)\in\z[G],\qquad \forall\, m\in\z_{\geq 1}.$$
\end{remark}
\medskip
Throughout the rest of this section, we fix an odd prime $p$ and an abelian extension $\ck/k$, where $\ck$ is a CM $\zp$--field and $k$ is a totally real number field.
Let $\cg:={\rm Gal}(\ck/k)$. We fix two finite, non--empty,
disjoint sets $S$ and $T$ of primes in $k$, such that $S$ contains $S_{\rm ram}(\ck/k)\cup S_\infty$. Note that, in particular,
$S$ contains the set $S_p$ of $p$--adic primes of $k$.
We let $\cs$ and $\ct$ denote
the sets of finite primes in $\ck$ sitting above primes in $S$ and $T$, respectively.
The main goal of this section is the proof of an Equivariant Main Conjecture for the data $(\ck/k, S, T, p)$. This statement expresses the first Fitting invariant
${\fit}_{\zp[[\cg]]^-}(T_p(\mk)^-)$ of the finitely generated module $T_p(\mk)^-$ over the (Noetherian, commutative) ring
$$\zp[[\cg]]^-:=\zp[[\cg]]/(1+j)$$
in terms of a certain equivariant $p$--adic $L$--function $\Theta_{S,T}^{(\infty)}$, which is an element of $\zp[[\cg]]^-$
canonically associated to the
data $(\ck/k, S, T, p)$. As usual, $j$ denotes the complex conjugation automorphism of $\ck$, viewed as an element
of $\cg$. The precise statement, to be proved in \S\ref{emc-section} below, is the following.
\begin{theorem}[The Equivariant Main Conjecture]\label{emc} Under the above hypotheses, we have the following equality of ideals in $\zp[[\cg]]^-$.
$${\rm Fit}_{\zp[[\cg]]^-}(T_p(\mk)^-)=(\Theta_{S, T}^{(\infty)}).$$
\end{theorem}
Let $\ck'$ be the cyclotomic $\zp$--extension of $k$ and $\Gamma_k:=G(\ck'/k)$.
As in the proof of Theorem \ref{projective}(2), the exact sequence
$$\xymatrix{1\ar[r] &G(\ck/\ck')\ar[r] &\cg\ar[r]^{\rm res_{\ck/\ck'}} &\Gamma_k\ar[r] &1
}$$
is split. We fix a splitting, i.e. we fix a subgroup $\Gamma$ of $\cg$ which is mapped isomorphically
to $\Gamma_k$ by ${\rm res_{\ck/\ck'}}$. We let $K:=\ck^\Gamma$ denote the maximal subfield of $\ck$ fixed by $\Gamma$.
Then, Galois theory combined with the fact that $\cg$ is abelian, implies that $K/k$ is Galois, $K\cdot\ck'=\ck$
and $\ck'\cap K=k$. If we let $G:=G(K/k)$, Galois restriction induces group isomorphisms $\cg\simeq G\times\Gamma_k$, $\Gamma\simeq\Gamma_k$, and $G(\ck/\ck')\simeq G$.
Below, we freely identify these groups via these isomorphisms.
Obviously, $K$ is a CM number field whose cyclotomic $\zp$--extension is $\ck$. We let $j$ denote the complex conjugation automorphism of $K$ as well.
We fix a topological generator $\gamma$ of $\Gamma$ and identify it via the Galois restriction
isomorphism with a generator of $\Gamma_k$. For any finite extension $\co$ of $\zp$, we have isomorphisms of compact $\co[G]$--algebras
\begin{equation}\label{power-series-iso}\co[[\cg]]\simeq \co[G][[\Gamma]]\simeq \co[G][[t]],\qquad \co[[\cg]]^\pm\simeq \co[G]^\pm[[\Gamma]]\simeq \co[G]^\pm[[t]].\end{equation}
Above, $\co[G]^\pm:=\co[G]/(1\mp j)$, $t$ is a variable, and the right-most isomorphisms send $\gamma$ to $(t+1)$, as usual.
\subsection{The work of Wiles on the Main Conjecture \cite{Wiles}} In order to simplify notations, in this section we assume that $\bmu_p\subseteq K$ (i.e. $\bmu_{p^\infty}\subseteq\ck$.)
Let $c:=\omega\cdot\kappa$ be the decomposition of the $p$--cyclotomic character $c:=c_p$ of $\cg$ into its Teichm\"uler component $\omega:=\omega_p$ and
its complement $\kappa:=\kappa_p$, as in \S\ref{appendix-twisting} in the Appendix. Note that, in this case, $\omega$ and $\kappa$ factor through $G$ and $\Gamma$, respectively. Let $u$ be the element in $(1+p\zp)$ given by
$u:=\kappa(\gamma)$.
We fix an embedding $\Bbb C\hookrightarrow\cp$. Via this embedding we identify $\widehat G(\Bbb C)$ and $\widehat G(\cp)$. A character $\psi\in\widehat G(\cp)$ is called even if $\psi(j)=1$ and odd otherwise. Let $\co$ be a fixed finite extension of $\zp$ which contains the values of all $\psi\in\widehat G(\cp)$. We fix a uniformizer $\pi$ in $\co$ and denote by $Q(\co)$ the field of fractions of $\co$.
In \cite{Deligne-Ribet}, Deligne and Ribet proved that for every character $\psi\in\widehat G(\cp)$, there exist power series
$G_{\psi, S}(t)$ and $H_{\psi, S}(t)$ in $\co[[t]]$, uniquely determined by the following properties.
\begin{equation}\label{power-series}
H_{\psi, S}=\left\{
\begin{array}{ll}
t, & \hbox{if $\psi=\boldsymbol 1_G$;} \\
1, & \hbox{otherwise;}
\end{array}
\right.
\qquad\quad
\frac{G_{\psi, S}(u^m-1)}{H_{\psi, S}(u^m-1)}=L_S(\psi\omega^{-m}, 1-m),\quad \forall\, m\in\z_{\geq 1.}
\end{equation}
The reader may consult \S1 in \cite{Wiles}
and \S4 in \cite{Popescu-CS} for the properties above and note that, since $K\cap\ck' =k$,
all the non trivial characters of $G$ are of ``type S'', in the terminology used in loc.cit.
By definition, the $S$--incomplete $p$--adic $L$--function $L_{p, S}(\psi, s)$ associated to an even character
$\psi\in\widehat G(\mathbb C)$ is given by
\begin{equation}\label{p-adicL} L_{p, S}(\psi, 1-s)=\frac{G_{\psi, S}(u^s-1)}{H_{\psi, S}(u^s-1)},\qquad s\in\zp.\end{equation}
The function $L_{p, S}(\psi, s)$ is $p$--adically analytic everywhere, except for a possible pole of order $1$
at $s=1$, if $\psi=\boldsymbol 1_G$.
\begin{remark}\label{odd-characters} As an immediate consequence of the functional equation for the
global $L$--functions $L_S(\chi, s)$ considered above, we have $L_S(\chi, 1-m)=0$, whenever $\chi$ and $1-m$ have the same parity, for all $m\in\z_{\geq 1}$ and all $\chi\in\widehat G(\C)$.
Assume that if $\psi$ is an odd character.
Then $\psi\omega^{-m}$ and $1-m$ have the same parity, for all $m\in\Bbb Z_{\geq 1}$. Consequently,
we have $L_S(\psi\omega^{-m}, 1-m)=0$, for all $m\in\z_{\geq 1}$. Therefore, $G_{\psi, S}(t)=0$ and $L_{p, S}(\psi, s)=0$, for all odd characters $\psi.$
\end{remark}
For simplicity, if $L$ is a subfield of $K$ containing $k$, we denote by $\cx_{L}$ the Galois
group of the maximal abelian pro--$p$ extension of $\cl:=L\cdot\ck'$ (the cyclotomic $\zp$--extension of
$L$), which is unramified away from the primes above those in $S$. We view $\cx_{L}$ as a module
over $\zp[[G(\cl/k)]]\simeq\zp[G(L/k)][[\Gamma]]$ (and consequently over $\zp[[\cg]]$) in the usual way.
Note that $\cx_K$ is the module we denoted by $\cxs$ in \S\ref{classical}. We will continue to use $\cxs$ instead of $\cx_K$ in what follows. The modules $\cx_{L}$ and $\cx_{L}^+$ are finitely generated, respectively
torsion and finitely generated modules over $\Lambda:=\zp[[\Gamma]]$ (a classical theorem of Iwasawa \cite{Iwasawa-zl}.)
Let $\psi\in\widehat G(\cp)$ be an even character. Let $K_\psi:= K^{\ker\psi}$ be the largest subfield of $K$ fixed by the kernel of $\psi$. We have $G(K_\psi/ k)\simeq F\times H$,
where $H$ is the $p$--Sylow subgroup of $G(K_\psi/k)$ and $F$ is its maximal subgroup of order coprime to $p$. Consequently, $\psi$ (viewed as a
character of $G(K_\psi/ k)$) splits as $\psi=\varphi\cdot\rho$, with $\varphi$ a faithful character of $R$ (of order coprime to $p$) and $\rho$ a faithful character of $H$ (of order
a power of $p$.) Let $h$ be a generator of $H$ and assume that ${\rm ord}(h)=p^n$, for some $n\in\Bbb Z_{\geq 0}$. Let $e_{\varphi}:=1/|F|\sum_{f\in F}\varphi(f)\cdot f^{-1}$ be the
idempotent associated to $\varphi$. Note that $e_{\varphi}\in \co[F]$. The following definition of ``character $\mu$--invariants'' of $\cxs$ is due
to Greenberg (see \cite{Neukirch}, pp. 656--657.)
\begin{definition}[Greenberg] Let $\psi\in\widehat G(\cp)$ be an even character. Then $\mu_{\psi, \co}(\cxs)$ is defined to be
the $\mu$--invariant $\mu_\co(\cxs^\psi)$ of the $\co[[\Gamma]]$--module
$$\cxs^\psi:=\left\{
\begin{array}{ll}
(h^{p^{n-1}}-1)e_\varphi(\cx_{K_\psi}\otimes_{\zp}\co), & \hbox{if $n\geq 1$;} \\
e_\varphi(\cx_{K_\psi}\otimes_{\zp}\co), & \hbox{otherwise.}
\end{array}
\right.
$$
\end{definition}
\noindent Note that if the character $\psi$ has order coprime to $p$, i.e. $\psi=\varphi$ and $n=0$ in the above notation,
then we have
$$e_\varphi(\cx_{K_\psi}\otimes_{\zp}\co)=\{x\in (\cx_{K_\psi}\otimes_{\zp}\co)\mid gx=\psi(g)x, \forall g\in G(K_\psi/k)\}.$$
This reconciles the definition above with Wiles's definition (see \cite{Wiles}, p. 497) of ``character $\mu$--invariants''
for characters of order coprime to $p$. (See also Definition 11.6.14 in \cite{Neukirch}.)
The following is the statement of the classical (non--equivariant) Main Conjecture for the module $\cxs^+$, proved by Wiles
in \cite{Wiles}. Below, we denote by $\mathfrak m_\gamma$ the multiplication-by-$\gamma$ automorphism of all the relevant vector spaces.
\begin{theorem}[Wiles]\label{Wiles} For every even character $\psi\in\widehat G(\C)$, we have
$$G_{\psi, S}(t) \sim \pi^{\mu_{\psi, \co}(\cxs)}\cdot{\rm det}_{Q(\co)}((t+1)-\mathfrak m_\gamma\mid e_{\psi}(\cxs\otimes_{\zp}Q(\co))),$$
where $\sim$ denotes association in divisibility in the ring $\co[[t]]$ and $e_{\psi}$ is the idempotent associated to $\psi$ in $Q(\co)[G]$.
\end{theorem}
\begin{proof} (Sketch.) For characters $\psi$ of order coprime to $p$, the statement above is the combination of Theorems 1.3 and 1.4 in \cite{Wiles}.
For arbitrary characters, \cite{Wiles} contains a proof of the statement above only up to a power of $\pi$ (Theorem 1.3. loc.cit.)
However, once one has the right definition of $\mu_{\psi, \co}(\cxs)$ (see above), one can deduce the statement above for arbitrary
characters $\psi$ without much difficulty from Theorems 1.3 and 1.4 in \cite{Wiles} restricted to the case of the trivial character and base fields $K_\psi$ and $K_{\psi^p}$, respectively (see Theorem 11.6.16 in \cite{Neukirch} for a proof.) \end{proof}
\begin{remark}\label{Wiles-Fitting}
As proved by Greenberg (see Proposition 1 in \cite{Greenberg}), we have an equality of monic polynomials in $\co[t]$
$${\rm det}_{Q(\co)}((t+1)-\frak m_\gamma\mid e_{\psi}(\cxs\otimes_{\zp}Q(\co)))={\rm det}_{Q(\co)}((t+1)-\mathfrak m_\gamma\mid e_{\psi}(\cx_{K_\psi}\otimes_{\zp}Q(\co))),$$
for all even $\psi$. Also, it is easy to see that
$$e_{\psi}(\cx_{K_\psi}\otimes_{\zp}Q(\co))=\cxs^\psi\otimes_\co Q(\co)\,,$$
for all even $\psi$. Consequently, the right hand-side of the equivalence $\sim$ in Theorem \ref{Wiles} is precisely the characteristic polynomial
${\rm char}_{\co[[\Gamma]]}(\cxs^\psi)$ of the $\co[[\Gamma]]$--module $\cxs^\psi$. On the other hand, by definition, $\cxs^\psi$ is an
$\co[[\Gamma]]$--submodule of $(\cx_{K_\psi}\otimes_{\zp} \co)$. As such, it is a torsion $\co[[\Gamma]]$--module with no finite non trivial $\co[[\Gamma]]$--submodules (a classical theorem of Iwasawa.)
Consequently, ${\rm pd}_{\co[[\Gamma]]} \cxs^\psi=1$ (see Lemma \ref{projective-dimension}) and
its first Fitting ideal ${\rm Fit}_{\co[[\Gamma]]}(\cxs^\psi)$ is principal, generated by ${\rm char}_{\co[[\Gamma]]}(\cxs^\psi)$ (see Lemma 2.4 in \cite{Popescu-CS} and also Proposition \ref{fitting-calculation}(1) in the Appendix.)
This leads to the following equivalent, and perhaps more elegant formulation of Theorem \ref{Wiles}:
$${\rm Fit}_{\co[[\Gamma]]}(\cxs^\psi)=(G_{\psi, S}(t)),\qquad \text{for all even $\psi$},$$
where $(G_{\psi, S}(t))$ denotes the principal ideal of $\co[[\Gamma]]$ generated by $G_{\psi, S}(t)$.
In light of this reformulation,
it would be natural to expect that a $G$--equivariant refinement of Theorem \ref{Wiles} would give the Fitting ideal
${\rm Fit}_{\zp[[\cg]]^+}(\cxs^+)$ in terms of an equivariant $p$--adic $L$--function. Unfortunately, in general, the $\zp[[\cg]]^+$--module
$\cxs^+$ has infinite projective dimension, which makes its Fitting ideal non--principal and difficult to compute. This is the main reason why
the module $\cxs^+$ is replaced with $T_p(\mk)^-$, a $\zp[[\cg]]^-$--module of projective dimension $1$ (see Theorem \ref{projective} above), whose Fitting
ideal is principal (see Proposition \ref{fitting-calculation}(1) in the Appendix). Moreover, $T_p(\mk)^-$ has a quotient isomorphic to $(\cxs^+)^\ast(1)$ (see Remark \ref{remark-link-classical}.)
\end{remark}
\begin{corollary}\label{Wiles+zero-mu} If the $\mu$--invariant $\mu_{\zp}(\cxs^+)$ of the $\zp[[\Gamma]]$--module $\cxs^+$ is $0$, then the following hold
for all even characters $\psi\in\widehat G(\cp)$.
\begin{enumerate}
\item $\mu_{\psi, \co}(\cxs)=0.$
\item $G_{\psi, S}(t) \sim {\rm det}_{Q(\co)}((t+1)-\mathfrak m_\gamma\mid e_{\psi}(\cxs\otimes_{\zp}Q(\co)))$ in $\co[[t]]$.
\end{enumerate}
\end{corollary}
\begin{proof}
(2) is a consequence of (1) and Theorem \ref{Wiles}.
Part (1) is proved as follows. It is easily seen that Galois restriction gives an isomorphism of
$\zp[[\cg]]$--modules $\cxs^+\simeq \cx_{K^+}$, where $K^+$ is the maximal real subfield
of $K$. Also, since $\psi$ is even, we have $K_\psi\subseteq K^+$ and Galois restriction
leads to an exact sequence of $\zp[[\cg]]$--modules
$$
\xymatrix{\cx_{K^+}\ar[r] &\cx_{K_\psi}\ar[r] & G(K_\psi\cap\ck'/k)\ar[r] &1
}.$$
Since $G(K_\psi\cap\ck'/k)$ is finite and $\mu_{\zp}(\cxs^+)=0$, we have $\mu_{\zp}(\cx_{K_\psi})=0$.
Consequently, we have $\mu_\co(\cx_{K_\psi}\otimes_{\zp} \co)=0$. Therefore $\mu_\co(e_{\varphi}(\cx_{K_\psi}\otimes_{\zp} \co))=0$
and $\mu_{\psi, \co}(\cxs)=0$. We have used the fact that the $\mu$--invariant of a quotient or submodule
of an Iwasawa module is at most equal to that of the module itself.
\end{proof}
\subsection{The relevant equivariant $p$--adic $L$--functions}\label{equivariant-L-functions} For simplicity, we will continue to assume that $\boldsymbol\mu_p\subseteq K$,
unless otherwise stated. If $R$ is a commutative ring with $1$, we denote by $Q(R)$ the total ring of fractions of $R$. By abusing notation,
we denote by $\iota, t_m: Q(\co[[\cg]])\simeq Q(\co[[\cg]])$ the unique $Q(\co)$--algebra automorphisms obtained by extending to $Q(\co)[[\cg]]$ the $\zp$--algebra
automorphisms $\iota, t_m: \zp[[\cg]]\simeq\zp[[\cg]]$ defined in \S\ref{appendix-twisting} of the Appendix, for all $m\in\z$.
As usual, we freely identify $\co[[\cg]]$
and $Q(\co[[\cg]])$ with $\co[G][[t]]$ and $Q(\co[G][[t]])$, respectively, via the first isomorphism in (\ref{power-series-iso}).
By abusing notation once again, we let $\psi:\co[G][[t]]\to \co[[t]]$ denote
the unique $\co[[t]]$--algebra morphism extending $\psi: G\to \co$, for all $\psi\in\widehat G(\cp)$.
\begin{remark}
Note that the non zero--divisors of $\co[G][[t]]$ (respectively $\co[G]$) are precisely those elements $f\in \co[G][[t]]$ (respectively $f\in \co[G]$)
with the property that $\psi(f)\ne 0$ in $\co[[t]]$ (respectively in $\co$),
for all $\psi\in\widehat G(\cp)$.
\end{remark}
\noindent We consider the following power series in $\frac{1}{|G|}\co[G][[t]]$:
$$G_S(t):=\sum_{\psi\in\widehat G(\cp)}G_{\psi, S}(t)\cdot e_\psi, \quad H_S(t):=\sum_{\psi\in\widehat G(\cp)}H_{\psi, S}(t)\cdot e_\psi=t\cdot e_{\boldsymbol 1_G}+(1-e_{\boldsymbol 1_G}).$$
Observe that, for all $m\in\z$, we have equalities
$$\iota(e_\psi)=e_{\psi^{-1}}, \quad t_m(e_\psi)=e_{\psi\omega^{-m}},\quad \iota(t)=(1+t)^{-1}-1,\quad t_m(t)=u^m(1+t)-1,$$
for all characters $\psi\in\widehat G(\cp)$. Consequently, we have the following, for all $m\in\z$.
\begin{eqnarray}
\nonumber (\iota\circ t_m)(G_S(t)) &=& \sum_{\psi\in\widehat G(\cp)}G_{\psi^{-1}\omega^m, S}(u^m(1+t)^{-1}-1)\cdot e_\psi, \\
\nonumber (\iota\circ t_m)(H_S(t)) &=& (u^m(1+t)^{-1}-1)\cdot e_{\omega^m} + (1-e_{\omega^m}).
\end{eqnarray}
For every $n\in\z_{\geq 0}$, we let $K_n$ denote the unique field, with $K\subseteq K_n\subseteq \ck$ and $[K_n:K]=p^n$. We let $G_n:=G(K_n/k)$ and note that
$G_n\simeq G\times \Gamma/\Gamma^{p^n}$. We denote by $\pi_n: \co[[\cg]]\twoheadrightarrow \co[G_n]$ the usual (surjective) $\co$--algebra morphism induced by Galois restriction.
By applying the remark above to the group rings $\co[[\cg]]$ and $\co[G_n]$, it is easily seen that $|G|(\iota\circ t_m)(H_S)$ is not a zero--divisor in $\co[[\cg]]$, for all $m\ne 0$.
Moreover,
$\pi_n(|G|(\iota\circ t_m)(H_S))$ is not a zero--divisor in $\co[G_n]$, for all $n$. Let
$$\vu:=\{f\in \co[[\cg]]\mid \pi_n(f)\text{ not a zero--divisor in }\co[G_n], \forall\, n\}, \qquad \vu_n:=\pi_n(\vu)\,.$$
It is easily proved that $\vu_n$ is in fact the set of all non zero--divisors of $\co[G_n]$. As a consequence, we have an equality $\vu_n^{-1}\co[G_n]=Q(\co[G_n])=Q(\co)[G_n]$.
The morphisms $\pi_n$ above can be uniquely extended to surjective $Q(\co)$--algebra morphisms (for simplicity, also denoted) $\pi_n: \vu^{-1}\co[[\cg]]\twoheadrightarrow Q(\co[G_n])$, whose projective
limit gives an injective $Q(\co)$--algebra morphism
$$\vu^{-1}\co[[\cg]]\hookrightarrow\underset{n}{\underset{\longleftarrow}\lim}\, Q(\co[G_n]).$$
Next, we consider the element
$$g_S:=\frac{(\iota\circ t_1)(G_S(t))}{(\iota\circ t_1)(H_S(t))}=\frac{|G|(\iota\circ t_1)(G_S(t))}{|G|(\iota\circ t_1)(H_S(t))}\in \vu^{-1}\co[[\cg]],$$
and describe its images $\pi_n(g_S)$ in $Q(\co[G_n])$, for all $n$. In order to do that, let
$$\Theta_S^{(n)}(s):=\Theta_{S, K_n/k}(s),\qquad \forall\, n\in\z_{\geq 0}$$
be the $S$--incomplete $G_n$--equivariant $L$--function associated to $K_n/k$. According to Siegel's Theorem \ref{siegel-theorem}, we have
$$\Theta_S^{(n)}(1-m)\in\q[G_n]\subseteq Q(\co[G_n]), \qquad \forall\, n\in\z_{\geq 0}, \quad \forall\, m\in\z_{\geq 1}.$$
Moreover, the inflation property of Artin $L$--functions implies that
$$(\Theta_S^{(n)}(1-m))_n\in\underset{n}{\underset{\longleftarrow}\lim}\, Q(\co[G_n]),\qquad \forall\, m\in\z_{\geq 1}.$$
We let $\Theta_S^{(\infty)}(1-m):=(\Theta_S^{(n)}(1-m))_n$ and view it as an element of $\underset{n}{\underset{\longleftarrow}\lim}\, Q(\co[G_n])$.
\begin{lemma}\label{twisting-gs} For all $m\in\z_{\geq 1}$, we have.
\begin{enumerate}\item $t_{1-m}(g_S)\in\vu^{-1}\co[[\cg]].$
\item $\Theta_S^{(\infty)}(1-m)=t_{1-m}(g_S)$ in $\vu^{-1}\co[[\cg]]$.
\end{enumerate}\end{lemma}
\begin{proof}
Part (1) follows from the observation that the element
$$t_{1-m}(|G|(\iota\circ t_1)(H_S))=|G|(\iota\circ t_m)(H_S)$$
belongs to $\vu$, for all $m\in\z_{\geq 1}$. Part (2) is equivalent to
$$\pi_n(t_{1-m}(g_S))=\Theta^{(n)}_S(1-m),$$
for all $m\in\z_{\geq 1}$ and all $n\in\z_{\geq 0}$.
This is Proposition 4.1 in \cite{Popescu-CS}. Note that in loc.cit. our power series
$(\iota\circ t_1)(G_S(t))$ and $(\iota\circ t_1)(H_S(t))$ are denoted $G_S(t)$ and $H_S(t)$, respectively.
However, $g_S$ denotes the same element in $\vu^{-1}\co[[\cg]]$.
\end{proof}
Next, we use the additional set of primes $T$ to eliminate the denominators of the elements $t_{1-m}(g_S)$ in
$\vu^{-1}\co[[\cg]]$, for all $m\in\z_{\geq 1}$. For that purpose, we consider
$$\delta_T^{(n)}:=\delta_{T, K^{(n)}/k},\quad \Theta_{S, T}^{(n)}:=\Theta_{S, T, K^{(n)}/k}=\delta_{T, K^{(n)}/k}\cdot\Theta_{S, K^{(n)}/k} ,$$
viewed as holomorphic functions $\C\to\C[G_n]$, for all $n\in\z_{\geq 0}$. Note that we have
$$\delta_T^{(\infty)}(1-m):=(\delta_T^{(n)}(1-m))_n\in \underset{n}{\underset{\longleftarrow}\lim}\, \zp[G_n]=\zp[[\cg]],$$
for all $m\in\z_{\geq 1}$, as an immediate consequence of the definition of the $\delta_T^{(n)}$'s. Consequently, since $T\cap S_p=\emptyset$,
Remark \ref{theta-st-remark} implies that
$$(\Theta_{S, T}^{(n)}(1-m))_n\in \underset{n}{\underset{\longleftarrow}\lim}\, \zp[G_n]=\zp[[\cg]], \qquad \forall\, m\in\z_{\geq 1}.$$
We let $\Theta_{S, T}^{(\infty)}(1-m):=(\Theta_{S, T}^{(n)}(1-m))_n$ and view these as elements in $\zp[[\cg]]$.
\begin{lemma}\label{twisting-theta-st} For all $m\in\z_{\geq 1}$, we have the following equalities in $\zp[[\cg]]$.
\begin{enumerate}
\item $\delta_T^{(\infty)}(1-m)=t_{1-m}(\delta_T^{(\infty)}(0))$.
\item $\Theta_{S, T}^{(\infty)}(1-m)=t_{1-m}(\delta_T^{(\infty)}(0))\cdot t_{1-m}(g_S)=t_{1-m}(\Theta_{S,T}^{(\infty)}(0)).$
\end{enumerate}
\end{lemma}
\begin{proof} Let $m$ be as above. From the definitions, we have
$$\delta_T^{(\infty)}(1-m)=\prod_{v\in T}(1-(\sigma_v^{(\infty)})^{-1}\cdot{{\bf N}v}^{m}),$$
where $\sigma_v^{(\infty)}$ is the Frobenius morphism associated to $v$ in $\cg$. Now, since $T\cap S_p=\emptyset$, we have
$c(\sigma_v^{(\infty)})={\bf N}v$, for all $v\in T$. Combined with the last displayed equality, this implies part (1) of the Lemma.
Part (2) is a direct consequence of part (1) combined with Lemma \ref{twisting-gs}(2).
\end{proof}
\begin{corollary}
We have the following.
\begin{enumerate}
\item $g_S\in \vu_{\zp}^{-1}\zp[[\cg]]$, where $\vu_{\zp}:=\vu\cap\zp[[\cg]]$.
\item $G_S\in 1/|G|\zp[[\cg]]$.
\end{enumerate}
\end{corollary}
\begin{proof} Part (1) follows from Lemma \ref{twisting-theta-st}(2) with $m=1$ and the obvious observation that $\delta_T^{(\infty)}(0)\in\vu_{\zp}$.
Part (2) is an immediate consequence of part (1) and the fact that $H_S\in 1/|G|\zp[[\cg]]$.\end{proof}
\begin{definition} The $S$--incomplete, respectively $S$--incomplete $T$--modified, $G$--equivariant $p$--adic $L$--functions $\Theta_S^{(\infty)}\in\vu_{\zp}^{-1}\zp[[\cg]]$ and $\Theta_{S,T}^{(\infty)}\in\zp[[\cg]]$
associated to $(\ck/k, S, T, p)$ are defined by
$$\Theta_S^{(\infty)}:=\Theta_S^{(\infty)}(0)=g_S, \qquad \Theta_{S,T}^{(\infty)}:=\Theta_{S,T}^{(\infty)}(0)\,.$$
For consistency, we let $\delta_T^{(\infty)}:=(\delta_T^{(n)}(0))_n\in\zp[[\cg]]$.
\end{definition}
\begin{remark} Note that, by Lemmas \ref{twisting-gs} and \ref{twisting-theta-st}, we have
$$\Theta_{S,T}^{(\infty)}=\delta_T^{(\infty)}\cdot\Theta_S^{(\infty)}$$
for all $m\in\z_{\geq 1}$. Obviously, the link between $\Theta_{S, T}^{(\infty)}$ and the classical $p$--adic $L$--functions $L_p(\psi, s)$ defined in \eqref{p-adicL} is given by
\begin{equation}\label{link-to-classical-Lp}\chi(\Theta_{S,T}^{(\infty)})=\chi(\delta_T^{(\infty)})\cdot\frac{G_{\chi^{-1}\omega, S}(u(1+t)^{-1}-1)}{H_{\chi^{-1}\omega, S}(u(1+t)^{-1}-1)}\,,
\end{equation}
for all characters $\chi\in\widehat G(\cp)$.
Also, note that since for all odd characters $\psi\in\widehat G_n(\cp)$ we have $G_{\psi, S}=0$ (see Remark \ref{odd-characters}) and $p$ is odd,
we have
$$\Theta_S^{(\infty)}\in \frac{1}{2}(1-j)\cdot\vu_{\zp}^{-1}\zp[[\cg]], \qquad \Theta_{S,T}^{(\infty)}\in\frac{1}{2}(1-j)\cdot\zp[[\cg]].$$
In what follows, we view $\Theta_{S,T}^{(\infty)}$ and $\Theta_S^{(\infty)}$ as elements of both $1/2(1-j)\zp[[\cg]]$ and $\zp[[\cg]]^-=\zp[[\cg]]/(1+j)$
via the ring isomorphism $1/2(1-j)\zp[[\cg]]\simeq \zp[[\cg]]^-$ given by reduction modulo $(1+j)$.
\end{remark}
\begin{remark}\label{no-mup}
In this section we assumed that $\bmu_p\subseteq K$ (i.e. $\bmu_{p^\infty}\subseteq \ck$). This hypothesis was used whenever the twisting
automorphisms $t_m$ of $\co[[\cg]]$ with respect to the various powers $c^m$ of the $p$--cyclotomic character $c$ of $\cg$ were needed.
\noindent However, even if $\bmu_p\not\subseteq K$, we can still construct any object $\ast$ in the list $G_S(t)$, $H_S(t)$, $\Theta_S^{(n)}(s)$, $\delta_T^{(n)}(s)$, $\Theta_{S,T}^{(n)}(s)$ for the data $(\ck/k, S, T, p)$, just as above. Let $\widetilde K:=K(\mu_p)$, $\widetilde\ck=\ck(\mu_p)$, $\widetilde G:=G(\widetilde K/k)$, $\widetilde G_n:=G(\widetilde K_n/k)$, $\widetilde\cg:=G(\widetilde\ck/k)$. Further,
let $\widetilde\ast$ denote the analogue of $\ast$ for $(\widetilde\ck/k, S, T, p)$. By abuse of notation, let
$\pi:\C[\widetilde G_n]\twoheadrightarrow \C[G_n]$, $\pi: \co[[\widetilde\cg]]\twoheadrightarrow \co[[\cg]]$, etc., denote the $\co$--algebra morphisms induced by Galois restriction. The inflation property of Artin $L$--functions
immediately implies that
$$\pi(\widetilde\ast)=\ast\,,$$
for all $\ast$ as above. Consequently, we can construct equivariant $p$--adic $L$--functions $\Theta_S^{(\infty)}:=(\Theta_S^{(n)}(0))_n\in\vu_{\zp}^{-1}\zp[[\cg]]^-$ and $\Theta_{S,T}^{(\infty)}:=(\Theta_{S,T}^{(n)}(0))_n\in\zp[[\cg]]^-$, for $(\ck/k, S, T, p)$. Obviously, these will satisfy
$$\pi(\widetilde\Theta_{S, T}^{(\infty)})=\Theta_{S, T}^{(\infty)}.$$
\end{remark}
\subsection{\bf The Proof of the Equivariant Main Conjecture}\label{emc-section} In this section, we prove Theorem \ref{emc}. Consequently,
we work under the hypothesis $\mu_{\ck}=0$. In particular, we have $\mu_{\zp}(\cxs^+)=0$. (Recall from classical Iwasawa theory that $\mu_{\zp}(\cxs^+)=\mu_{K,p}^-$, for all
sets $S$ as above.)
\begin{lemma}
It suffices to prove Theorem \ref{emc} under the assumption that $\bmu_p\subseteq\ck$.
\end{lemma}
\begin{proof} Assume that $\bmu_p\not\subseteq\ck$. With notations as in remark \label{no-mup} above, let $\Delta:=G(\widetilde \ck/\ck)$.
Also let $\cm:=\mk$ and $\widetilde{\cm}:=\cm^{\widetilde\ck}_{\widetilde\cs, \widetilde\ct}$, where $\widetilde\cs$ and $\widetilde\ct$ are the sets of
primes in $\widetilde\ck$ sitting above primes in $\cs$ and $\ct$, respectively. Since
$$\ker(\pi: \zp[[\widetilde\cg]]^-\twoheadrightarrow \zp[[\cg]]^-)=I_{\Delta}\zp[[\widetilde\cg]]^-,$$
where $I_{\Delta}$ is the augmentation ideal in $\zp[\Delta]$, Corollary \ref{coinvariants} implies that we have isomorphisms
of $\zp[[\cg]]^-$--modules
$$T_p(\cm)^-\simeq T_p(\widetilde\cm)^-/I_{\Delta} T_p(\widetilde\cm)^-\simeq T_p(\widetilde\cm)^-\otimes_{\zp[[\widetilde\cg]]^-}\zp[[\cg]]^-.$$
Consequently, the second equality in \eqref{fitt-base-change} (Appendix) implies that we have
$${\rm Fit}_{\zp[[\cg]]^-}T_p(\cm)^-=\pi({\rm Fit}_{\zp[[\widetilde\cg]]^-}T_p(\widetilde\cm)^-).$$
Now the last equality in Remark \ref{no-mup} shows that Theorem \ref{emc} for the
data $(\widetilde\ck/k, S, T, p)$ implies the same result for the data $(\ck/k, S, T, p)$.
\end{proof}
\noindent As a consequence of the above Lemma, we may and will assume that $\bmu_p\subseteq\ck$ throughout the rest of this section. Now, note that Theorem \ref{projective}(1) shows that the $\zp[G]^-$--module
$T_p(\mk)^-$ is projective. Consequently, if we apply Proposition \ref{fitting-calculation}(1) (see Appendix) for the semi-local ring $R:=\zp[G]^-$, $g:=\gamma$ and the $R[[\Gamma]]$--module $M:=T_p(\mk)^-$, we conclude that Theorem \ref{emc} is equivalent to the following association in divisibility in $\zp[[\cg]]^-\simeq \zp[G]^-[[t]]$:
\begin{equation}\label{equivalent-association} \Theta_{S, T}^{(\infty)}\sim {\rm det}_{\zp[G]^-}((t+1)-\mathfrak m_\gamma\mid T_p(\mk)^-).\end{equation}
\begin{lemma}\label{G}For all odd characters $\chi\in\widehat G(\cp)$, we have the following in $\co[[t]]$.\begin{enumerate}
\item $G_{\chi^{-1}\omega, S}(u(1+t)^{-1}-1)\sim \chi\left({\rm det}_{Q(\co)[G]}((t+1)-\mathfrak m_\gamma\mid T_p(\cm^{\ck}_{\cs, \emptyset})^-\otimes_{\zp}Q(\co))\right).$
\item $\mu(G_{\chi^{-1}\omega, S}(u(1+t)^{-1}-1))=0$, with notations as in Corollary \ref{corollary-association}.
\end{enumerate}
\end{lemma}
\begin{proof} (1) Combine Corollary \ref{Wiles+zero-mu}(2)
and Lemma \ref{twist-poly-lemma}(3), applied to the lattice
$\cl:=\cxs^+\otimes_{\zp}\co$ and $n=1$ to conclude that
$$G_{\chi^{-1}\omega, S}(u(1+t)^{-1}-1)\sim \chi\left({\rm det}_{Q(\co)[G]}((t+1)-\mathfrak m_\gamma\mid (\cxs^+)^\ast(1)\otimes_{\zp}Q(\co))\right ).$$
Now use the $\zp[[\cg]]$--module isomorphism $T_p(\cm^{\ck}_{\cs, \emptyset})^-\simeq (\cxs^+)^\ast(1)$ given by Lemma \ref{link-classical} to conclude the proof of (1).
(2) This follows from (1) and the fact that the right-hand-side in (1) is a monic polynomial in $\co[t]$ (see Lemma \ref{twist-poly-lemma} in the Appendix.)
\end{proof}
\begin{lemma}\label{H} If $Q(\co)$ is endowed with the trivial $\cg$--action, then:
\begin{enumerate} \item $H_S(t)={\rm det}_{Q(\co)[G]}((t+1)-\mathfrak m_{\gamma}\mid Q(\co))$.
\item For all odd $\chi\in\widehat G(\cp)$, we have an association in divisibility in $\co[[t]]$:
$$H_{\chi^{-1}\omega, S}(u(1+t)^{-1}-1)\sim \chi({\rm det}_{Q(\co)[G]}((t+1)-\mathfrak m_{\gamma}\mid Q(\co)(1)))\,.$$
\item For all odd $\chi\in\widehat G(\cp)$, we have $\mu(H_{\chi^{-1}\omega, S}(u(1+t)^{-1}-1))=0$, with notations as in Corollary \ref{corollary-association}.
\end{enumerate}\end{lemma}
\begin{proof} Part (1) is clear from the equality $H_S(t):=t\cdot e_{\mathbf 1_G}+ (1-e_{{\mathbf 1}_G})$. Part (2) is a direct consequence of part (1),
and Lemma \ref{twist-poly-lemma}(3) applied to $\cl:=\co$ (with trivial $\cg$--action), $V=Q(\co)$, and $n=1$. Observe that $Q(\co)^\ast(1)\simeq Q(\co)(1)$.
Part (3) follows from (2) and the fact that the right-hand-side in (2) is a monic polynomial in $\co[t]$ (see Lemma \ref{twist-poly-lemma} in the Appendix.)
\end{proof}
\begin{lemma}\label{delta}
The $\zp[[\cg]]$--module $T_p(\Delta_{\ck, \ct})$ satisfies the following.
\begin{enumerate}
\item $T_p(\Delta_{\ck, \ct})\simeq\bigoplus_{v\in T}\zp[[\cg]]/(\delta_v^{(\infty})$, where $\delta_v^{(\infty)}:=(1-(\sigma_v^{(\infty})^{-1}\cdot{\bf N}v)$.
\item ${\rm Fit}_{\zp[[\cg]]} T_p(\Delta_{\ck, \ct})=(\delta_T^{(\infty})$.
\item ${\rm pd}_{\zp[[\cg]]}T_p(\Delta_{\ck, \ct})=1$ and ${\rm pd}_{\zp[G]}T_p(\Delta_{\ck, \ct})=0$.
\item $\delta_T^{(\infty)}\sim {\rm det}_{\zp[G]}((t+1)-\mathfrak m_\gamma\mid T_p(\Delta_{\ck, \ct}))$ in $\zp[[\cg]]$
\item $\mu(\chi(\delta_T^{(\infty)}))=0$, for all $\chi\in\widehat G(\cp)$.
\end{enumerate}
\end{lemma}
\begin{proof}
(1) For every $v\in T$, let us fix a prime $w(v)\in \ct$ sitting above $v$. Then, we have an obvious isomorphism of $\z[[\cg]]$--modules
$$\Delta_{\ck, \ct}\simeq \bigoplus_{v\in T}\left(\kappa(w(v))^\times\otimes_{\z[[\cg_v]]}\z[[\cg]]\right),$$
where $\cg_v$ is the decomposition group associated to $v$ in $\cg$ (topologically generated by
$\sigma_v^{(\infty})$) and $\kappa(w(v))$ denotes, as usual, the residue field associated to $w(v)$. This induces an isomorphism of $\zp[[\cg]]$--modules
$$T_p(\Delta_{\ck, \ct})\simeq \bigoplus_{v\in T}\left(T_p(\kappa(w(v))^\times)\otimes_{\zp[[\cg_v]]}\zp[[\cg]]\right)\,.$$
The isomorphism above combined with Lemma \ref{twisting-Fitting}(4) (see Appendix) applied to the extension $\kappa(w(v))/\kappa(v)$ of Galois group $\cg_v$,
for $q:={\rm card}(\kappa(v))$ and $\sigma_q=\sigma_v^{(\infty)}$ concludes the proof of (1).
(2) is a direct consequence of (1) and the definition of the Fitting ideal.
(3)
Note that if $\alpha\in\zp$, such that $\sigma_v^{(\infty)}=\gamma^\alpha$ and $\sigma_v$ is the Frobenius associated to $v$ in $G$,
then we have
$$\delta_v^{(\infty)}\sim (\sigma_v\cdot(t+1)^\alpha-{\bf N}v)\quad \text{in $\zp[[\cg]]\simeq\zp[G][[t]]$}.$$
Consequently,
$\chi(\delta_v^{(\infty)})=(\chi(\sigma_v)(1+t)^\alpha-{\bf N}v)$ is not a zero divisor in $\zp(\chi)[[t]]$, for all $\chi\in\widehat G(\cp)$ and therefore
$\delta_v^{(\infty)}$ is not a zero divisor in $\zp[[\cg]]$, for all $v\in T$. This observation combined with the isomorphism in part (1)
leads to the equality ${\rm pd}_{\zp[[\cg]]}T_p(\Delta_{\ck, \ct})=1$. Now, in order to obtain the equality ${\rm pd}_{\zp[G]}T_p(\Delta_{\ck, \ct})=0$, one applies Lemma \ref{projective-dimension}
for $M:=T_p(\Delta_{\ck, \ct})$ and $H:=G$, keeping in mind that $T_p(\Delta_{\ck, \ct})$ is $\zp$--free.
(4) This is a direct consequence of (2) and (3) combined with Proposition \ref{fitting-calculation}(1) applied for $R:=\zp[G]$, $g:=\gamma$ and the $R[[\Gamma]]=\zp[[\cg]]$--module
$M:=T_p(\Delta_{\ck, \ct})$.
(5) This is a direct consequence of (4) and the fact that the right-hand-side in (4) is a monic polynomial in $\zp[G][[t]]$ (see Remark \ref{monicity-remark} in the Appendix.)
\end{proof}
\begin{remark}\label{remark-delta} Note that the proofs of (1), (2) and the first equality in (3) of the above Lemma are independent on our
assumption that $\cg\simeq G\times \Gamma$.
\end{remark}
Now, we are ready to prove \eqref{equivalent-association} above. Let
$$\Theta:=\Theta_{S, T}^{(\infty)},\qquad F:= {\rm det}_{\zp[G]^-}((t+1)-\mathfrak m_\gamma\mid T_p(\mk)^-).$$
We view $\Theta$ and $F$ as power series in $\zp[G]^-[[t]]$ and plan on using Corollary \ref{corollary-association} (see Appendix) to show that $\Theta\sim F$. Note that, in fact,
$F$ is a monic polynomial in $\zp[G]^-[t]$ (see Remark \ref{monicity-remark} in the Appendix). Consequently, with notations as in Corollary \ref{corollary-association}, we have
\begin{equation}\label{mu-F}\mu(\chi(F))=0,\qquad \forall\, \chi\in\widehat G(\cp), \text{ $\chi$ odd.}\end{equation}
Now, Lemma \ref{G}(2), Lemma \ref{H}(3) and Lemma \ref{delta}(5) combined with equality \eqref{link-to-classical-Lp} above show that
\begin{equation}\label{mu-Theta}\mu(\chi(\Theta))=0,\qquad \forall\, \chi\in\widehat G(\cp), \text{ $\chi$ odd.}\end{equation}
Now, exact sequence \eqref{sequence-empty-T} combined with equality \eqref{link-to-classical-Lp} and Lemma \ref{G}(1), Lemma \ref{H}(2) and Lemma \ref{delta}(4) show that the following holds in $\co[[t]]$.
\begin{equation}\label{Theta-F}\chi(\Theta)\sim \chi\left( {\rm det}_{Q(\co)[G]^-}((t+1)-\mathfrak m_\gamma\mid T_p(\mk)^-\otimes_{\zp}Q(\co)\right)=\chi(F),\end{equation}
for all odd $\chi\in\widehat G(\cp)$. Now, as a consequence of \eqref{mu-F}, \eqref{mu-Theta} and \eqref{Theta-F}, Corollary \ref{corollary-association} in the Appendix leads to
$$\Theta\sim F \text{ in $\zp[G]^-[[t]]$,}$$
which concludes the proof of \eqref{equivalent-association} and that of Theorem \ref{emc}. \hfill $\Box$
\begin{corollary}\label{full-group-ring}
Under the hypotheses of Theorem \ref{emc}, we have
$${\rm Fit}_{\zp[[\cg]]}((T_p(\mk)^-)^\ast)={\rm Fit}_{\zp[[\cg]]}(T_p(\mk)^-)=\left(\Theta_{S,T}^{(\infty)},\quad \frac{1}{2}(1+j)\right)$$
where $(T_p(\mk)^-)^\ast:={\rm Hom}_{\zp}(T_p(\mk)^-, \zp)$ with the coinvariant
$\cg$--action.
\end{corollary}
\begin{proof}
Theorem \ref{projective}(1) permits us to apply Proposition \ref{fitting-calculation}(3) (see Appendix) to $M:=T_p(\mk)^-$. The latter implies the first equality above.
The second equality is a direct consequence of Theorem \ref{emc}. Note that in the statement of the Corollary $\Theta_{S, T}^{(\infty)}$ is viewed as an element of
$\zp[[\cg]]$.
\end{proof}
\section{Applications of the Equivariant Main Conjecture}\label{applications-section}
In this section, we give two application of Theorem \ref{emc} proved above: the first is a proof of a refinement of an imprimitive version of the Brumer--Stark Conjecture,
away from its $2$--primary part (see Theorem \ref{refined-BS} below); the second is a proof of a refinement of the Coates-Sinnott Conjecture, away from its $2$--primary part
(see Theorem \ref{cs-theorem} and Corollary \ref{cs-corollary} below.) Of course, these theorems are proved under the assumption that the relevant classical Iwasawa $\mu$--invariants vanish.
In what follows, if $M$ is a $\zp[[\mathcal H]]$--module, for some prime $p$ and some abelian profinite group $\mathcal H$,
then $M^\vee:={\rm Hom}_{\zp}(M, \qp/\zp)$ and $M^\ast:={\rm Hom}_{\zp}(M, \zp)$ denote its Pontrjagin and $\zp$--module dual, respectively. depending on the context, $M^\vee$ and $M^\ast$ will
be endowed either with the covariant $\mathcal H$--action $({}^\sigma f)(m):=f({}^\sigma m)$ or the contravariant one $({}^\sigma f)(m):=f({}^{\sigma^{-1}} m)$, for all $m\in M$, $\sigma\in\mathcal H$, and all $f\in M^\vee$ and $f\in M^\ast$, respectively.
Throughout this section, we let $K/k$ be an abelian extension of number fields of Galois group $G$ and let $S$ be a fixed finite set of primes in $k$, containing
the set $S_{\rm ram}(K/k)\cup S_\infty$. We will use the notations introduced at the beginning of \S\ref{EMC}.
\subsection{\bf The Brumer-Stark Conjecture} We fix a non-empty, finite set $T$ of primes in $k$, such that $S\cap T=\emptyset$. We assume that
$T$ contains either at least two primes of distinct residual characteristics or a prime of residual characteristic coprime to $w_K:=|\mu_K|$, where $\mu_K$ denotes the group of roots of unity in $K$.
(It is easily seen that this last condition is equivalent to the non-existence of non trivial roots of unity in $K$ which are congruent to $1$ modulo all primes in $T$.)
For simplicity, we let $\Theta_{S,T}:=\Theta_{S, T, K/k}$ denote
the $S$--incomplete, $T$--modified $G$--equivariant $L$--function associated to $(K/k, S, T)$ at the beginning of \S\ref{EMC}. Note that, under
our assumptions, Remark \ref{theta-st-remark} for $m=1$ gives
$$\Theta_{S, T}(0)\in\z[G].$$
In order to simplify notations, we let $S$ and $T$ also denote the sets of primes in $K$ sitting above primes in the original sets $S$ and $T$, respectively. Similarly, depending on the context, $S_\infty$ will denote the set of infinite primes in either $K$ or $k$. As in \S4.3 of \cite{Popescu-PCMI}, to the set of data $(K,T)$, one can associate a $T$--modified Arakelov class-group (Chow group) ${\rm CH}^1(K)^0_T$, endowed with a natural $\z[G]$--module structure, as follows. First, one defines the divisor group
$${\rm Div}(K)_T:=(\bigoplus_{w\not\in S_\infty\cup T}\z\cdot w)\bigoplus(\bigoplus_{w\in S_{\infty}}\Bbb R \cdot w)\,,$$
where the direct sums are taken over all the primes $w$ in $K$. Then, one defines
$${\rm deg}_{K, T}: {\rm Div}(K)_T\to \Bbb R$$
to be the unique degree map which is $\z$--linear on the left direct summand and $\Bbb R$--linear on the right and satisfies
$${\rm deg}_{K, T}(w):=\left\{
\begin{array}{ll}
\log|{\bf N}w|, & \hbox{if $w$ is a finite prime;} \\
1 & \hbox{if $w$ is an infinite prime.}
\end{array}
\right.$$
We let ${\rm Div}(K)^0_T:=\ker({\rm deg}_{K, T})$. Next, one defines a divisor map
$${\rm div}_K: K_T^\times\to {\rm Div}(K)^0_T, \quad {\rm div}_K(x):=\sum_{w\not\in S_\infty\cup T}{\rm ord_w}(x)\cdot w +\sum_{w\in S_{\infty}}(-\log|x|_w)\cdot w,$$
where ${\rm ord}_w$ and $|\cdot|_w$ denote the canonically normalized valuation and metric associated to $w$, respectively and $K_T^\times$ denotes the subgroup of $K^\times$ consisting of elements congruent to $1$ modulo all primes in $T$. Finally, one defines
$${\rm CH}^1(K)_T^0:=\frac{{\rm Div}(K)^0_T}{{\rm div}_K(K_T^\times)}\,.$$
If endowed with the natural quotient topology, this is a compact group whose volume is the absolute value of the leading term $\zeta_{S, T, k}^\ast(0)$
at $s=0$ of the modified zeta function $\zeta_{S,T,k}:=\Theta_{S, T, k/k}: \C\to \C$ associated to $k$ (see loc.cit.)
\begin{remark} Note the difference, both in notation and definition, between the finite divisor group ${\mathcal Div}_{K,T}$ and divisor map ${div}_{K}$ defined
in \S\ref{ideal-class-groups} and, respectively, the Arakelov divisor group ${\rm Div}(K)_T$ and divisor map ${\rm div}_{K}$ defined above. The link between the finite $T$--modified class-group
$C_{K,T}$ defined in \S\ref{ideal-class-groups} and the compact $T$--modified Arakelov class--group ${\rm CH}^1(K)^0_T$ defined above is captured by the following obvious exact sequence of $\z[G]$--modules.
$$\xymatrix{ 0\ar[r] &\dfrac{{\rm Div}(K)^0(S_\infty)}{{\rm div}_{K}(U_{K,T})}\ar[r] & {\rm CH}^1(K)^0_T\ar[r] & C_{K, T}\ar[r] &0.
}$$
Above, ${\rm Div}(K)^0(S_\infty):=\bigoplus_{v\in S_\infty}\Bbb R\cdot v$ denotes the $\Bbb R$--vector space of Arakelov divisors of degree $0$ supported on $S_\infty$ (denoted by $\Bbb RX_{S_\infty}$ in \cite{Popescu-PCMI})
and $U_{K,T}$ denotes the group of units in $K$ which are congruent to $1$ modulo all the primes in $T$.
(See \S4.3 of \cite{Popescu-PCMI} for the exact sequence above.)
\end{remark}
In \S4.3 of \cite{Popescu-PCMI}, we proved that the Brumer--Stark Conjecture ${\bf BrSt}(K/k, S)$ for the data $(K/k, S)$ (as stated, for example, in Chpt. IV, \S6 of \cite{Tate-Stark})
is equivalent to the following statement.
\begin{conjecture}[Brumer-Stark]\label{BS} Let $(K/k, S)$ be as above. Then, for all sets $T$ satisfying the above conditions, we have
\begin{equation*}{\bf BrSt}(K/k, S, T):\qquad \Theta_{S, T}(0)\in{\rm Ann}_{\z[G]} {\rm CH}^1(K)^0_T\,.\end{equation*}
\end{conjecture}
\medskip
\begin{remark}\label{BS-reduction-remark} Let us fix $(K/k, S, T)$ as above. It is easily seen that if $S$ contains at least two primes which split completely in $K/k$, then $\Theta_{S,T}(0)=0$.
If $S$ contains exactly one prime which splits completely in $K/k$, then $\Theta_{S,T}(0)=0$ unless $k$ is
an imaginary quadratic field, $K/k$ is unramified everywhere and $S=S_\infty$. In the latter case, the proof of Conjecture ${\bf BrSt}(K/k, S, T)$ is an easy, albeit instructive exercise: Indeed, according
to Corollary 4.3.3 (top change) in \cite{Popescu-PCMI}, one can assume that $K$ is the Hilbert class--field of $k$. In that case, it is easily seen that
$$\Theta_{S,T}(0)=\dfrac{\prod_{v\in T}(1-{\bf N}v)}{w_k}\cdot {{\bf N}_G},$$
where $v$ runs through primes in $k$, ${\bf N}_G:=\sum_{\sigma\in G}\sigma$ is the usual norm element in $\z[G]$, and $w_k$ is the cardinality of the group of roots of unity $\bmu_k$ in $k$. Now, the desired
annihilation result follows from the fact that ideals in $k$ become principal in $K$ (Hilbert's capitulation theorem) and the exact sequence of abelian groups
$$\xymatrix{1\ar[r] &\bmu_k\ar[r] &\Delta_{k, T}\ar[r] &C_{k, T}\ar[r] &C_k\ar[r] &1.
}$$
(See exact sequence \eqref{t-sequence} and note that $U_k=\bmu_k$ in this case.)
We leave the remaining details to the interested reader.
As a consequence, it suffices to study the Brumer-Stark Conjecture in the case where {\bf $k$ is totally real and $K$ is totally imaginary.} (Otherwise,
$S$ will contain at least one infinite prime which splits completely in $K/k$ and the above considerations settle the conjecture in that case.)
\end{remark}
Now, Proposition 4.3.7 in \cite{Popescu-PCMI} provides us with the following additional reduction.
\begin{proposition}\label{BS-reduction-CM} Let $(K/k, S, T)$ be as above. Assume that $k$ is totally real and $K$ is totally imaginary. Let $K^{CM}$ be the maximal
CM subfield of $K$, let $G_{CM}:=G(K^{CM}/k)$ and $\widetilde \Theta_{S, T}:=\Theta_{S, T, \widetilde K/k}.$ Then, the following
are equivalent, for all primes $p>2$.
\begin{enumerate}
\item $\Theta_{S, T}(0)\in{\rm Ann}_{\zp[G]} ({\rm CH}^1(K)^0_T\otimes\zp)$.
\item $\widetilde\Theta_{S, T}(0)\in{\rm Ann}_{\zp[G_{CM}]} (C_{K^{CM}, T}\otimes\zp)^-$.
\end{enumerate}
\end{proposition}
\begin{proof} See Proposition 4.3.7 in \cite{Popescu-PCMI}. The upper script ``${}^-$'' in (2) refers to the action of the unique complex conjugation
morphism in $G^{CM}$ on $(C_{K^{CM}, T}\otimes\zp)$.
\end{proof}
As a consequence of the Equivariant Main Conjecture (Theorem \ref{emc}), we can prove the following refinement of statement (2) in the
Proposition above, under certain hypotheses (see below.)
In what follows, $S_p$ denotes the set of $p$--adic primes in $k$ and $\mu_{K,p}$
denotes the Iwasawa $\mu$--invariant associated to the number field $K$ and the prime $p$.
\begin{theorem}\label{refined-BS} Let $(K/k, S, T)$ be as above and let $p$ be an odd prime. Assume that $K$ is CM, $k$ is totally real, $S_p\subseteq S$ and
$\mu_{K,p}=0$. Then, we have
$$\label{theta-fitting} \overline{\bf BrSt}(K/k, S, T, p):\qquad \Theta_{S, T}(0)\in {\rm Fit}_{\zp[G]}\, ((A_{K,T}^-)^\vee),$$
where $A_{K,T}:=(C_{K,T}\otimes\zp)$ and
the dual is endowed with the covariant $G$--action.
\end{theorem}
\begin{proof} In what follows, all occurring $\zp$--module or Pontrjagin duals are endowed with the {\bf covariant} action by the appropriate groups.
We let $\ck$ be the cyclotomic $\zp$--extension of $K$ and $\cg:={\rm Gal}(\ck/k)$. As usual, we let $\cs$ and $\ct$ denote the sets of finite primes
in $\ck$ sitting above primes in $S$ and $T$, respectively. Note that the hypotheses of Theorem \ref{emc} are satisfied by the data $(\ck/k, S, T, p)$.
Consequently, Corollary \ref{full-group-ring} implies that
\begin{equation}\label{theta-fitting}\Theta_{S, T}^{(\infty)}\in {\rm Fit}_{\zp[[\cg]]}((T_p(\mk)^-)^\ast)\,.\end{equation}
Now, recall that, with notations as in \S\ref{ideal-class-groups} and under the current hypotheses,
the $\zp[[\cg]]$--module
$$\ca_{\ck, \ct}^-\simeq\, \underset{n}{\underrightarrow\lim}\, A_{K_n, T_n}^-$$
is a torsion, divisible $\zp$--module of finite co-rank and that the transition maps in the injective limit above
are injective (see Lemma \ref{no-capitulation}.) Consequently, since $A_{K,T}^-$ is finite, we can fix $n\in\Bbb N$, such that we have an inclusion of $\zp[[\cg]]$--modules
$$A_{K,T}^- \subseteq \ca_{\ck, \ct}^-[p^n].$$
The inclusion above induces a natural surjection of $\zp[[\cg]]$--modules
\begin{equation}\label{surjection}\ca_{\ck, \ct}^-[p^n]^{\,\vee}\twoheadrightarrow (A_{K,T}^-)^\vee. \end{equation}
Next, we need the following elementary result.
\begin{lemma}\label{Pontrjakin-zp-dual} Let $p$ be a prime, $\mathcal H$ a profinite abelian group, and $M$ a $\zp[[\mathcal H]]$--module. Assume that $M$ is $\zp$--torsion, divisible, of finite
corank. Then there exist canonical isomorphisms of $\zp[[\mathcal H]]$--modules
$$M[p^m]^{\,\vee} \simeq T_p(M)^\ast\otimes_{\zp}\z/p^m\z\,, \qquad\text{ for all }m\in\z_{\geq 1}.$$
\end{lemma}
\begin{proof} Fix $m\in\z_{\geq 1}$. There is a canonical isomorphism of $\zp[[\mathcal H]]$-modules
$$T_p(M)^\ast\simeq M^\vee.$$ This induces a
canonical isomorphism $T_p(M)^\ast/p^m\simeq M^\vee/p^m$. However, the exact functor $\ast\to {\rm Hom}_{\zp}(\ast, \qp/\zp)$
applied to the exact sequence
$$\xymatrix {0\ar[r] &M[p^m]\ar[r] &M\ar[r]^{\times p^m} &M\ar[r] &0}$$
induces an isomorphism $M^\vee/p^m\simeq M[p^m]^\vee$. This concludes the proof.
\end{proof}
\noindent
Consequently, we obtain the following surjective morphisms of $\zp[[\cg]]$--modules.
\begin{equation}\label{surjections}(T_p(\mk)^-)^\ast\twoheadrightarrow T_p(\ca_{\ck, \ct}^-)^\ast\twoheadrightarrow \ca_{\ck, \ct}^-[p^n]^{\,\vee}\twoheadrightarrow (A_{K,T}^-)^\vee.\end{equation}
The first surjection above is obtained by taking $\zp$--duals in exact sequence \eqref{motive-exact-sequence}.
Note that $T_p(\ca_{\ck, \ct}^-)=T_p(\ca_{\ck, \ct})^-$. Also, note that the exact sequence \eqref{motive-exact-sequence} is split in the category of $\zp$--modules (because ${\mathcal Div}_{\ck}(\cs\setminus\cs_p)^-\otimes\zp$ is $\zp$--free) and therefore it stays exact after taking $\zp$--duals. The second surjection is given by Lemma \ref{Pontrjakin-zp-dual} applied to $M:=\ca_{\ck, \ct}^-$, $m:=n$, $\mathcal H:=\cg$. Finally, the third surjection is
\eqref{surjection} above.
Now, apply the first part of \eqref{fitt-base-change} in the Appendix for $R:=\zp[[\cg]]$, $M:=(T_p(\mk)^-)^\ast$ and $M':=(A_{K,T}^-)^\vee$, in combination with the surjections \eqref{surjections}
as well as \eqref{theta-fitting} above to conclude that we have
$$\Theta_{S,T}^{(\infty)}\in {\rm Fit}_{\zp[[\cg]]}T_p(\mk)^- = {\rm Fit}_{\zp[[\cg]]}(T_p(\mk)^-)^\ast \subseteq {\rm Fit}_{\zp[[\cg]]}\, ((A_{K,T}^-)^\vee).$$
Now, consider the projection $\pi:\zp[[\cg]]\twoheadrightarrow\zp[G]$ given by Galois restriction. By the definition of $\Theta_{S, T}^{(\infty)}$, we have
$$\pi(\Theta_{S,T}^{(\infty)})=\Theta_{S,T}(0),\qquad A_{K,T}^-\simeq A_{K,T}^-\otimes_{\zp[[\cg]]}\zp[G],$$
where the isomorphism is viewed in the category of $\zp[G]$--modules. Consequently, \eqref{fitt-base-change} (see Appendix)
applied for $M:=A_{K,T}^-$ and $\rho:=\pi$ implies that
$$\Theta_{S,T}(0)\in {\rm Fit}_{\zp[G]}\, ((A_{K,T}^-)^\vee),$$
which concludes the proof of the Theorem.
\end{proof}
\begin{corollary}
Assume that $(K/k, S, T)$ satisfy the hypotheses in Conjecture \ref{BS}. Let $p$ be a prime. If $k$ is totally real and $K$ is totally imaginary, assume that
$p$ is odd, $S_p\subseteq S$ and $\mu_{K^{CM},p}=0$. Then, we have:
$${\bf BrSt}(K/k, S, T, p):\qquad \Theta_{S,T}(0)\in{\rm Ann}_{\zp[G]}({\rm CH}^1(K)_T^0\otimes\zp).$$
\end{corollary}
\begin{proof} Combine Theorem \ref{refined-BS} with Proposition \ref{BS-reduction-CM} and Remark \ref{BS-reduction-remark} and note that
$${\rm Fit}_{\zp[G]}(M^\vee)\subseteq {\rm Ann}_{\zp[G]}(M),$$
for any $\zp[G]$--module $M$, if $M^\vee$ is endowed with the covariant $G$--action.
\end{proof}
\begin{remark} Some cases of the Brumer-Stark Conjecture over an arbitrary totally real number field were also settled by the first
author in \cite{Greither-Fitting} with different methods and working under somewhat more restrictive hypotheses. See Theorem 10 in loc.cit.
\end{remark}
\subsection{The Coates-Sinnott Conjecture.} As usual, $K/k$ is an abelian extension of number fields of
Galois group $G$, and $S$ is a finite set of primes in $k$, such that $S_{\rm ram}(K/k)\cup S_\infty\subseteq S$.
If $p$ is a prime number, we let ${\rm H}^i_{et}(\co_{K, S}[1/p], \zp(n))$ denote the $i$--th \'etale cohomology
group of the affine scheme ${\rm Spec}(\co_{K,S}[1/p])={\rm Spec}(\co_{K,S\cup S_p})$ with coefficients in the \'etale $p$--adic sheaf $\zp(n)$, for all $i\in\z_{\geq 0}$ and
all $n\in\z_{\geq 2}$. Also, for every $m\in\z_{\geq 0}$, we let ${\rm K}_m(\co_{K,S})$ denote the $m$--th Quillen ${\rm K}$--group of $\co_{K,S}$.
For the definitions and properties of these \'etale cohomology groups and ${\rm K}$--groups, the reader may consult Kolster's excellent survey article \cite{Kolster}.
For a discussion closer in spirit to the current section, the reader may consult \cite{Popescu-CS} as well. In the present context, all these \'etale cohomology and ${\rm K}$--theory groups come endowed with natural
$\zp[G]$--module, respectively $\z[G]$--module structures.
\noindent Quillen showed in \cite{Quillen} that ${\rm K}_m(\co_{K,S})$ is a finitely generated abelian group, for all $m$. Borel showed in \cite{Borel} that we have the following remarkable equalities
\begin{equation}\label{Borel}{\rm rank}_{\z}\, {\rm K}_{2n-i}(\co_{K,S})=\left\{
\begin{array}{ll}
0, & \hbox{if $i=2$;} \\
{\rm ord}_{s=(1-n)}\, \zeta_{K, S}(s), & \hbox{if $i=1$,}
\end{array}
\right.\end{equation}
for all $n\geq 2$, where $\zeta_{K,S}$ is the $S$--incomplete zeta--function associated to $K$. For all primes $p>2$, Soul\'e \cite{Soule} and later Dwyer-Friedlander \cite{Dwyer-Friedlander}
constructed surjective
$\zp[G]$--linear $p$--adic Chern character morphisms
\begin{equation}\label{chern}
{\rm ch}^i_{p,n}: {\rm K}_{2n-i}(\co_{K,S})\otimes\zp\twoheadrightarrow {\rm H}^i_{et}(\co_{K, S}[1/p], \zp(n)), \quad \forall\, i=1,2,\quad \forall\, n\in\z_{\geq 2}.
\end{equation}
Soul\'e proved in \cite{Soule-regulators} that these morphisms have finite kernels.
Similar morphisms have been defined for $p=2$ (see \cite{Dwyer-Friedlander}.) Their kernels and cokernels are finite but non trivial, in general.
\noindent Consequently, the group ${\rm H}^2_{et}(\co_{K, S}[1/p], \zp(n))$ is finite while ${\rm H}^1_{et}(\co_{K, S}[1/p], \zp(n))$ is finitely generated over $\zp$ of $\zp$--rank equal
to ${\rm ord}_{s=(1-n)}\, \zeta_{K, S}(s)$, for all
$p$ and all $n\geq 2$. If $i\geq 3$, the group ${\rm H}^i_{et}(\co_{K, S}[1/p], \zp(n))$
vanishes for $p> 2$ and is a finite, $2$-primary group for $p= 2$ (see \cite{Kolster}, \S2.)
With notations as in \S\ref{EMC}, we have the following.
\begin{lemma}\label{CS-lemma}For all $n\geq 2$ and all primes $p$, the following hold.
\begin{enumerate}
\item We have a $\zp[G]$--module isomorphism
$${\rm H}^1_{et}(\co_{K, S}[1/p], \zp(n))_{\rm tors}\simeq \left(\qp/\zp(n)\right)^{G_K}.$$
\item We have an equality of $\zp[G]$--ideals
$${\rm Ann}_{\zp[G]}({\rm H}^1_{et}(\co_{K, S}[1/p], \zp(n))_{\rm tors})=\langle\delta_{T, K/k}(1-n)\mid T \rangle,$$
where $T$ runs through all the finite, non-empty sets of primes in $k$ which are disjoint from $S\cap S_p$.
\item ${\rm Ann}_{\zp[G]}({\rm H}^1_{et}(\co_{K, S}[1/p], \zp(n))_{\rm tors})\cdot\Theta_{S, K/k}(1-n)\subseteq \zp[G].$
\end{enumerate}
\end{lemma}
\begin{proof} For (1), see \S2 in \cite{Kolster}. Note that in loc.cit., the author deals with the \'etale cohomology groups
of ${\rm Spec}(\co_K[1/p])$ rather than those of ${\rm Spec}(\co_{K,S}[1/p])$. However, for all $n$ and $p$ as above, Soul\'e's localization sequence in \'etale
cohomology (see \cite{Soule}) establishes an isomorphism of $\zp[G]$--modules
\begin{equation}\label{independent-h1}{\rm H}^1_{et}(\co_{K}[1/p], \zp(n))\simeq {\rm H}^1_{et}(\co_{K, S}[1/p], \zp(n)).\end{equation}
Part (2) follows from part (1) and a well--known lemma of Coates (see Lemma 2.3 in \cite{Coates} and its proof.) Also, see the proof
of Corollary \ref{Deligne-Ribet-rewrite} above.
Part (3) is a direct consequence of part (2) and Corollary \ref{Deligne-Ribet-rewrite} above.
\end{proof}
\begin{conjecture}[Coates-Sinnott, cohomological version]\label{CS-coh}
For all $(K/k, S, p\,, n)$ as above, the following holds.
\begin{eqnarray}{\bf CS}(K/k, S, p\,, n):\quad
\nonumber {\rm Ann}_{\zp[G]}({\rm H}^1_{et}(\co_{K, S}[1/p], \zp(n))_{\rm tors})\cdot\Theta_{S, K/k}(1-n)\subseteq \\
\nonumber \subseteq {\rm Ann}_{\zp[G]}({\rm H}^2_{et}(\co_{K,
S}[1/p], \zp(n))).
\end{eqnarray}
\end{conjecture}
\noindent With notations as above, let $e_n(K/k)$ be the idempotent in $\zp[G]$ given by
$$e_n(K/k):=\left\{
\begin{array}{ll}
\prod_{v\in S_\infty(k)}\frac{1}{2}(1+(-1)^n\sigma_v), & \hbox{if $k$ is totally real;} \\
0, & \hbox{otherwise,}
\end{array}
\right.$$
where $\sigma_v$ is the generator of the decomposition group $G_v$, for all $v\in S_\infty(k)$. The main goal of this
section is a proof of the following refinement of the conjecture above, under the expected hypotheses.
\begin{theorem}\label{cs-theorem} Assume that $(K/k, S, p\,, n)$ are as above. If $k$ is totally real, assume that $p>2$ and
$\mu_{K(\bmu_p)^{CM},p}=0$, where $K(\bmu_p)^{CM}$ is the maximal $CM$--subfield of $K(\bmu_p)$. Then, the following holds.
\begin{eqnarray}{\bf \overline{CS}}(K/k, S, p\,, n):\quad
\nonumber {\rm Ann}_{\zp[G]}({\rm H}^1_{et}(\co_{K, S}[1/p], \zp(n))_{\rm tors})\cdot\Theta_{S, K/k}(1-n)= \\
\nonumber =e_n(K/k)\cdot {\rm Fit}_{\zp[G]}({\rm H}^2_{et}(\co_{K,
S}[1/p], \zp(n))).
\end{eqnarray}
\end{theorem}
\begin{proof} We begin by making a few useful reduction steps.
\begin{lemma}\label{Sp} It suffices to prove statement $\overline{\bf CS}(K/k, S, p\,, n)$ under the assumption that $S_p\subseteq S$, where
$S_p$ denotes the set of $p$--adic primes in $k$.
\end{lemma}
\begin{proof} This is an immediate consequence of \eqref{independent-h1} and the obvious equality in $\zp[G]$
$$\Theta_{S\cup S_p, K/k}(1-n) = \Theta_{S,K/k}(1-n)\cdot \prod_{v\in S_p\setminus S}(1-\sigma_v^{-1}\cdot({\bf N}v)^{n}),$$
for all $n$ and $p$ as above. Indeed, for all $v\in S_p\setminus S$, the element $(1-\sigma_v^{-1}\cdot({\bf N}v)^{n})$ is a divisor of
$(1-p^m)$ in $\zp[G]$, for a large $m$ (say, $m$ such that $p^m=|\kappa(w)|^n$, for $w$ prime in $K$ dividing $v$) and it is therefore a unit in $\zp[G]$. Therefore,
the elements $\Theta_{S\cup S_p, K/k}(1-n)$ and $\Theta_{S,K/k}(1-n)$ differ by a unit in $\zp[G]$.
\end{proof}
\begin{lemma}\label{k-totally-real}
It suffices to prove statement $\overline{\bf CS}(K/k, S, p\,, n)$ under the assumption that $k$ is a totally real number field.
\end{lemma}
\begin{proof} Indeed, the functional equation satisfied by the $S$-incomplete $L$--function associated to a character $\chi\in\widehat G(\C)$ implies the following formula for the order of vanishing at $s=(1-n)$,
for all $n\in\z_{\geq 2}$.
\begin{equation}\label{orders-of-vanishing}
{\rm ord}_{s=(1-n)}L_S(\chi, s)=\left\{
\begin{array}{ll}
r_2(k)+a(\chi)^+, & \hbox{if $n$ is odd;} \\
r_2(k)+a(\chi)^-, & \hbox{if $n$ is even,}
\end{array}
\right.
\end{equation}
where $r_2(k)$ denotes the number of complex infinite places in $k$,
$$a(\chi)^+={\rm card}\{v\in S_\infty(k)\mid \chi\mid_{G_v}=\mathbf 1_{G_v}\}, \quad a(\chi)^-={\rm card}\{v\in S_\infty(k)\mid \chi\mid_{G_v}\ne\mathbf 1_{G_v}\}.$$
Note that $a(\chi)^++a(\chi)^-=r_1(k)$, the number of real
infinite places in $k$. The above formula shows that if $r_2(k)>0$ (i.e. if $k$ is not totally real), then we have $\Theta_{S, K/k}(1-n)=0$, for all $n\in\z_{\geq 2}$. This concludes the proof of the Lemma.
\end{proof}
\begin{lemma}\label{top-change} Assume that $\widetilde K/k$ is an abelian extension of number fields, of Galois group $\widetilde G$, with
$K\subseteq \widetilde K$ and $S_{\rm ram}(\widetilde K/k)\subseteq S$. Then, for all $p>2$ and $n$ as above,
$$\overline{\bf CS}(\widetilde K/k, S, p\,, n)\Rightarrow \overline{\bf CS}(K/k, S, p\,, n).$$
\end{lemma}
\begin{proof} Let us fix $p>2$ and $n$ as above. In order to simplify notations, we let ${\rm H}^i_K:={\rm H}^i_{et}(\co_{K, S}[1/p], \zp(n))$ and similarly
for ${\rm H}^i_{\widetilde K}$, for all $i=1,2$. Let $H:=G(\widetilde K/K)$. Galois restriction induces the usual $\zp$--algebra morphism
$\pi: \zp[\widetilde G]\twoheadrightarrow \zp[G]$, whose kernel is the relative augmentation ideal $I_H$ associated to $H$ in $\zp[\widetilde G]$. Proposition 2.10 in \cite{Kolster}
gives a canonical $\zp[G]$--module isomorphism
$${\rm H}^2_K\simeq ({\rm H}^2_{\widetilde K})_H,\qquad {\rm H}^1_K\simeq ({\rm H}^1_{\widetilde K})^H$$
where $M^H:=M[I_H]$ and $M_H:=M/I_HM\simeq M\otimes_{\zp[\widetilde G]}\zp[G]$ denote the modules of $H$--invariants and $H$--coinvariants, respectively,
associated to any $\zp[\widetilde G]$--module
$M$. The isomorphism above combined with the second equality in \eqref{fitt-base-change} of the
Appendix (applied for $\rho:=\pi$, $R:=\zp[\widetilde G]$, $R':=\zp[G]$, and $M:={\rm H}^2_K$ gives
\begin{equation}\label{h2-coinvariants}\pi({\rm Fit}_{\zp[\widetilde G]}({\rm H}^2_{\widetilde K}))={\rm Fit}_{\zp[G]}({\rm H}^2_{K}).
\end{equation}
The above equality combined with the obvious equalities $\pi(e_n(\widetilde K/k))=e_n(K/k)$, $\pi(\Theta_{S,\widetilde K/k}(1-n))=\Theta_{S,K/k}(1-n)$ and
\begin{equation}\label{h1-invariants} \pi({\rm Ann}_{\zp[\widetilde G]}( {\rm H}^1_{\widetilde K})_{\rm tors})={\rm Ann}_{\zp[G]}( {\rm H}^1_{K})_{\rm tors}
\end{equation}
concludes the proof of the Lemma. (Apply Lemma \ref{CS-lemma}(2) for the last equality.)
\end{proof}
\begin{lemma}\label{K-CM} With notations as in the previous Lemma, assume that $k$ is totally real, $\widetilde K$ is totally imaginary and
$K:={\widetilde K}^{\,CM}$ is the maximal CM subfield of $\widetilde K$. Then, for all $p>2$ and $n$ as above, we have
$$\overline{\bf CS}(\widetilde K/k, S, p\,, n) \Leftrightarrow \overline{\bf CS}(K/k, S, p\,, n).$$
\end{lemma}
\begin{proof} Let us fix $p>2$ and $n$ as above. We use the notations in the proof of the previous Lemma.
Note that $H$ is the $2$--group generated by the set
$$\{\sigma_{v}\cdot \sigma_{v'}\mid v, v'\in S_\infty(k), v\ne v'\}.$$
This shows that, for all $\chi\in\widehat G(\C)$, we have
$$\chi\mid_H\ne\mathbf 1_H\quad \Rightarrow \quad L_{S}(\chi, 1-n)=0, \quad \chi(e_n(\widetilde K/k))=0.$$
Indeed, if $\chi\mid_H\ne\mathbf 1_H$, then there exist $v, v'\in S_\infty(k)$, with $v\ne v'$, such that
$\chi(\sigma_v)=1$ and $\chi(\sigma_{v'})\ne 1$. On one hand, this shows that $\chi(e_n(\widetilde K/k))=0$. On the other hand, it shows that $a(\chi)^+\geq 1$ and $a(\chi)^-\geq 1$. Now, \eqref{orders-of-vanishing}
implies that $L_S(\chi, 1-n)=0$, for all $n\geq 2$. Consequently, if $e_H:=|H|^{-1}\cdot \sum_{h\in H}h$ is the idempotent element
associated to $H$ in $\zp[\widetilde G]$ (note that $p\nmid |H|$), we have
$$\Theta_{S,\widetilde K/k}(1-n)\in e_H\qp[\widetilde G], \qquad e_n(\widetilde K/k)\in e_H\zp[\widetilde G].$$
Now, note that since $\zp[\widetilde G]=e_H\zp[\widetilde G]\oplus (1-e_H)\zp[\widetilde G]$ and $I_H=(1-e_H)\zp[\widetilde G]$, the map $\pi$ establishes a ring isomorphism $\pi: e_H\zp[\widetilde G]\simeq \zp[G]$.
Consequently, \eqref{h2-coinvariants} and \eqref{h1-invariants} imply that $\pi$ establishes an isomorphism at the level of $e_H\zp[\widetilde G]$--ideals
$$\pi: e_H{\rm Fit}_{\zp[\widetilde G]}({\rm H}^2_{\widetilde K})\simeq {\rm Fit}_{\zp[G]}({\rm H}^2_{K}),\quad
\pi: e_H{\rm Ann}_{\zp[\widetilde G]}({\rm H}^1_{\widetilde K})_{\rm tors}\simeq {\rm Ann}_{\zp[G]}({\rm H}^1_{K})_{\rm tors} .$$
Now, the equalities $\pi(e_n(\widetilde K/k))=e_n(K/k)$ and $\pi(\Theta_{S,\widetilde K/k}(1-n))=\Theta_{S,K/k}(1-n)$ conclude the proof of the Lemma.
\end{proof}
Now, we are ready to return to the proof of Theorem \ref{cs-theorem}. We fix $(K/k,S, p)$ as in the theorem. According to the previous four Lemmas, we may assume that $k$ is totally real
(otherwise, the statement is trivially true, according to Lemma \ref{k-totally-real}), $S_p\subseteq S$ (see Lemma \ref{Sp}), $K$ is CM and
$\bmu_p\subseteq K$ (otherwise, we replace $K$ by $K(\bmu_p)^{CM}$ and apply Lemmas \ref{K-CM} and \ref{top-change}, respectively.) Under these hypotheses,
we assume in addition that $p>2$ and $\mu_{K,p}=0$.
As usual, we let $\ck$ denote the cyclotomic $\zp$--extension of $K$, $\cg:=G(\ck/k)$ and $\Gamma:=G(\ck/K)$. We fix a finite, nonempty set $T$ set of primes in $K$, such that $S\cap T=\emptyset.$ Also,
we let $\cs$ and $\ct$
denote the sets of finite primes in $\ck$ sitting above primes in $S$ and $T$, respectively. For simplicity, we let $\Theta_S:=\Theta_{S, K/k}$, $\Theta_{S,T}:=\Theta_{S,T,K/k}$ and $\delta_T:=\delta_{T,K/k}$
and resume using all the notations introduced in \S\ref{EMC}. In particular, $j\in\cg$ denotes the unique complex conjugation automorphism of $\ck$ (identified as usual with the complex
conjugation automorphism $j\in G$ of $K$.) Note that we have
$$e_n:=e_n(K/k)=\frac{1}{2}(1+(-1)^n j), \qquad \forall\, n\in\z.$$
We view $e_n$ either as an element of $\zp[[\cg]]$ or $\zp[G]$, for all $n\in\z$.
Observe that if $M$ is a $\zp[[\cg]]$--module and $n\in\z$,
then $e_nM=M^+$, if $n$ is even and $e_nM=M^-$, if $n$ is odd. Also, note that as a consequence of \eqref{orders-of-vanishing} we have
\begin{equation}\label{theta-component}\Theta_{S}(1-n)=e_n\cdot \Theta_{S}(1-n), \quad
\Theta_{S,T}(1-n)=e_n\cdot \Theta_{S,T}(1-n),\qquad \forall\, n\in\z_{\geq 2}.
\end{equation}
In what follows, all the occurring $\zp$--module duals or Pontrjagin duals are endowed with the {\bf contravariant}
actions by the appropriate groups, unless stated otherwise.
We will need the following elementary result.
\begin{lemma}\label{elementary-lemma} Let $M$ be a $\zp[[\cg]]$--module. Assume that $M$ is $\zp$--free of finite rank and that $M_{\Gamma}$ is finite. Then,
\begin{enumerate}\item $M^\Gamma=0$.
\item $(M_{\Gamma})^\vee\simeq (M^\ast)_{\Gamma},$ as $\zp[G]$--modules.
\item If ${\rm pd}_{\zp[[\cg]]}M\leq 1$, then ${\rm pd}_{\zp[G]}M_{\Gamma}\leq 1$.
\end{enumerate}
\end{lemma}
\begin{proof} Part (1) is immediate. For part (2), see Lemma 5.18 in \cite{GP}. Part (3) is an immediate consequence of
$\zp[[\cg]]^{\,\Gamma}=0$ and $\zp[[\cg]]_{\,\Gamma}\simeq\zp[G]$. \end{proof}
\begin{proposition}\label{dual-coinvariants}
Let $n\in\z_{\geq 2}$, $p>2$, and ${\rm H}^i_{K}:={\rm H}_{et}^i(\co_{K,S}, \zp(n))$, for all $i=1,2$. Then, we have $\zp[G]$--module isomorphisms
\begin{enumerate}
\item $e_n\cdot {\rm H}^2_K\simeq (\cxs^+(-n)_{\Gamma})^\vee\simeq (\cxs^+(-n)^\ast)_{\Gamma};$\medskip
\item $e_n\cdot {\rm H}^1_K=e_n\cdot ({\rm H}^1_K)_{\rm tors}\simeq({\rm H}^1_K)_{\rm tors}\simeq\zp(n)_\Gamma,$
\end{enumerate}
where $M^\Gamma$ and $M_{\Gamma}$ denote the $\zp[G]$--modules of $\Gamma$--invariants and $\Gamma$--coinvariants,
respectively, for all $\zp[[\cg]]$--modules $M$.
\end{proposition}
\begin{proof} Let ${\rm H}^i_{K^+}:={\rm H}_{et}^i(\co_{K^+,S}, \zp(n))$, for all $i=1,2$, where $K^+:=K^{j=1}$, as usual. Then, Proposition 2.9 in \cite{Kolster} shows that there is a canonical $\zp[G]$--module
isomorphism ${\rm H}^1_{K^+}\simeq ({\rm H}^1_K)^+$. Consequently, \eqref{chern} and \eqref{Borel} together with \eqref{orders-of-vanishing} give
$${\rm rank}_{\zp}\, (e_n\cdot {\rm H}^1_K)=\left\{
\begin{array}{ll}
r_2(K)-r_1(K^+) , & \hbox{if $n$ is odd;}\\
r_2(K^+) , & \hbox{if $n$ is even}
\end{array}
\right\} = 0.
$$ Consequently, $e_n{\rm H}^1_K$ is a finite group. On one hand, this implies the equality in part (2) of the Proposition. On the other hand, via the spectral sequence argument on pp. 237--238 of \cite{Kolster}
this leads to isomorphisms of $\zp[G]$--modules
$$e_n\cdot {\rm H}^2_{K}\simeq e_n\cdot{\rm H}^1_{et}(\co_{K,S}, \qp/\zp(n))\simeq e_n\cdot {\rm H}^1(G_{\ck, \cs}^{(p)}, \qp/\zp(n))^{\Gamma},$$
where $G_{\ck, \cs}^{(p)}$ is the Galois group of the maximal pro--$p$ extension of $\ck$ which is unramified away from $\cs$ and the right--most cohomology group
is a Galois cohomology group. Now, since $\bmu_{p^\infty}\subseteq\ck$ and therefore $G_{\ck, \cs}^{(p)}$ acts trivially on $\qp/\zp(n)$ and since
$\cxs$ is the maximal abelian quotient of $G_{\ck, \cs}^{(p)}$, we have the following isomorphisms of $\zp[G]$--modules.
\begin{eqnarray}
\nonumber {\rm H}^1(G_{\ck, \cs}^{(p)}, \qp/\zp(n))^{\Gamma}\simeq & {\rm Hom}_{\zp}(G_{\ck, \cs}^{(p)}, \qp/\zp(n))^{\Gamma} & \\
\nonumber \simeq & {\rm Hom}_{\zp}(\cxs, \qp/\zp(n))^{\Gamma} &\simeq\quad (\cxs(-n)_{\Gamma})^\vee.
\end{eqnarray}
Now, the first isomorphism in part (1) of the Proposition follows from the last two displayed isomorphisms and the obvious equality $e_n\cdot\cxs(-n)=\cxs^+(-n)$.
The second isomorphism in part (1) follows from Lemma \ref{elementary-lemma}(2) applied to $M:=\cxs^+(-n).$
Note that the finiteness of ${\rm H}^2_K$ and the first isomorphism in part (1) of the Proposition imply that $\cxs^+(-n)_{\Gamma}$ is finite. Also, $\cxs^+(-n)$ is
finitely generated over $\zp$ (a theorem of Iwasawa) and therefore $\zp$--free of finite rank (a consequence of $\mu_{K,p}=0$.)
This concludes the proof of part (1) of the Proposition.
The isomorphism in part (2) of the Proposition is an immediate consequence of the isomorphism $({\rm H}^1_K)_{\rm tors}\simeq(\qp/\zp(n))^\Gamma$ (see Lemma \ref{CS-lemma}(1) and recall that $\bmu_{p^\infty}\subseteq\ck$)
and the obvious equality $\qp(n)^\Gamma=\qp(n)_{\Gamma}=0$.
\end{proof}
Note that the hypotheses of Lemma \ref{link-classical} and Theorem \ref{emc} are satisfied by the data $(\ck/k, \cs, \ct, p)$.
Lemma \ref{link-classical} and exact sequence \eqref{sequence-empty-T} tensored with $\zp(n-1)$ lead to the following four term exact sequence of $\zp[[\cg]]$--modules.
\begin{equation}\label{four-term-sequence} 0\to \zp(n)\to T_p(\Delta_{\ck, \ct})^-(n-1)
\to T_p(\mk)^-(n-1)\to \cxs^+(-n)^\ast
\to 0.\end{equation}
We intend to apply Proposition \ref{four-term-sequence-fitting} in the Appendix to the sequence of $\Gamma$--coinvariants associated to the exact sequence above.
First, let us note that each of the four modules in the exact sequence above is $\zp$--free. Most importantly, the module of $\Gamma$--coinvariants associated to each of these is finite.
Indeed, $\zp(n)_{\Gamma}$ and $(\cxs^+(-n)^\ast)_{\Gamma}$ are finite, due to Proposition \ref{dual-coinvariants}. On the other hand,
Lemma \ref{delta}(1) and Remark \ref{remark-delta} combined with Lemma \ref{twisting-Fitting}(1) in the Appendix imply that we have a $\zp[G]$--module isomorphism
$$T_p(\Delta_{\ck, \ct})(n-1)_{\Gamma}\simeq\bigoplus_{v\in T}\zp[G]/(1-\sigma_v^{-1}\cdot N_v^{n}).$$
Since $n\geq 2$, the element $(1-\sigma_v^{-1}\cdot N_v^{n})$ is not a zero--divisor in $\zp[G]$, for all $v\in T$. Consequently, the above isomorphism
implies that $T_p(\Delta_{\ck, \ct})(n-1)_{\Gamma}$ is finite. Therefore, its direct summand $T_p(\Delta_{\ck, \ct})^-(n-1)_{\Gamma}$ is indeed finite. Now, the finiteness
of $T_p(\mk)^-(n-1)_{\Gamma}$ follows from the exact sequence
$$T_p(\Delta_{\ck, \ct})^-(n-1)_{\Gamma}
\to T_p(\mk)^-(n-1)_{\Gamma}\to (\cxs^+(-n)^\ast)_{\Gamma}
\to 0.$$
Now, Lemma \ref{elementary-lemma} implies that the $\Gamma$--invariants of the four modules in \eqref{four-term-sequence} are trivial. Consequently, we obtain the following exact sequence of finite $\zp[G]$--modules
at the level of $\Gamma$--coinvariants
\begin{equation}\label{four-term-sequence-coninvariants} 0\to e_n\cdot{\rm H}^1_K\to T_p(\Delta_{\ck, \ct})^-(n-1)_{\Gamma}
\to T_p(\mk)^-(n-1)_{\Gamma}\to e_n\cdot {\rm H}^2_K
\to 0.\end{equation}
Above, we have used Proposition \ref{dual-coinvariants} to identify the $\Gamma$--coinvariants of the end terms of \eqref{four-term-sequence} with $e_n\cdot{\rm H}^1_K$ and
$e_n\cdot{\rm H}^2_K$, respectively.
Since ${\rm pd}_{\zp[[\cg]]}T_p(\Delta_{\ck, T})^-\leq 1$ and ${\rm pd}_{\zp[[\cg]]} T_p(\mk)^-\leq 1$ (see Lemma \ref{delta}(3) and Remark \ref{remark-delta} for the first and Theorem \ref{projective}(2) for the second),
we have
$${\rm pd}_{\zp[G]}\,(T_p(\Delta_{\ck, \ct})^-(n-1)_{\Gamma})\leq 1, \qquad {\rm pd}_{\zp[G]}\, (T_p(\mk)^-(n-1)_{\Gamma})\leq 1,$$
as a consequence of \ref{twisting-Fitting}(3) and Lemma \ref{elementary-lemma}(3).
Consequently, we may apply Proposition \ref{four-term-sequence-fitting} in the Appendix to the
exact sequence \eqref{four-term-sequence-coninvariants}.
This way, we obtain
\begin{eqnarray}\label{product-fitting}
& {\rm Fit}_{\zp[G]}({e_n{\rm H}^1_K}^\vee)\cdot {\rm Fit}_{\zp[G]}(T_p(\mk)^-(n-1)_{\Gamma})=\\
\nonumber & ={\rm Fit}_{\zp[G]}(e_n{\rm H}^2_K)\cdot {\rm Fit}_{\zp[G]}(T_p(\Delta_{\ck, \ct})^-(n-1)_{\Gamma}).
\end{eqnarray}
where the dual is endowed with the covariant $G$--action. Now, let us note that for any $\zp[[\cg]]$--module $M$ and $\zp[G]$--module $N$, we have
$$e_n\cdot M(n-1)=M^-(n-1), \qquad e_n\cdot{\rm Fit}_{\zp[G]}(N)={\rm Fit}_{e_n\zp[G]}(e_n\cdot N).$$
Consequently, if we combine Lemma \ref{twisting-Fitting}(2) in the Appendix with Corollary \ref{full-group-ring}
and Lemma \ref{delta}(2), respectively, we obtain
\begin{eqnarray}
\nonumber & e_n\cdot{\rm Fit}_{\zp[G]}(T_p(\mk)^-(n-1)_{\Gamma}) = e_n\cdot (\pi\circ t_{1-n}(\Theta_{S,T}^{(\infty)}))=(\Theta_{S,T}(1-n)),\\
\nonumber & e_n\cdot{\rm Fit}_{\zp[G]}(T_p(\Delta_{\ck, \ct})^-(n-1)_{\Gamma})= e_n\cdot (\pi\circ t_{1-n}(\delta_{T}^{(\infty)}))=e_n\cdot(\delta_T(1-n)),
\end{eqnarray}
where $\pi:\zp[[\cg]]\to\zp[G]$ is the usual projection. Note that above we have used the second equality in \eqref{theta-component}. Consequently, \eqref{product-fitting} implies that
$${\rm Fit}_{\zp[G]}({e_n{\rm H}^1_K}^\vee)\cdot\Theta_{S,T}(1-n)=e_n{\rm Fit}_{\zp[G]}({\rm H}^2_K)\cdot(\delta_T(1-n)).$$
However, since $\Theta_{S,T}(1-n)=\delta_T(1-n)\cdot\Theta_S(1-n)$ and $\delta_T(1-n)$ is not a zero--divisor in $\zp[G]$ (an easy exercise !), the last equality implies
$${\rm Fit}_{\zp[G]}(({\rm H}^1_K)_{\rm tors}^{\,\,\vee})\cdot\Theta_S(1-n)=e_n{\rm Fit}_{\zp[G]}({\rm H}^2_K).$$
Now, since $({\rm H}^1_K)_{\rm tors}$ is a cyclic module (see Lemma \ref{elementary-lemma}(2)) and the dual is endowed with the covariant
$G$--action, we have equalities
$${\rm Fit}_{\zp[G]}(({\rm H}^1_K)_{\rm tors}^{\,\,\vee})={\rm Ann}_{\zp[G]}(({\rm H}^1_K)_{\rm tors}^{\,\,\vee})={\rm Ann}_{\zp[G]}(({\rm H}^1_K)_{\rm tors}).$$
When combined with the last displayed equality, this concludes the proof of Theorem \ref{cs-theorem} (our refinement of the cohomological Coates-Sinnott Conjecture.)
\end{proof}
\begin{remark} Results somewhat weaker than our Theorem \ref{cs-theorem} were obtained with different methods in \cite{Burns-Greither}, Corollary 2 (which imposes restrictions upon $K/k$ and $p$) and
\cite{NQD}, Th\'eor\`eme 4.3 (which imposes restrictions upon $K/k$ and $n$.)
In both cases, the vanishing of the appropriate Iwasawa $\mu$--invariant is assumed.
\end{remark}
Finally, we would like to mention that the well known Quillen-Lichtenbaum Conjecture states that the Chern character maps ${\rm ch}^i_{p,n}$ (see \eqref{chern} above) are isomorphisms,
for all $p>2$ and $n\geq 2$. On the other hand, it is known that the Quillen-Lichtenbaum Conjecture is a consequence of the Bloch-Kato Conjecture for finitely generated fields (e.g., see Theorem 2.7 in \cite{Kolster}.)
To our knowledge, recent work of Rost and Voevodsky has lead to a proof of the Bloch-Kato Conjecture.
\noindent The following is an immediate consequence of Theorem \ref{cs-theorem}.
\begin{corollary}[a refined ${\rm K}$--theoretic Coates--Sinnott Conjecture] \label{cs-corollary} Let $K/k$ be an abelian extension of number fields of
Galois group $G$. Let $S$ be a finite set of primes in $k$, such that $S_\infty(k)\cup S_{\rm ram}(K/k)\subseteq S$. If $k$ is totally real, assume that
$\mu_{K(\bmu_p)^{CM},p}=0$, for all primes $p>2$. Also, assume that the Quillen-Lichtenbaum Conjecture holds. Then, for all $n\in\z_{\geq 2}$, we have the following equality
of $\z[1/2][G]$--ideals.
\begin{eqnarray}
\nonumber \z[1/2]\,{\rm Ann}_{\z[G]}({\rm K}_{2n-1}(\co_{K,S})_{\rm tors})\cdot\Theta_S(1-n)=\\
\nonumber =e_n(K/k)\cdot \z[1/2]\,{\rm Fit}_{\z[G]}({\rm K}_{2n-2}(\co_{K,S})).
\end{eqnarray}
\end{corollary}
\noindent As the reader will notice right away, the ${\rm K}$--theoretic statement above is closer in spirit to the conjecture originally formulated by Coates and Sinnott in \cite{Coates-Sinnott}.
The difference is the presence of a Fitting ideal rather than an annihilator on the right-hand side and an equality rather than an inclusion of ideals. These differences justify the use of the term ``refined'' above.
\section{Appendix: Algebraic Ingredients}
\subsection{Determinants and Fitting Ideals}\label{appendix-fitting}
Let $R$ be a commutative ring with $1$, $P$ a finitely generated, projective $R$--module and $f\in{\rm End}_R(P)$. Then,
the determinant $\det_R(f\mid P)$ of $f$ acting on $P$ is defined as follows. We take a finitely generated $R$--module $Q$,
such that $P\oplus Q$ is a (finitely generated) free $R$--module, then we let $f\oplus\mathbf 1_Q\in{\rm End}_R(P\oplus Q)$, where
$\mathbf 1_Q$ is the identity of $Q$, and define
$${\rm det}_R(f\mid P):={\rm det}_R(f\oplus \mathbf 1_Q\mid P\oplus Q)\,.$$
It is easy to check (use Schanuel's Lemma !) that the definition above does not depend on $Q$. Now, one can use the same strategy to define
the characteristic polynomial ${\rm det}_R(X-f\mid P)\in R[X]$ of variable $X$. Indeed, $P\otimes_R R[X]$ is a finitely
generated, projective $R[X]$--module. One defines
$${\rm det}_R(X-f\mid P):={\rm det}_{R[X]}({\rm id}_P\otimes X-f\otimes 1\mid P\otimes_R R[X])\,.$$
For any $P$, $R$ and $f$ as above and any $R$--algebra $R'$, we have base-change equalities
\begin{equation}\label{det-base-change}
\begin{array}{c}
{\rm det}_R(f\mid P)={\rm det}_{R'}(f\otimes\mathbf 1_{R'}\mid P\otimes_R R')\,, \\
{}\\
{\rm det}_R(X-f\mid P)={\rm det}_{R'}(X- (f\otimes\mathbf 1_{R'})\mid P\otimes_R R')\,.
\end{array}
\end{equation}
\begin{remark}\label{monicity-remark} It is easily checked that the polynomial $F(X):={\rm det}_R(X-f\mid P)$ in $R[X]$ defined above is always a monic polynomial.
\end{remark}
Now, for any $R$ as above and any finitely presented $R$--module $M$, the first Fitting invariant (ideal) ${\rm Fit}_R(M)$ of $M$ over $R$ is defined as follows.
First, one considers a finite presentation of $M$
$$\xymatrix{
R^n\ar[r]^\phi &R^m\ar[r] &M\ar[r] &0}.$$ By definition, the
Fitting ideal ${\rm Fit}_R(M)$ is the ideal in $R$ generated by the
determinants of all the $m\times m$ minors of the matrix $A_\phi$
associated to $\phi$ with respect to $R$--bases of $R^n$ and
$R^m$. It is well-known that the definition does not depend on the
chosen presentation or bases, and
\begin{equation}\label{ann-fitt}
{\rm Ann}_R(M)^m\subseteq{\rm Fit}_R(M)\subseteq {\rm Ann}_R(M)\,.
\end{equation}
Two elementary properties of Fitting ideals which are used throughout the paper state that
if $M\twoheadrightarrow M'$ is a surjective morphism of finitely presented $R$--modules and $\rho:R\to R'$ is a morphism of commutative rings with $1$,
then one has
\begin{equation}\label{fitt-base-change}{\rm Fit}_R(M)\subseteq {\rm Fit}_R(M'),\qquad {\rm Fit}_{R'}(M\otimes_R R')=\rho({\rm Fit}_R(M))R'.\end{equation}
For more details on general properties of Fitting ideals, the reader can consult the Appendix of \cite{Mazur-Wiles}.
In what follows, if $R$ is a commutative topological ring and $\Gamma$ is
a profinite group, then the profinite group algebra
$$R[[\Gamma]]:=\underset{\mathfrak H}{\underset\longleftarrow\lim} R[\Gamma/\mathfrak H],$$
where $\Gamma/\mathfrak H$ are all the finite quotients of
$\Gamma$ by (open and) closed subgroups $\mathfrak H$, is viewed as a
topological $R$--algebra endowed with the usual projective limit
topology.
Below, by a semi-local ring $R$ we mean a direct sum of finitely many local rings. Examples of such rings include $\co[G]$ and $\co[G]^\pm:=\co[G]/(1\mp j)$, where $G$ is a finite, abelian group,
$\co$ is a finite integral extension of $\zp$, for some odd prime $p$ and $j$ is an element of order $2$ in $G$.
\begin{proposition}\label{fitting-calculation}
Let $R$ be a commutative, semi-local, compact topological ring and $\Gamma$ a pro-cyclic group of topological generator $g$. Let $M$ be a topological $R[[\Gamma]]$--module,
which is projective and finitely generated as an $R$--module. Let $$F(X):={\rm det}_{R}(X-\mathfrak m_g\mid M),$$ where $\mathfrak m_g$ is the $R[[\Gamma]]$--module automorphism of $M$ given
by multiplication by $g$. Then, the following hold.
\begin{enumerate}\item $M$ is finitely presented as an $R[[\Gamma]]$--module. Also, if we let $F(g)$ be the image of $F(X)$ via the $R$--algebra morphism $R[X]\to R[[\Gamma]]$ sending $X$ to $g$, we have an equality of $R[[\Gamma]]$-ideals
$${\rm Fit}_{R[[\Gamma]]}(M)=(F(g))\,.$$
\item
Let $M^\ast_R:={\rm Hom}_R(M, R)$, viewed as a
topological $R[[\Gamma]]$--module with the covariant
$\Gamma$--action, given by $\sigma\cdot
f(x):=f(\sigma\cdot x)$, for all $f\in M^\ast_R$,
$\sigma\in\Gamma$ and $x\in M$. Then, we have
$${\rm Fit}_{R[[\Gamma]]}(M)={\rm Fit}_{R[[\Gamma]]}(M^\ast_R)\,.$$
\item Assume that $R=\zp[G]$, where $G$ is
a finite, abelian group and $p$ is a prime number. Let $M^\ast:={\rm Hom}_{\zp}(M, \zp)$, viewed as an
$R[[\Gamma]]\simeq\zp[[G\times\Gamma]]$--module with the covariant $G\times\Gamma$--action. Then, we have
$${\rm Fit}_{R[[\Gamma]]}(M^\ast)={\rm Fit}_{R[[\Gamma]]}(M).$$
\end{enumerate}
\end{proposition}
\begin{proof} See Proposition 4.1 and Corollary 4.2 in \cite{GP}.\end{proof}
\begin{proposition}\label{four-term-sequence-fitting}
Let $R:=\zp[G]$, for some finite abelian group $G$ and prime number $p$. Assume that we have an exact sequence
of finite $R$--modules
$$\xymatrix{0\ar[r] &A\ar[r] &P\ar[r] &P'\ar[r] &A'\ar[r] &0.
}$$ Further, assume that ${\rm pd}_{\zp[G]}P\leq 1$ and ${\rm pd}_{\zp[G]}P'\leq 1$. Then, we have
$${\rm Fit}_R(A^\vee)\cdot{\rm Fit}_R(P')={\rm Fit}_R(A)\cdot{\rm Fit}_R(P),$$
where the dual $A^\vee:={\rm Hom}(A, \qp/\zp)$ is endowed with the covariant $G$--action.
\end{proposition}
\begin{proof}See Lemma 5, p.179 of \cite{Burns-Greither}. In loc.cit. this result is proved for
general finitely generated $\zp$--algebras $R$ which are $\zp$--free and relatively Gorenstein over $\zp$. Also, note
that ${\rm pd}_{\zp[G]}P=0$ if and only if $P=0$ (and similarly for $P'$), in which case the equality above is immediate.\end{proof}
\subsection{\bf Twisting}\label{appendix-twisting} In what follows, we fix an odd prime $p$, a field $k$ of characteristic different from $p$ and a Galois extension
$\ck/k$. We let $\cg:={\rm Gal}(\ck/k)$ and assume that the group of $p$--power roots of unity $\bmu_{p^\infty}$ is contained
in $\ck$. As usual, we denote by $c_p:\cg\to\zp^\times={\rm Aut}(\bmu_{p^\infty})$ the $p$--cyclotomic character
of $\cg$ and decompose $c_p:=\omega_p\cdot\kappa_p$ in its tame (Teichm\"uller) and wild components, $\omega_p:\cg\to\bmu_{p-1}$ and $\kappa_p:\cg\to(1+p\zp)$, respectively.
For simplicity, we let $c:=c_p$, $\omega:=\omega_p$ and $\kappa:=\kappa_p$.
Let $\co$ be a finite, integral extension of $\zp$. We let $Q(\co)$ denote its field of fractions. We consider the unique continuous isomorphisms
of $\co$--algebras
$${t_n}: \co[[\cg]]\overset\sim\longrightarrow \co[[\cg]],\qquad \iota: \co[[\cg]]\overset\sim\longrightarrow \co[[\cg]]^{\rm op}$$
satisfying $t_n(g)=c_p(g)^n\cdot g$ and $\iota(g)=g^{-1}$, for all $g\in \cg$ and all $n\in\z$.
For any $\co[[\cg]]$--module $M$ and any $n\in\z$, we let $M(n)$ denote the usual Tate twist of $M$ by $c_p^n$. More precisely, $M(n):=M$ with a new $\co[[\cg]]$--action given by
$\lambda\ast x:=t_n(\lambda)\cdot x$, for all $\lambda\in \co[[\cg]]$ and $x\in M$. Also, we let $M^\ast:={\rm Hom}_\co(M, \co)$ and view it as an $\co[[\cg]]$--module with the contra-variant $\cg$--action
given by $g\cdot f(x)=f(\iota(g)\cdot x)$, for all $f\in M^\ast$, $x\in M$ and $g\in\cg$. Throughout, $\co$ and $Q(\co)$ are viewed as a $\co[[\cg]]$--modules
with the trivial $\cg$--action. Also, if $M$ and $N$ are $\co[[\cg]]$--modules, the $M\otimes_\co N$ is viewed as an $\co[[\cg]]$--module with the diagonal $\cg$--action.
\begin{lemma}\label{twisting-Fitting}
Assume that $\cg$ is abelian and $M$ is a finitely presented $\co[[\cg]]$--module. Then, for all $n\in\z$, the following hold.
\begin{enumerate}
\item ${\rm Ann}_{\co[[\cg]]}(M(n))=t_{-n}\left({\rm Ann}_{\co[[\cg]]}(M) \right)$;
\item ${\rm Fit}_{\co[[\cg]]}(M(n))=t_{-n}\left({\rm Fit}_{\co[[\cg]]}(M) \right)$;
\item ${\rm pd}_{\co[[\cg]]}M={\rm pd}_{\co[[\cg]]} M(n)$;
\item If $k:=\Bbb F_q$ is the finite field of $q$ elements ($p\nmid q$), $\sigma_q\in\cg$ is the $q$--power Frobenius
automorphism of $\ck$, and $M:=T_p(\ck^\times)$, then
$${\rm Fit}_{\zp[[\cg]]}(M(n))={\rm Ann}_{\zp[[\cg]]}(M(n))=(1-q^n\cdot\sigma_q^{-1}).$$
\end{enumerate}
\end{lemma}
\begin{proof} For the easy proof of (1) and (2), see Lemma 3.1 in \cite{Popescu-CS}. Part (3) is an immediate consequence of the existence of a unique
$\zp[[\cg]]$--module isomorphism $\zp[[\cg]]\simeq\zp[[\cg]](n)$ which sends $g\to c(g)^ng$, for all $g\in\cg$. Part (4) is derived from part (2) as follows.
Under the hypotheses of part (4), $\cg$ is a pro--cyclic group of (topological) generator $\sigma_q$. Since $\zp$ is a cyclic
$\zp[[\cg]]$--module, (\ref{ann-fitt}) gives
$$\quad {\rm Fit}_{\zp[[\cg]]}(\zp)={\rm Ann}_{\zp[[\cg]]}(\zp)=(1-\sigma_q^{-1})\,.$$
Now, note that $T_p(\ck^\times)=T_p(\bmu_{p^\infty})=\zp(1)$ and $c(\sigma_q)=q$, and apply (1) with $M:=\zp$.
\end{proof}
Throughout the rest of this subsection, we will assume that $\ck$ is a finite, abelian extension of Galois group $G$ of the cyclotomic
$\zp$--extension $\ck'$ of $k$ and that $\cg$ is abelian. Consequently, there is a non-canonical group isomorphism $\cg\simeq G\times\Gamma$,
where $\Gamma\simeq\zp$. We fix a topological generator $\gamma$ of $\Gamma$. We identify $\co[[\cg]]$, $\co[G][[\Gamma]]$ and $\co[G][[t]]$ via the obvious $\co[G]$--algebra isomorphisms
$$\co[[\cg]]\simeq \co[G][[\Gamma]]\simeq \co[G][[t]], \qquad \gamma\to (t+1).$$
We let $K:=\ck^\Gamma$ and identify $G$ and $\Gamma$ with $G(K/k)$ and $G(\ck'/k)$ via the usual Galois--restriction isomorphisms.
Note that since $\bmu_{p^\infty}\subseteq\ck$, we have $\bmu_p\subseteq K$. Consequently, $\omega$ and $\kappa$ factor through $G$ and $\Gamma$, respectively.
For simplicity, we assume that $\co$ contains the values of all the irreducible $\cp$--valued characters of $G$. We denote by $\widehat G(\cp)$ the set
of all $\cp$--valued irreducible characters of $G$ and, for all $\chi\in\widehat G(\cp)$, we let $e_\chi:=1/|G|\sum_{\sigma\in G}\chi(\sigma)\cdot\sigma^{-1}$ denote the idempotent associated to $\chi$ in $Q(\co)[G]$.
Each $\chi$ as above will be extended to the unique $Q(\co)[X]$-algebra and $Q(\co)\otimes_\co\co[[\Gamma]]$--algebra morphisms
$$\chi: Q(\co)[G][X]\to Q(\co)[X], \qquad \chi: Q(\co)\otimes_\co\co[[\cg]]\to Q(\co)\otimes_\co\co[[\Gamma]]$$
which send $g\to \chi(g)$, for all $g\in G$, respectively. Also, for a polynomial $P\in Q(\co)[G][X]$, we denote by $P(\gamma)$ (respectively $P(t+1)$) its image
via the unique $Q(\co)[G]$--algebra morphism $Q(\co)[G][X]\to Q(\co)\otimes_\co\co[[\cg]]$ which sends $X\to\gamma$ (respectively $X\to (t+1)$.)
Next, we let $\cl$ denote an $\co[[\cg]]$--module, which is free of finite rank as an $\co$--module. We consider the following $Q(\co)$--vector spaces
$$V:=Q(\co)\otimes_\co\cl, \qquad V^\ast:={\rm Hom}_{Q(\co)}(V, Q(\co))\simeq Q(\co)\otimes_\co\cl^\ast,$$
endowed with the usual $Q(\co)\otimes_\co\co[[\cg]]$--module structures.
We consider the following polynomials in $Q(\co)[G][X]$ and $Q(\co)[X]$, respectively:
$$P_V(X):={\rm det}_{Q(\co)[G]}(X-\mathfrak m_\gamma\mid V), \quad P_{V, \chi}(X):=\chi(P_V(X))={\rm det}_{Q(\co)}(X-\mathfrak m_\gamma\mid e_\chi V),$$
where $\mathfrak m_\gamma$ denotes the automorphism of $V$ and $e_\chi V$ given by multiplication with $\gamma$, for all $\chi\in\widehat G(\cp)$.
Observe that, for all $\chi\in\widehat G(\cp)$ and all $n\in\z$, we have
\begin{equation}\label{twist-poly}
\chi(t_n(P_{V}(\gamma)))=P_{V, \chi\omega^n}(\kappa(\gamma)^n\gamma), \quad \chi((\iota\circ t_n)(P_{V}(\gamma)))=P_{V, \chi^{-1}\omega^n}(\kappa(\gamma)^n\gamma^{-1}).\end{equation}
\begin{lemma}\label{twist-poly-lemma} The following hold for all $n\in\z$ and all $\chi\in\widehat G(\cp)$.
\begin{enumerate}
\item If $\cl$ is a projective $\co[G]$--module, then $P_{V(n)}(X),\, P_{V^\ast(n)}(X)$ are monic polynomials in $\co[G][X]$.
\item $P_{V(n), \chi}(X),\, P_{V*(n), \chi}$ are monic polynomials in $\co[X]$.
\item $ P_{V(n), \chi}(\gamma)\sim \chi(t_{-n}(P_{V}(\gamma)))$ and $P_{V^\ast(n), \chi}(\gamma)\sim \chi((\iota\circ t_n)(P_{V}(\gamma)))$, where ``$\sim$'' denotes association in divisibility in the ring $\co[[\Gamma]]$.
\end{enumerate}
\end{lemma}
\begin{proof} Fix an $n\in\z$ and note that we have $Q(\co)\otimes_\co\co[[\cg]]$--module isomorphisms
\begin{equation}\label{isos}
V(n)\simeq Q(\co)[G]\otimes_{\co[G]}\cl(n), \qquad V^\ast(n)\simeq Q(\co)[G]\otimes_{\co[G]}\cl^\ast(n).
\end{equation}
In order to prove (1), first note that if $\cl$ is $\co[G]$--projective, then the $\co[G]$--modules $\cl(n)\simeq\cl\otimes_{\co}\co(n)$ and
$\cl^\ast(n)\simeq\cl(-n)^\ast$ are projective as well. Indeed, since the modules in question are $\co$--free and $\co$ is a PID, their $\co[G]$--projectivity is equivalent to their $G$--cohomological triviality
(See \cite{Serre-Local}, Ch. IX, \S5, Theorem 7 for $\z[G]$--modules. The same proof works for $\co[G]$--modules, for a general PID $\co$.)
Now, apply the Corollary to Proposition 1 in \cite{Serre-Local}, Ch. IX, \S3 to arrive at the desired result.
(Loc.cit. deals with the case of $\z[G]$--modules. The same argument works for $R[G]$--modules, where $R$ is a PID, in particular $R=\co$.) Now, part (1) follows by applying \eqref{det-base-change} to conclude that
$$P_{V(n)}(X)={\rm det}_{\co[G]}(X-\mathfrak m_\gamma\mid \cl(n)), \qquad P_{V^\ast(n)}(X)={\rm det}_{\co[G]}(X-\mathfrak m_\gamma\mid \cl^\ast(n))\,.$$
The monicity follows from Remark \ref{monicity-remark}.
Part (2) follows similarly: First, one uses \eqref{isos} to conclude that there are isomorphisms of $Q(\co)\otimes_\co[[\Gamma]]$--modules
$e_\chi \cdot V(n)\simeq Q(\co)\otimes_\co e_\chi\cdot\cl(n)$ and $e_\chi\cdot V^\ast(n)\simeq Q(\co)\otimes_\co e_\chi\cdot\cl^\ast(n)$, where $e_\chi\cdot\cl(n)$ and $e_\chi\cdot\cl^\ast(n)$
are viewed as (necessarily free) $\co$--submodules of maximal rank in $e_\chi\cdot V(n)$ and $e_\chi\cdot V^\ast(n)$, respectively. Then, \eqref{det-base-change} gives the equalities
$$P_{V(n), \chi}(X)={\rm det}_{\co}(X-\mathfrak m_\gamma\mid e_\chi\cdot\cl(n)), \qquad P_{V^\ast(n), \chi}(X)={\rm det}_{\co}(X-\mathfrak m_\gamma\mid e_\chi\cdot\cl^\ast(n)),$$
which conclude the proof of part (2).
Next, we prove the second ``$\sim$'' in part (3). The first ``$\sim$'' is proved similarly. For every $\chi\in\widehat G(\cp)$, we fix an $\co$--basis $\mathbf x_\chi$ of $e_\chi\cl$. Note that $\mathbf x_\chi$ is also a $Q(\co)$--basis of $e_\chi V$. Let $A_{\gamma, \chi}$ be the matrix of
$\mathfrak m_\gamma$ restricted to $e_\chi V$ with respect to this basis. Then, $A_{\gamma, \chi}\in {\rm GL}_{m_\chi}(\co)$, where $m_\chi={\rm rk}_\co e_\chi\cl={\rm dim}_{Q(\co)}e_\chi V$. It is easily proved that for all $n$ and $\chi$ as above, the matrix of $\mathfrak m_\gamma$ restricted to $e_\chi\cdot V^\ast(n)=(e_{\chi^{-1}\omega^n} V)^\ast$ with respect to the basis $\mathbf x_{\chi^{-1}\omega^n}^\ast$ (dual basis of $\mathbf x_{\chi^{-1}\omega^n}$) is $\kappa(\gamma)^n\cdot(A_{\gamma, \chi^{-1}\omega^n}^{-1})^t$,
where ${}^t$ stands for transposition. Consequently, \eqref{twist-poly} combined with the fact that $\det(A_{\gamma, \chi^{-1}\omega^n})\in \co^\times$ and $\gamma\in \co[[\Gamma]]^\times$ imply that the following hold in $\co[[\Gamma]]$:
\begin{eqnarray}
\nonumber P_{V^\ast(n), \chi}(\gamma) &=&\det(\gamma\cdot I_{m_{\chi^{-1}\omega^n}}-\kappa(\gamma)^n\cdot(A_{\gamma, \chi^{-1}\omega^n}^{-1})^t) \\
\nonumber &\sim & \det(\kappa(\gamma)^n\gamma^{-1}\cdot I_{m_{\chi^{-1}\omega^n}}-A_{\gamma, \chi^{-1}\omega^n})\\
\nonumber &=& \chi((\iota\circ t_n)(P_{V}(\gamma)))
\end{eqnarray} This concludes the proof of the Lemma.
\end{proof}
\subsection{\bf Equivariant Power Series.}\label{appendix-power-series} (Compare with \S2 in \cite{Burns-Greither}.) Let $G$ be an arbitrary finite abelian group, $p$ a prime, and $\co$ a finite, integral extension of $\zp$ which contains the values of all characters $\chi\in\widehat G(\cp)$. We let $\pi$ be a uniformizer of $\co$. We identify the set of $\zp$--algebra morphisms ${\rm Hom}(\zp[G], \co)$ with the set
$\chi\in \widehat G(\cp)$ of $\cp$--valued characters of $G$ in the obvious manner. We let $I$ denote a radical ideal of $\zp[G]$ of pure codimension $1$.
It is easily seen that this means that $I=\cap_{\chi\in\mathcal F}\ker(\chi)$, for
some set $\mathcal F\subseteq {\rm Hom}(\zp[G], \co)$. In what follows, we let $\fa:=\zp[G]/I$.
Obviously, $\mathcal F$ coincides with the set of $\zp$--algebra morphisms ${\rm Hom}(\fa, \co)$.
Also, we have an injective $\zp$--algebra morphism
$$\fa\longrightarrow \oplus_{\chi}\, \co,\qquad x\longrightarrow (\chi(x))_{\chi\in {\rm Hom}(\fa, \co)}.$$
For every $\chi\in{\rm Hom}(\fa, \co)$, we abuse notation once again and
let $\chi$ also denote the unique $\fa[[t]]$--algebra morphism
$$\chi: \fa[[t]]\longrightarrow \co[[t]],$$ which sends $x\to\chi(x)$, for all $x\in \fa.$
\begin{definition}\label{mu-power-series}
\begin{enumerate}
\item The $\mu$--invariant $\mu(f)$ of a power series $f\in \co[[t]]$
is the largest exponent $r\in\Bbb Z_{\geq 0}$, such that $f\in\pi^r\co[[t]]$.
\item A power series $F\in \fa[[t]]$ is said to have $\mu$--invariant equal to $0$ (and we write $\mu(F)=0$) if
$$\mu(\chi(F))=0, \qquad\text{ for all }\chi\in{\rm Hom}(\fa, \co).$$
\item A polynomial $F\in \fa[t]$ is said to be Weierstrass if $\chi(F)$ is a Weierstrass polynomial in $\co[t]$
(i.e. $\chi(F)$ is monic and all its non--leading coefficients are divisible by $\pi$), for all $\chi$ as above.
\end{enumerate}
\end{definition}
\smallskip
\begin{remark} The rings $\fa$ considered above are the most general reduced quotients of $\zp[G]$ of pure Krull dimension $1$.
It is easy to see that any ring $\fa$ as above is {\rm admissible}, in the sense of \cite{Burns-Greither}, \S2.
Also, it is easy to prove that our definition of power series $F\in\fa[[t]]$ of $\mu$--invariant equal to $0$ is equivalent with
the definition in loc.cit., for all rings $\fa$ as above.
\end{remark}
In the following, if $R$ is a commutative ring with $1$, and $f, g\in R$, we write ``$f\sim g$ in $R$'' to mean that $f$ and $g$ are associated in divisibility in $R$,
i.e. there exists a unit $u\in R^\times$, such that $f=u\cdot g$.
\begin{lemma}\label{associated-in-divisibility} Assume that $F, \Theta\in \fa[[t]]$, such that $\mu(F)=\mu(\Theta)=0$ and
$$\chi(F)\sim\chi(\Theta) \text{ in } \co[[t]], \qquad \text{ for all }\,\chi\in{\rm Hom}(\fa, \co).$$
Then, we have $F\sim \Theta$ in $\fa[[t]]$.
\end{lemma}
\begin{proof} Since $\mu(F)=\mu(\Theta)=0$, Proposition 2.1 in \cite{Burns-Greither} (the equivariant Weierstrass preparation theorem) shows that there exist unique Weierstrass polynomials $f, g\in \fa[t]$ and units $u, v\in \fa[[t]]^\times$, such that
$$F=u\cdot f, \quad \Theta= v\cdot \theta.$$
This implies that we have Weierstrass decompositions in $\co[[t]]$
$$\chi(F)=\chi(u)\cdot\chi(f), \quad \chi(\Theta)= \chi(v)\cdot\chi(\theta),$$
with $\chi(u), \chi(v)\in \co[[t]]^\times$ and $\chi(f), \chi(\theta)$ Weierstrass polynomials in $\co[t]$, for all $\chi\in{\rm Hom}(\fa, \co)$. However, since the Weierstrass decomposition is unique
in $\co[[t]]$ (according to the classical Weierstrass preparation theorem), our hypotheses combined with the above equalities imply that
$$\chi(f)=\chi(\theta), \qquad \text{ for all }\, \chi\in{\rm Hom}(\fa, \co).$$
Consequently, we have $f=\theta$ and $F=uv^{-1}\cdot \Theta$. Therefore, $F\sim\Theta$ in $\fa[[t]]$.
\end{proof}
Now, let us assume that $p$ is odd and that $G$ has an element $j$ of order $2$. Let $\zp[G]^-:=\zp[G]/(1+j)$ and call
a character $\chi\in\widehat G(\cp)$ odd if $\chi(j)=-1$. Clearly, $I:=(1+j)=\cap_{\chi}\ker(\chi)$, where
$\chi$ runs through the odd characters of $G$. Therefore, $\fa:=\zp[G]^-$ is a reduced quotient of $\zp[G]$ of pure Krull dimension $1$. Therefore, the following is a direct consequence of the above Lemma.
\begin{corollary}\label{corollary-association} Let $F, \Theta\in\zp[G]^-[[t]]$, such that
$$\mu(\chi(F))=\mu(\chi(\Theta))=0, \qquad\chi(F)\sim\chi(\Theta) \text{ in $\co[[t]]$,}$$
for all odd $\chi\in\widehat G(\cp)$.
Then, we have $F\sim\Theta$ in $\zp[G]^-[[t]]$.
\end{corollary}
\bibliographystyle{plain}
|
\section{Introduction}
Modern developments and applications of quantum mechanics often involve complex chemical and even biological systems driven by laser fields (see for instance \cite{Assion}). Solving (numerically) the time dependent Schrodinger equation (TDSE) for such time dependent systems becomes then very time consuming and sometimes even impossible.
Finding numerical simplifications is an active research. One can for instance mention the multi-configuration time-dependent Hartree (MCTDH) algorithm \cite{Meyer}.
Techniques that lead to an efficient propagation of a time-dependent problem often involve the Floquet theory which allows one to incorporate fast oscillations of the external field (for instance such as the optical oscillations of a laser field) in an enlarged Hilbert space \cite{shirley}. For instance, it permits an adiabatic separation between the fast field oscillation dynamics and the slow time modulation
of the field envelope (adiabatic Floquet theory \cite{reviewguerin}). This Floquet technique can be used alternatively to treat the full time-dependence of the field, which is referred to as the $(t,t')$ approach \cite{peskin}.
Relevant processes are most often expected to be described in a small subspace, often named active space, through effective Hamiltonians. One can mention in particular the time-dependent wave operator theory (TDWOT) as a tool to extract dynamical active spaces \cite{reviewgeorges2}.
A few years ago Jolicard et al. \cite{CATM} have proposed the ``Constrained Adiabatic Trajectory Method'' (CATM)
for solving the TDSE for a time-dependent potential.
Since we use a quantum mechanical approach the trajectory studied in the
CATM is not a classical one but rather a constrained path followed by the wavefunction as it develops in time in a
composite Hilbert space which we describe below.
We here investigate that method
extending it for an initial condition as a general superposition of states for a small system,
and emphasizing its principal novel feature, the
use of a complex absorbing potential which is itself time-dependent.
The usual approach is to propagate the wavefunction in small time steps, with the Hamiltonian
considered as constant over each step \cite{leforestier,kosloff}. The radically different approach of the CATM
is to limit the time development to only one term in a Floquet expansion of the wavefunction,
achieving this by a careful choice of the varying absorbing potential.
The problem of integrating the TDSE then becomes that of finding one eigenvector of the system's
Floquet Hamiltonian. The method has some affinities with the $(t,t')$ approach \cite{peskin}
but represents a modification and improvement of it. The method finds the wavefunction
at regular grid points throughout the interaction period, the principal requirement being to work
with a sufficient number of points to describe the time-varying Hamiltonian and
to allow the stable use of Fast Fourier transforms.
In brief, the technique requires the wavefunction corresponding to the dynamics to connect to a single Floquet state, referred to as a constrained Floquet state (CFS), through the use of an artificial absorbing potential (or optical potential). The second role of the absorbing potential is to dilate the Floquet spectrum isolating well the eigenvalue corresponding to the CFS from the other ones. In practice one thus needs to find this CFS to determine the dynamics.
In Section II, on the basis of Ref. \cite{CATM}, we summarize the technique CATM with its corresponding Floquet representation, and recall the result when the initial condition is a single eigenstate of the free system. In Section III, we extend the technique to a more general initial state, as a superposition of eigenstates of the free system. This is analyzed for a two-state system.
A forthcoming work will treat the case of systems of higher dimension.
In Section IV, we give an analytic treatment of the effect of the absorbing potential on the Floquet spectrum for a two-state model. The numerical limitations of the method and its accuracy are analyzed in section V.
We study examples with two- and three-level models which illustrate the dual role played by the optical potential in Section VI. Section VII is devoted to the conclusion.
\section{The constrained adiabatic trajectory method}
\subsection{The Floquet representation}
We assume a system of Hamiltonian $H(q,t)$ (where the quantum coordinates have been denoted by $q$) defined in a basis $\{|j\rangle\}$. This Hamiltonian can be usually decomposed as $H(q,t)=H_0(q)+W(q,t)$ featuring a free system $H_0(q)$ subjected to an external time dependent field corresponding to the interaction potential $W(q,t)$. In that case, $\{|j\rangle\}$ correspond to the states of the free system. We assume that the interaction potential $W(q,t)$ acts on a duration $t\in[0,T]$ referred to as the physical duration in the following.
We define an extra time interval $[T,T']$ after the physical interaction time during which (i) we add an artificial time-dependent absorbing (or optical) potential ${\cal{V}}(q,t)$ satisfying
${\cal{V}}(q,0\le t\le T)={\cal{V}}(q,T')=0$, and (ii) we extend continuously the interaction such that $W(q,T')=W(q,0)$. This construction features a periodic Hamiltonian $H(q,T')=H(q,0)$.
We can thus define the corresponding Floquet Hamiltonian (or quasi-energy operator) on the extended Hilbert space (product
of the original Hilbert space, i.e. associated to the free system, by the space of $T'$-periodic functions)~\cite{reviewguerin}:
\begin{equation}
H_F(q,t)=H_0(q)+W(q,t)+{\cal{V}}(q,t)-i\hbar \frac{\partial}{\partial t}.
\label{hamiltonian}
\end{equation}
We consider the entire duration of the interaction+absorbing potential as a fundamental period $T'$ ($\omega_0 = 2\pi / T'$), contrary
to the traditional Floquet scheme in which $T'$ is associated with the period of an external field (such as the optical period of a laser field).
The Floquet states can be indexed with a double labelling $j,n$ linked to a finite basis representation of the decoupled parts of \eqref{hamiltonian}, i.e. to
the free-system eigenstates
($j\leftrightarrow \vert j \rangle$) and to the operator $-i\hbar \partial_t$ (corresponding here to a Fourier basis, $n \leftrightarrow |n\rangle\equiv\vert e^{-i n \omega_0 t} \rangle$).
Thus a complete basis is formed with the eigenstates $\{ \vert \lambda_{j,n} (q,t) \rangle\}$~of~$H_F$~:
\begin{equation}
H_F \vert \lambda_{j,n} (q,t) \rangle = \hbar\omega_{\lambda_{j,n}} \vert \lambda_{j,n} (q,t) \rangle.
\label{eq2}
\end{equation}
Using the periodicity properties of the Floquet theorem ($\ket{\lambda_{j,n} (q,t)}=\ket{\lambda_{j,0} (q,t)}e^{in\omega_0t}$, $\omega_{\lambda_{j,n}}=\omega_{\lambda_{j,0}}+n\omega_0$),
we can rigorously expand the solution of the time dependent Schr\"odinger equation (TDSE)
with an initial limitation to the first Brillouin zone \cite{shirley},
\begin{eqnarray}
\label{dev}
\vert \Psi (q,t)\rangle &=& \sum_{j} \langle \lambda_{j,0} (q,0) \vert \Psi (q,0) \rangle
e^{-i \omega_{\lambda_{j,0}} t} \vert \lambda_{j,0} (q,t) \rangle.\quad\
\end{eqnarray}
(Here we consider for simplicity only a bound spectrum, that can feature however imaginary parts;
the extension to a system with bound and continuous spectrum is in principle direct assuming a discretization of the continua).
Usually, a great number of $\vert \lambda_{j,0} \rangle$ is necessary to reconstruct $\vert \Psi (q,t) \rangle$. An interesting practical application of Eq. \eqref{dev}
is the development of a very reduced number of Floquet vectors, and in
the best case of only one, which is the key idea of the CATM.
The method developed in this paper deals with the case of a single constrained Floquet state (CFS) and labeled $\ell$ in the expansion \eqref{dev}.
In this case, the CFS has to match,
when projected at $t=0$, with the initial boundary conditions required for the wavefunction $\Psi(t=0)$:
\begin{equation}
\langle \lambda_{j,0} (q,0) \vert \Psi (q,0) \rangle=\delta_{j,\ell},
\end{equation}
i.e.
\begin{equation}
\vert \Psi(q,t) \rangle = e^{-i \omega_{\lambda} t } \vert \lambda (q,t) \rangle,
\label{wavefunction}
\end{equation}
where we have omitted in the latter the index $\ell$ for simplicity: $\omega_{\lambda}\equiv \omega_{\lambda_{\ell,0}}$, $\vert \lambda (q,t) \rangle\equiv \vert \lambda_{\ell,0} (q,t) \rangle$.
We will show below that in practice we do not get the exact equality \eqref{wavefunction} but a proportionality through a well defined complex phase.
\subsection{Initial condition as an eigenstate of the free system}
Jolicard et al. \cite{CATM} provided the matching with the initial condition for
an initial state equal to a single state $\vert i \rangle$ of the $\{\vert j \rangle\}$ basis, i.e. $\vert \Psi (q,0) \rangle = \vert i \rangle$.
The connection between the Floquet eigenstate and the required initial state is made thanks to the addition of the absorbing potential ${\cal{V}}$ on the extra interval $[T,T']$.
Below we summarize this scheme and extend it in the next section to any required initial condition for the particular case of a two-state system.
In order to satisfy Eq.(\ref{wavefunction}) (with a proportionality instead of the equality), it is sufficient to have the connection at $t=0$:
\begin{equation}
\vert \lambda (q,0)\rangle \propto \vert i \rangle.
\label{CI1}
\end{equation}
We remark that $\vert \lambda (q,t)\rangle$ is a Floquet vector of the extended Hilbert space, but fixing $t$ to a specific value leads to a component of this vector of dimension of the original Hilbert space.
Eq.(\ref{CI1}) suggests the use of the following form for the absorbing potential:
\begin{equation}
{\cal{V} }(t) = \sum_{j\neq i} -i V_{\text{opt}}(t) \vert j \rangle \langle j \vert
\label{potentiel}
\end{equation}
with $V_{\text{opt}}(t)$ zero over $[0,T]$ and positive over $[T,T']$.
As shown in \cite{CATM}, provided that
\begin{equation}
\label{cond}
\frac{1}{\hbar} \int_{T}^{T'} V_{\text{opt}}(t) dt \gg |\Im (\omega_{\lambda})| (T'-T)
\end{equation}
we can be sure that all channel except $\vert i \rangle$
are absorbed and that Eq.(\ref{CI1}) is satisfied to a good approximation (as will be tested in section \ref{inspection}).
\section{Extension of the CATM to a general initial condition: The two-state case}
\label{Extension}
If we wish to work with CATM in the case of an initial condition as a state superposition of the free system, i.e.
\begin{equation}
\ket{\Psi(0)} = \sum_{j}c_j \ket{j},
\end{equation}
then simple forms as (\ref{potentiel}) no longer work. (From now on, we do not write explicitly the dependence on the $q$ coordinates.)
We provide below the relevant absorbing potential that should be used for a two-state system of Hamiltonian
\begin{equation}
H(t)=\hbar\left(
\begin{array}{ll}
\Delta_1(t) & \Omega(t) \\
\Omega^{\ast}(t) & \Delta_2(t)
\end{array}
\right).
\end{equation}
We consider the most general case with diagonal terms that are time dependent (due to Stark shifts of the states for instance) and complex (i.e. including their lifetime).
We assume that the coupling $\Omega(t)$ is in general different from zero only over the physical time interval $[0,T]$. During the extra time interval, the diagonal terms have to be continuously varied such that they recover their initial value in order to guarantee the periodicity: $\Delta_j(T')=\Delta_j(0)$.
We show below that it is possible to treat any initial condition by adding
the following absorbing potential over the supplementary interval $[T,T']$:
\begin{eqnarray}
\label{potgen}
{\cal{V}}(t) & =& \left(
\begin{array}{ll}
0 & 0 \\
- \frac{c_2e^{i\int_t^{T'}\Delta_2(t')dt'}}{c_1e^{i\int_t^{T'}\Delta_1(t')dt'}} & 1
\end{array}
\right) \times \left(-i V_{\text{opt}}(t)\right) \\
\text{ with } & &V_{\text{opt}}(t) > 0 \quad \forall t \in ]T,T'[ \nonumber\\
& &V_{\text{opt}}(t) = 0 \quad \forall t \in [0,T]. \nonumber
\end{eqnarray}
The operator
\begin{eqnarray}
\Pi(t) & =& \left(
\begin{array}{ll}
0 & 0 \\
- \frac{c_2e^{i\int_t^{T'}\Delta_2(t')dt'}}{c_1e^{i\int_t^{T'}\Delta_1(t')dt'}} & 1
\end{array}
\right)
\end{eqnarray}
involved in this definition \eqref{potgen} is a non-orthogonal (i.e. non self-adjoint) projector, i.e. $\Pi^2=\Pi$, whose kernel is the initial state up to a phase correction:
\begin{equation}
\Pi(t)\left(
\begin{array}{c}
c_1 e^{i\int_t^{T'}\Delta_1(t')dt'}\\
c_2e^{i\int_t^{T'}\Delta_2(t')dt'}
\end{array}
\right)=0.
\end{equation}
For this case it is indeed possible to obtain the analytical asymptotic form of the Floquet eigenvector
over the extra interval $[T,T']$, where the Hamiltonian contains just the free system Hamiltonian and
the absorbing potential. With the above definition and writing Floquet components
$\langle j \vert \lambda (t) \rangle = \lambda_j (t)$, one must solve on $[T,T']$
the following system :
\begin{subequations}
\begin{eqnarray}
\frac{\partial \lambda_1 (t)}{\partial t} &=& i(\omega_{\lambda}-\Delta_1(t)) \lambda_1 (t) \\
\frac{\partial \lambda_2 (t)}{\partial t} &=&
\frac{V_{\text{opt}}(t)}{\hbar} \frac{c_2e^{i\int_t^{T'}\Delta_2(t')dt'}}{c_1e^{i\int_t^{T'}\Delta_1(t')dt'}} \lambda_1 (t) \nonumber\\
&&- \left(
\frac{V_{\text{opt}}(t)}{\hbar} -i\left(\omega_{\lambda}-\Delta_2(t)\right)\right) \lambda_2 (t)\qquad
\end{eqnarray}
\end{subequations}
The first component follows an exponential law:
$ \lambda_1 (t) = \lambda_1(T) e^{i(\omega_{\lambda}(t-T)-\int_T^t\Delta_1(t')dt')}$.
This function can be introduced in the second equation and making use of
the identity $\int_T^t V_{\text{opt}}(t') e^{\frac{1}{\hbar} \int_T^{t'} V_{\text{opt}}(t'') dt''}
dt'=\hbar \left( e^{\frac{1}{\hbar} \int_T^t V_{\text{opt}} (t') dt'}-1\right)$, we find
\begin{align}
\lambda_2(t)&=\lambda_2(T) e^{i (\omega_{\lambda}(t-T)-\int_T^t\Delta_2(t')dt')} e^{-\frac{1}{\hbar}\int_T^t V_{\text{opt}}(t')dt'} \nonumber\\
&+\frac{c_2}{c_1} \lambda_1(T) e^{i\omega_{\lambda}(t-T)}
e^{i\int_{T'}^T\Delta_1(t')dt'}e^{i\int_t^{T'}\Delta_2(t')dt'}\nonumber\\
&\times(1-e^{-\frac{1}{\hbar}\int_T^t V_{\text{opt}}(t')dt'}).
\label{expdec}
\end{align}
Taking into account the $\lambda$ periodicity $\lambda_j(T')\equiv \lambda_j(0)$, we obtain
\begin{align}
\label{ratio21}
\frac{\lambda_2(0)}{\lambda_1(0)}&=\frac{\lambda_2(T)}{\lambda_1(T)} e^{-\frac{1}{\hbar}\int_T^{T'} V_{\text{opt}}(t)dt} e^{i\int_T^{T'}(\Delta_1(t)-\Delta_2(t))dt}\nonumber\\
&+\frac{c_2}{c_1} (1-e^{-\frac{1}{\hbar}\int_T^{T'} V_{\text{opt}}(t)dt}),
\end{align}
which, in the limits
\begin{subequations}
\label{cond_}
\begin{align}
\label{cond_1} \frac{1}{\hbar}\int_{T}^{T'} V_{\text{opt}}(t) dt &\gg 1,\\
\label{cond_2} \frac{1}{\hbar}\int_{T}^{T'} V_{\text{opt}}(t) dt &\gg \int_T^{T'}[\Im(\Delta_2(t))-\Im(\Delta_1(t))]dt
\end{align}
\end{subequations}
and for $\lambda_2(T)$ and $\lambda_1(T)$ of the same order, leads to
\begin{equation}
\frac{\lambda_2(0)}{\lambda_1(0)}\leadsto\frac{c_2}{c_1}.
\end{equation}
We remark that, denoting the state-vector of the original TDSE $\ket{\Psi(t)}\equiv\left(\begin{array}{ll}
a_1(t)\\
a_2(t)
\end{array}\right)$, the connection to a single Floquet vector
\eqref{wavefunction} leads to $\lambda_2(T)/\lambda_1(T)=a_2(T)/a_1(T)$, i.e. to the ratio of the amplitude at the end of the process. If this ratio becomes very large, which corresponds to the specific case of an efficient population transfer to state 2, the condition \eqref{cond_2} is not sufficient. It should be replaced in general by the condition:
\begin{align}
\label{cond_2_} \frac{1}{\hbar}\int_{T}^{T'} V_{\text{opt}}(t) dt &\gg \int_T^{T'}[\Im(\Delta_2(t))-\Im(\Delta_1(t))]dt\nonumber\\
&+\log(a_2(T))-\log(a_1(T)).
\end{align}
This is discussed in more detail in Section V.B.
For an initial condition as a single state of the free system, ie. $c_2=0$, $c_1=1$, one recovers $\lambda_2(0)\ll\lambda_1(0)$ \cite{CATM}.
In this case, we must note that conditions (\ref{cond_}) are less restrictive than condition (\ref{cond}).
This is due to the fact that conditions (\ref{cond_}) are obtained constraining a ratio of two components, whereas in \cite{CATM} we wished to absorb the components, with an error lower than the computer accuracy.
If the conditions \eqref{cond_} are satisfied, then we can force any eigenstate $\ket{\lambda}$
to obey the final condition
\begin{subequations}
\begin{eqnarray}
\lambda_1(0) &=& \lambda_1(T)e^{i(\omega_{\lambda}(T'-T)-\int_T^{T'}\Delta_1(t)dt)} \\
\lambda_2(0) & \leadsto & \lambda_1(T) e^{i(\omega_{\lambda}(T'-T)-\int_T^{T'}\Delta_1(t)dt)} \frac{c_2}{c_1}
\end{eqnarray}
\end{subequations}
Thus, apart from a global constant $\lambda_1(T)$ which results from the diagonalization procedure, an exponentially decreasing term and a global phase,
we obtain
\begin{equation}
\ket{\lambda (t=0)} \propto \ket{\psi (t=0)}.
\end{equation}
This approximate proportionality is sufficient to impose
the required initial connection to the Floquet eigenvector \eqref{CI1}. This will be illustrated
by an example given in section \ref{inspection}.
\section{Isolating one eigenvalue in the Floquet spectrum \label{eigenvalues}}
The second role of the absorbing potential is to dilate the Floquet spectrum and so isolate the ``connected'' eigenvalue $\hbar\omega_{\lambda}$ (i.e. the one associated to the eigenvector $\vert \lambda(t) \rangle$ connected to the initial condition) from the other eigenvalue (denoted $\hbar\omega_{\lambda'}$ associated to $\vert \lambda' (t) \rangle$).
We consider for simplicity the initial condition as a single bound state $\ket{1}$ of $H_0$. The absorbing potential takes the form set out in Eq. (\ref{potentiel}).
We start connecting the solution $\vert \Psi (q,t)\rangle$ to the Floquet vector. This is achieved by solving the stationary problem (in the first Brillouin zone):
\begin{subequations}
\label{0Tprime}
\begin{eqnarray}
\label{0T}
&&\text{for $t\in[0,T]$}:\nonumber\\
&&\left[- i \frac{\partial}{\partial t}+\left(
\begin{array}{ll}
\Delta_1(t) & \Omega(t) \\
\Omega^{\ast}(t) & \Delta_2(t)
\end{array}
\right)\right]\ket{\lambda(t)}=\omega_{\lambda}\ket{\lambda(t)},\\
&&\text{for $t\in[T,T']$}:\nonumber\\
\label{TTprime}& &\left[- i \frac{\partial}{\partial t}+\left(
\begin{array}{ll}
\Delta_1(t) & 0 \\
0 & \Delta_2(t)-\frac{i}{\hbar}V_{\text{opt}}(t)
\end{array}
\right)\right]\ket{\lambda(t)}=\omega_{\lambda}\ket{\lambda(t)}.\nonumber\\
\end{eqnarray}
\end{subequations}
In the region $t\in[T,T']$, we obtain from \eqref{TTprime} (see the preceding section):
\begin{subequations}
\begin{eqnarray}
\lambda_1 (t) &=& \lambda_1(T) e^{i(\omega_{\lambda}(t-T)-\int_T^t\Delta_1(t')dt')},\\
\lambda_2 (t) &=& \lambda_2(T) e^{i(\omega_{\lambda}(t-T)-\int_T^t\Delta_2(t')dt')}
e^{-\frac{1}{\hbar} \int_T^t V_{\text{opt}} (t') dt'}.\qquad
\end{eqnarray}
\end{subequations}
\subsection{Decoupled channels}
The situation is the easiest to follow in the
elementary case in which the channels are not coupled ($\Omega(t)=0$) and with constant diagonal terms $\Delta_i$.
Thus we make the instant $T$ coincide with $t=0$, to study the influence of the optical potential alone
on the interval $[T=0,T']$ without any physical coupling terms.
In this particular case, with $T=0$ and $t=T'$ the previous system becomes:
\begin{subequations}
\begin{eqnarray}
\lambda_1 (T') &=& \lambda_1(0) e^{i(\omega_{\lambda}-\Delta_1)T'},\\
\lambda_2 (T') &=& \lambda_2(0) e^{i(\omega_{\lambda}-\Delta_2)T'}
e^{-\frac{1}{\hbar} \int_0^{T'} V_{\text{opt}} (t') dt'}.\qquad
\end{eqnarray}
\end{subequations}
The same equations can be written for the other eigenstate $\ket{\lambda'}$.
Floquet eigenvectors must be periodic, i.e. $\lambda_i(T') = \lambda_i(0)$. Thus each Floquet eigenvalue
must satisfy simultaneously two conditions:
\begin{subequations}
\begin{eqnarray}
1 &=& e^{ i (\omega_{\lambda} - \Delta_1)T'} \quad \text{ if} \; \lambda_1(0) \neq 0 \\
1 &=& e^{ i ((\omega_{\lambda} - \Delta_2)T' + \frac{i}{\hbar} \int_0^{T'} V_{\text{opt}}(t) dt)} \quad \text{if} \;\lambda_2(0)\neq 0
\end{eqnarray}
\end{subequations}
The only solution is to have only one non-zero components for each eigenvector:
\begin{subequations}
\begin{eqnarray}
\lambda_1(0)&\neq&0 \quad \text{and}\quad \lambda_2(0)=0 \quad \text{i.e. } \omega_{\lambda} = \Delta_1 \\
\lambda'_1(0)&=&0 \quad \text{and}\quad \lambda'_2(0)\neq0 \nonumber \\
\text{i.e. } \quad &\omega_{\lambda'}& = \Delta_2 - \frac{i}{\hbar T'} \int_0^{T'} V_{\text{opt}}(t)dt
\end{eqnarray}
\end{subequations}
The terms $\frac{2k\pi}{T'}$ are not mentioned because we work in a given Brillouin zone.
In this simpliest case, the extension to a $N-$dimension system is straightforward: all the eigenvalues connected to absorbed channels possess an imaginary term proportional to $\frac{1}{T'}\int_0^{T'} V_{\text{opt}}(t)dt$.
Thus we expect to obtain a dispersion of the eigenvalues in the complex plane which will leave the other eigenvalues
distant from the ``connected`` eigenvalue $\omega_{\lambda}$.
\subsection{General case}
In the present case of a 2-level coupled system described by Eq.(\ref{0Tprime}), it is possible to go further in the analytical description.
In the region $t\in[0,T]$, one can rewrite \eqref{0T} as
\begin{equation}
\label{0T_}
\left[- i \frac{\partial}{\partial t}+\left(
\begin{array}{ll}
\Delta_1(t) & \Omega(t) \\
\Omega^{\ast}(t) & \Delta_2(t)
\end{array}
\right)\right]\ket{\lambda(t)}e^{-i\omega_{\lambda}t}=0,\\
\end{equation}
that is as the same form of the original TDSE of solution $\ket{\Psi(t)}\equiv\left(\begin{array}{ll}
a_1(t)\\
a_2(t)
\end{array}\right)$.
We connect the two solutions invoking the initial conditions $a_1(0)=1$, $a_2(0)=0$, and $ \lambda_1 (0) = \lambda_1(T) e^{i(\omega_{\lambda}(T'-T)-\int_T^{T'}\Delta_1(t)dt)}$, $\lambda_2(0)\simeq 0$ (from the preceding section):
\begin{equation}
\left(\begin{array}{ll}
a_1(t)\\
a_2(t)
\end{array}\right)\lambda_1(T)e^{i(\omega_{\lambda}(T'-T)-\int_T^{T'}\Delta_1(t')dt')}
=\left(\begin{array}{ll}
\lambda_1(t)\\
\lambda_2(t)
\end{array}\right)e^{-i\omega_{\lambda}t}.
\end{equation}
The latter equation is just the proof of the Floquet theorem for our specific two-state problem. Considering the final physical time $t=T$, we get
\begin{equation}
a_1(T)=e^{i\int_T^{T'}\Delta_1(t)dt}e^{-i\omega_{\lambda}T'},
\label{a_1T}
\end{equation}
that is we connect the imaginary part of the eigenvalue $\omega_{\lambda}$ to the final probability amplitude:
\begin{equation}
\label{Elambda}
\Im{(\omega_{\lambda})}=\frac{1}{T'}\int_T^{T'}\Im{(\Delta_1(t))}dt+\frac{1}{T'}\log(|a_1(T)|).
\end{equation}
To get the counterpart relation for the other (``non-connected'') eigenvalue $\omega_{\lambda'}$, we reformulate the complete calculation with the adjoint of $H_F(t)$ (using $\partial_t^{\dagger}=-\partial_t$) :
\begin{equation}
H_F^{\dagger}(t)=H_0^{\dagger}+W^{\dagger}(t)+{\cal{V}}^{\dagger}(t)-i\hbar \frac{\partial}{\partial t}.
\label{hamiltonian_adj}
\end{equation}
of eigenstates $\{ \vert \widetilde\lambda_{j,n} (t) \rangle\}$
\begin{equation}
H_F^{\dagger} \vert \widetilde\lambda_{j,n} (t) \rangle = \hbar\omega_{\lambda_{j,n}}^{\ast} \vert \widetilde\lambda_{j,n} (t) \rangle,
\end{equation}
where $(.)^{\ast}$ denotes the complex conjugate. For real energies of $H_0$ and real elements in $W(t)$, this latter equation corresponds to the same original problem as before but
with the use of an exponentially diverging potential ${\cal{V}}^{\dagger}(t)$.
We have then for the components of $\vert \widetilde\lambda' (t) \rangle$ (denoted as the eigenvector associated to the eigenvalue $\hbar\omega_{\lambda'}^{\ast}$,
$\vert \widetilde\lambda' (t) \rangle$ is different from $\vert \lambda' (t) \rangle$ in general):
\begin{subequations}
\begin{eqnarray}
\widetilde\lambda'_1 (t) &=& \widetilde\lambda'_1(T) e^{i(\omega_{\lambda'}^{\ast}(t-T)-\int_T^t\Delta_1^{\ast}(t')dt')},\\
\widetilde\lambda'_2 (t) &=& \widetilde\lambda'_2(T) e^{i(\omega_{\lambda'}^{\ast}(t-T)-\int_T^t\Delta_2^{\ast}(t')dt')}
e^{\frac{1}{\hbar} \int_T^t V_{\text{opt}} (t') dt'},\qquad
\end{eqnarray}
\end{subequations}
which leads in the limits \eqref{cond_} to
\begin{subequations}
\begin{eqnarray}
\widetilde\lambda'_1 (0)& \ll &\widetilde\lambda'_2 (0),\\
\widetilde\lambda'_2 (0) &=& \widetilde\lambda'_2(T) e^{i(\omega_{\lambda'}^{\ast}(T'-T)-\int_T^{T'}\Delta_2^{\ast}(t)dt)} \nonumber\\
&&e^{+\frac{1}{\hbar} \int_T^{T'} V_{\text{opt}} (t) dt}.
\end{eqnarray}
\end{subequations}
It corresponds to the Schr\"odinger equation
\begin{equation}
\label{0Tp}
\left[- i \frac{\partial}{\partial t}+\left(
\begin{array}{ll}
\Delta_1^{\ast}(t) & \Omega^{\ast}(t) \\
\Omega(t) & \Delta_2^{\ast}(t)
\end{array}
\right)\right]\left(\begin{array}{c}a_1'(t)\\a_2'(t)\end{array}\right)=0\\
\end{equation}
with the initial condition $a_1'(0)=0,a_2'(0)=1$ for which we get
\begin{equation}
a_2'(T)=e^{i\int_T^{T'}\Delta_2^{\ast}(t)dt}e^{-i\omega_{\lambda'}^{\ast}T'}
e^{-\frac{1}{\hbar} \int_T^{T'} V_{\text{opt}} (t) dt}.
\end{equation}
One can connect it to $a_1(T)$ as described in appendix \ref{appA} which induces
\begin{align}
\label{a_1T_}
a_1(T) =e^{-i\int_0^T[\Delta_1(t)+\Delta_2(t)]dt} e^{-i\int_T^{T'}\Delta_2(t)dt} \nonumber \\
e^{+i\omega_{\lambda'}T'}
e^{-\frac{1}{\hbar} \int_T^{T'} V_{\text{opt}} (t) dt}.
\end{align}
Identifying \eqref{a_1T} and \eqref{a_1T_} leads to
\begin{align}
\Im{(\omega_{\lambda'})}&=\frac{1}{T'}\left(\int_0^{T'}\Im{(\Delta_2(t'))}dt'+\int_0^T \Im(\Delta_1(t'))dt'\right)
\nonumber\\
&-\frac{1}{T'}\log(|a_1(T)|) -\frac{1}{\hbar T'} \int_T^{T'} V_{\text{opt}} (t') dt',
\end{align}
which gives a relation between the imaginary parts of the two eigenvalues:
\begin{align}
\label{Elambdap}
\Im{(\omega_{\lambda'})}&=-\frac{1}{\hbar T'} \int_T^{T'} V_{\text{opt}} (t') dt'
-\Im{(\omega_{\lambda})}\nonumber\\
&+ \frac{1}{T'}\int_0^{T'} \Im(\Delta_1(t')+\Delta_2(t'))dt'.
\end{align}
This central relation shows that the connected eigenvalue will be in general well isolated from the other one for a large enough area of the absorbing potential. More precisely we have $-\Im{(\omega_{\lambda'})}\gg -\Im{(\omega_{\lambda})}$ when
\begin{align}
\label{cond_isol}
&\frac{1}{\hbar T'} \int_T^{T'} V_{\text{opt}} (t') dt'\gg
-2\Im{(\omega_{\lambda})}\nonumber\\
&\qquad\qquad+\frac{1}{T'}\int_0^{T'} \Im(\Delta_1(t')+\Delta_2(t'))dt'
\end{align}
This feature will be useful in numerical calculations; in particular it will improve the rate of convergence of
the wave operator method \cite{reviewgeorges2} when applied to the location of the thus isolated connected eigenvalue.
However the separation between the imaginary parts of eigenvalues can be not so efficient in practice for specific cases of good population transfer. This is analyzed in the following section.
\section{Numerical limitations and accuracy \label{limitations}}
In this section we study the numerical limitations of the method, restricting the discussion to the situation $c_1(0)=1,c_2(0)=0$.
We consider for simplicity the situation $\Im{(\Delta_2(t))}=\Im{(\Delta_1(t))}=0$.
\subsection{General cases}
The accuracy of the method can be roughly estimated from the imperfect initial connection with the eigenvector $\ket{\lambda}$, that is from the small quantity $\lambda_2(0)$. In general, when $\lambda_2(T)$ and $\lambda_1(T)$ are of the same order, we obtain for the error in the final amplitude from \eqref{ratio21}:
\begin{equation}
\label{acc2}
\left|a_1(T)-a_1^{\text{(CATM)}}(T)\right|\propto e^{-\frac{1}{\hbar} \int_T^{T'} V_{\text{opt}} (t') dt'},
\end{equation}
where $a_1^{\text{(CATM)}}(T)$ is the probability amplitude of state 1 at the end of the physical process obtained from the CATM method.
This is shown to give a correct estimation of the accuracy of the method when it is tested numerically (see next section).
We remark that this estimation does not obviously take into account the grid size effect. This is studied numerically in the next section.
\subsection{Case of good population transfer}
The estimation \eqref{acc2} is not valid when the population transfer at the end of the process is efficient: $|a_1(T)|\to0$, since, in Eq. \eqref{ratio21}, we have then $|\lambda_2(T)/\lambda_1(T)|\gg1$.
The area of the optical potential should be large enough to satisfy the connectivity to a unique Floquet eigenvector: $\lambda_2(0)/\lambda_1(0)\leadsto0$, that is, from \eqref{cond_2_}
\begin{equation}
\label{cond_2__}
\frac{1}{\hbar}\int_{T}^{T'} V_{\text{opt}}(t) dt \gg -\log(a_1(T)).
\end{equation}
One limiting case is when there is no separation between the imaginary parts of the eigenvalues:
\begin{equation}
\label{equalIm}
\Im{(\omega_{\lambda'})}=\Im{(\omega_{\lambda})},
\end{equation}
leading to
\begin{align}
&\frac{1}{\hbar}\int_T^{T'} V_{\text{opt}} (t') dt'=
-2\log(|a_1^{\text{(CATM)}}(T)|)).
\end{align}
This equation shows that, in this case of equal quasienergies, the inequality \eqref{cond_2__} is satisfied with only a factor 2. More precisely, we have
\begin{equation}
\frac{\lambda_2(0)}{\lambda_1(0)}\approx e^{-\frac{1}{2\hbar}\int_{T}^{T'} V_{\text{opt}}(t) dt}.
\end{equation}
Thus, one can still satisfy $\lambda_2(0)/\lambda_1(0)\leadsto0$ to get the connection to a unique Floquet eigenvector to a good accuracy by imposing
\begin{equation}
\label{cond_2___}
\frac{1}{2\hbar}\int_{T}^{T'} V_{\text{opt}}(t) dt \gg 1.
\end{equation}
This condition \eqref{cond_2___}, a bit more restrictive than \eqref{cond_1} is thus sufficient to obtain a quite good relative accuracy of the solution in the case of good population transfer, even if in that case the imaginary parts of the Floquet eigenvalues are close together.
We can use this limiting case \eqref{equalIm} to estimate the absolute accuracy of the method.
Assuming $\Im{(\omega_{\lambda'})}\le\Im{(\omega_{\lambda})}$, we get
\begin{equation}
\label{acc}
|a_1^{\text{(CATM)}}(T)|\ge e^{-\frac{1}{2\hbar} \int_T^{T'} V_{\text{opt}} (t') dt'},
\end{equation}
that is we cannot obtain numerically a population $|a_1^{\text{(CATM)}}(T)|^2$ of state 1 at the end of the physical process smaller than $e^{-\frac{1}{\hbar} \int_T^{T'} V_{\text{opt}} (t') dt'}$,
which gives thus a numerical limitation of the depopulation of the initial state.
\section{Numerical investigation \label{inspection}}
The method is investigated numerically in this section through the examples of two- and three- state systems driven by a time-dependent field. They can correspond for instance to atoms submitted to resonant laser pulses in the rotating wave approximation (RWA) \cite{shore,guerin}.
\subsection{Some results for selected examples \label{exemples}}
The first example is a two-state system $\{\ket{1},\ket{2}\}$ which is subjected to a pulsed coupling of frequency little detuned with the transition frequency. The detuning is denoted $\Delta$ and $\Omega$ is the coupling (Rabi frequency). In the dressed state picture of the RWA the Hamiltonian is (in units such that $\hbar=1$)
\begin{equation}
H = \left(
\begin{array}{ll}
0 & \Omega \\
\Omega & \Delta
\end{array} \right)
= \left(
\begin{array}{ll}
0 & \Omega_0 \sin ^2 \left( \frac{\pi t}{T} \right) \\
\Omega_0 \sin ^2 \left( \frac{\pi t}{T} \right) & \Delta_0 \cos \left( \frac{\pi t}{T} +\phi_0 \right)
\end{array} \right)
\end{equation}
We will consider as initial condition (i) $\ket{\Psi (t=0)} = \ket{1}$,
from which we expect a final quasi-inversion of population for large enough $\Omega_0T$ and $\Delta_0T$ (adiabatic passage, see \cite{shore,guerin}), and (ii) the more complicated situation $\ket{\Psi (0)} = c_1 \ket{1} + c_2 \ket{2}$.
The second example is that of a 3 level system $\{\ket{1} , \ket{2},\ket{3}\}$ driven by two near-resonant laser fields with Rabi frequencies
$\Omega_p$ and $\Omega_s$, tuned to the transitions $1\leftrightarrow 2$ and $2\leftrightarrow 3$ respectively.
We allow a detuning $\Delta$ between the transition frequency $1\rightarrow2$ and the laser frequency and assume a two-photon resonance. The initial state is $\ket{1}$.
Here the RWA Hamiltonian takes the form :
\begin{equation}
H = \left(
\begin{array}{lll}
0 & \Omega_p & 0 \\
\Omega_p & \Delta & \Omega_s \\
0 & \Omega_s & 0
\end{array} \right)
\end{equation}
We study two situations, on one hand the intuitive case: we first turn on the coupling between levels $1$ and $2$, then
between levels $2$ and $3$,
\begin{eqnarray}
\Omega_p & =& \Omega_0 \sin ^2 \left( \frac{\pi t}{T_1} \right) \quad \forall t \in [0,T_1] \quad (0 \text{ elsewhere}) \nonumber\\
\Omega_s &=& \Omega_0 \sin ^2 \left( \frac{\pi t-T_1/2}{T_1} \right) \quad \forall t \in \left[\frac{1}{2}T_1,\frac{3}{2}T_1 \right] \nonumber\\
\Delta&=&\Delta_0
\end{eqnarray}
With $\Omega_0 T_1= 20$ and $\Delta_0 T_1=0$,
we expect to observe oscillations without complete population exchange to state $\ket{3}$.
With $\Delta_0 T_1=20$ a partial transfer to $\ket{3}$ occurs with less oscillations.
On the other hand the STIRAP case (Stimulated Raman Adiabatic Passage) is exactly the inverse of the first configuration \cite{stirap}:
\begin{eqnarray}
\Omega_p & =& \Omega_0 \sin ^2\left( \frac{\pi t-T_1/2}{T_1} \right) \quad \forall t \in \left[\frac{1}{2}T_1,\frac{3}{2}T_1\right] \nonumber \\
\Omega_s &=& \Omega_0 \sin ^2 \left( \frac{\pi t}{T_1} \right) \quad \forall t \in [0,T_1] \nonumber\\
\Delta & = &0
\end{eqnarray}
With $\Omega_0 T_1= 20$ and $\Delta_0 T_1=0$, STIRAP allows a large transfer of the population to $\ket{3}$.
The total physical time interval $T$ here is $3/2$ times the period $T_1$ of the sine function $[0,T]=[0,3/2T_1]$;
the additional time interval will begin at $3/2T_1$ for a duration of $T_1$.
In the subsequent discussion we use the labels (i) and (ii) for the 2 level system with initial state $\ket{1}$
and the superposition of states, respectively. The label (iii) and (iv) refer to the 3 level system in the ``intuitive'' or STIRAP situations, respectively.
\subsection{Calculating with CATM}
From a technical point of view the calculation involves the five following steps:
\begin{itemize}
\item Construction of the matrix representation of the Floquet Hamiltonian (some details are given in appendix \ref{appB})
\item Diagonalization of the Floquet matrix
\item Selection of $N$ Floquet eigenstates belonging to the first Brillouin zone (for a problem with $N$ levels)
\item Detection of the appropriate ``connected'' Floquet eigenstate, i.e. corresponding to the smallest imaginary part of the eigenvalue in absolute value as a criterion
\item Production of the wavefunction via Eq. (\ref{wavefunction})
\end{itemize}
In principle only one vector computation is needed. For our small-scale examples we can easily use direct complete diagonalisation. However, for larger systems the time-dependent wave operator can be used to find the required eigenstate of the corresponding large matrix.
\subsection{A comparison with direct integration}
We analyse the results obtained with the Floquet eigenstate which possesses the smallest value
of $\vert \Im (\omega_{\lambda}) \vert$, as predicted by the theory. Next we calculate the populations
\begin{equation}
p_n(t)=\vert \langle n \ket{\Psi (t)} \vert ^2
\end{equation}
and the relative phases
\begin{equation}
\beta_n(t)=\arg \left( \langle n \ket{\Psi (t)} \right)
\end{equation}
for all the previously presented situations.
We compare the CATM results for the population and phase with those of a direct integration
using the propagation equation
\begin{equation}
\label{Direct}
\ket{ \Psi (t+\Delta t)} = e^{ -i \hbar^{-1} H \left(t+\frac{\Delta t}{2}\right) \Delta t} \ket{ \Psi (t)}
\end{equation}
with $\Delta t$ a sufficiently small time-step.
For the CATM calculation, the size of the Fourier basis set was $N=256$ which is ample for both stable computation and
graphical representation.
\subsubsection{Two-state model}
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig1_CATM_ss.eps}
\caption{Evolution of populations (a) and phases (b) for the 2-level system (i) with the initial state $\ket{1}$, and $\Omega_0 T = 10$ and $\Delta_0 T= 10$.
Exact [i.e. numerical with the direct integration, cf. \eqref{Direct}] results ($p_1$ and $\beta_1$: short dashes, $p_2$ and $\beta_2$: dots)
and CATM results ($p_1$ and $\beta_1$: solid line, $p_2$ and $\beta_2$: long dashes)
for various amplitudes $V_0$ of the time-dependent absorbing potential: (I) $V_0 T=0$ (II) $V_0 T=10$ (III) $V_0 T=40$.}
\label{fig1}
\end{figure}
For the 2 level system (i) the results are shown on Fig. \ref{fig1}.
In frames (a-I) and (b-I), it is evident that without the absorbing potential the use of a single Floquet state is not sufficient. On frames (a-II) and (b-II) we can
observe the effects of the absorbing potential.
The initial populations approach $p_1(0)=1$ and $p_2(0)=0$, showing however a small difference
of a few percent from the reference calculation results. Phases begin to agree with those of the reference calculation
but the difference remains important, especially at the beginning.
For the last case (a-III and b-III), one cannot detect any difference between
the CATM and the reference results at the scale of the figure.
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig2_CATM_ss.eps}
\caption{Same as Fig. \ref{fig1}, but for the 2-level system (ii), and $\Omega T = 10$, $\Delta_0 T =0$ with the initial state $\sqrt{0.75}\,\ket{1}+\sqrt{0.25}\,\ket{2}$,
for various amplitudes $V_0$ of the time-dependent absorbing potential: (I) $V_0 T=0$ (II) $V_0 T=10$ (III) $V_0 T=40$.}
\label{fig2}
\end{figure}
Fig. \ref{fig2} shows the same quantities
for the initial condition $\ket{\Psi(0)} = c_1 \ket{1} + c_2 \ket{2}$, $c_1=\sqrt{0.75}$ and $c_2=\sqrt{0.25}$. We have used the absorbing potential
given by Eq.(\ref{potgen}). The previous comments about the efficiency of the method remain valid.
Fig. \ref{fig2} illustrates clearly the efficiency of the chosen matrix in reproducing
the boundary conditions.
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig3_CATM_ss.ps}
\caption{Integrated logarithmic error estimation between direct integration and CATM with 512 time grid points as a function of $V_0$,
calculated with the first component $\langle 1 \ket{\Psi}$ for the 2-level system (i).
Errors on population, see Eq. \eqref{epsp} (solid line), and on angles, see Eq. \eqref{epsa} (dashed line). We remark that these errors follow the anticipated exponential law Eq. \eqref{acc2} (dotted line) until they reach a plateau due to grid effects of CATM}
\label{fig6}
\end{figure}
We now give a more precise analysis of how the exact solution is approached.
To this end we define a measure of the difference between the CATM results and the direct integration results.
For the single component $\langle 1 \ket{\Psi}$ calculated by the two methods we define the integrated difference of population and
of angle:
\begin{subequations}
\begin{eqnarray}
\label{epsp}
\epsilon_{p} &=& \frac{1}{T}\int_0^T \left(\vert \langle 1
\ket{\Psi(t)}_{\text{CATM}} \vert ^2 - \vert \langle 1 \ket{\Psi(t)} \vert ^2\right) dt \\
\label{epsa} \epsilon_{a}& =& \frac{1}{T}\int_0^T \left[ \arg\left( \langle 1 \ket{\Psi(t)}_{\text{CATM}} \right) -
\arg \left( \langle 1 \ket{\Psi(t)} \right) \right] dt\qquad\
\end{eqnarray}
\end{subequations}
These quantities are represented on Fig. \ref{fig6} as a function of the absorbing potential amplitude $V_0$.
With the logarithmic scale, we observe a quasi-linear law for $V_0\in[10,35]$ in consistency with Eq. \eqref{acc2}.
The error estimates next reach plateaus which are interpreted by the grid effects due to the finite basis representation of the time in the CATM method.
Indeed we can lower the level of the plateaus by increasing the number of the grid points (not shown).
\subsubsection{Three-state model}
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig4_CATM_ss.eps}
\caption{Evolution of the populations $p_n(t)$ the three-level model (iii); exact (numerical) results $p_1$ (dots), $p_2$ (long dot-dashes), $p_3$ (dot-dashes)
and CATM results $p_1$ (solid line), $p_2$ (long dashes), $p_3$ (short dashes) without detuning and for various amplitude of the absorbing potential,
I: $V_0 T_1=0$, II: $5$, III: $10$, IV: $40$. }
\label{fig3}
\end{figure}
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig5_CATM_ss.eps}
\caption{Same as Fig. \ref{fig3}, but with a detuning $\Delta_0T_1=20$.}
\label{fig4}
\end{figure}
For the 3 level system the evolution of the population in the three-level model (iii) (as defined in section \ref{exemples})
is shown in Fig. \ref{fig3} and Fig. \ref{fig4}, without or with detuning ($\Delta_0 T_1=0$ or $\Delta_0T_1 =20$).
The selected field amplitude was $\Omega_0 T_1 = 20$ and the absorbing potential was gradually turned on from $V_0T_1=0$ to $V_0T_1=40$.
Here again, if the absorption is not sufficient, the results are wrong.
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig6_CATM_ss.eps}
\caption{Evolution of the populations $p_n(t)=\vert \langle n \ket{\Psi (t)} \vert ^2$ in the STIRAP model (iv) ; exact (numerical) results $p_1$ (dots), $p_2$ (long dot-dashes), $p_3$ (short dot-dashes)
and CATM results $p_1$ (solid line), $p_2$ (long dashes), $p_3$ (short dashes) without detuning and for various amplitude of the absorbing potential,
I : $V_0 T_1=0$, II : $5$, III : $10$, IV : $40$. }
\label{fig5}
\end{figure}
The results for the STIRAP model (iv) (as defined in section \ref{exemples}) are displayed in Fig \ref{fig5}.
The coupling terms between levels $2$ and $3$ are turned on before the coupling terms between $1$ and $2$ and a relatively large population inversion is observed.
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig7_CATM_ss.eps}
\caption{(a) $\Re(\omega_{\lambda})$, (b) $\Im(\omega_{\lambda})$,
(c) $\Re(\omega_{\lambda'})$ and (d) $\Im(\omega_{\lambda'})$, versus $V_0$ (all in units of $1/T_1$) for the two-level model (i). When $V_0$ grows, $\omega_{\lambda'}$ moves away from $\omega_{\lambda}$ acquiring a imaginary part proportional to $V_0$.}
\label{eig1}
\end{figure}
\subsection{The expansion of the spectrum}
We now analyse the effect of dilatation of the eigenvalues by the the absorbing potential, that is the feature of separating the ``connected'' eigenvalue with respect to the other ones.
Fig. \ref{eig1} shows the Floquet eigenvalues $\{\omega_{\lambda}\}$ and $\{\omega_{\lambda'}\}$
in the first Brillouin zone calculated for the two-level model (i) as functions of $V_0$ .
Apart for small absorbing potential amplitudes where one notices an ambiguity concerning
the labelling of eigenvalue \cite{viennot}, $\Im(\omega_{\lambda})$ is a constant value in agreement with Eq. \eqref{Elambda},
and $\Im(\omega_{\lambda'})$ shows a linear evolution as predicted by
Eq. \eqref{Elambdap}.
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig8_CATM_ss.ps}
\caption{$\Im(\omega_{\lambda_1})$ versus $V_0$ for the three-level model (iii). }
\label{eig2}
\end{figure}
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig9_CATM_ss.ps}
\caption{$\Im(\omega_{\lambda_2})$ (solid line) and $\Im(\omega_{\lambda_3})$ (dashed line) versus $V_0$ for the three-level model (iii).}
\label{eig3}
\end{figure}
Figures \ref{eig2} and \ref{eig3} refer to the three-level system (iii) and show
the same features.
Concentrating on the imaginary part of the ``connected'' Floquet eigenvalue (Fig. \ref{eig2})
we see that, after a region of stabilization, $\Im(\omega_{\lambda_1})$ is no longer affected by the growth of the absorbing potential.
By contrast both $\Im(\omega_{\lambda_2})$ and $\Im(\omega_{\lambda_3})$ acquire imaginary parts which are linear with respect to $V_0$.
This feature will be useful in practice for large systems, in particular
if a wave operator method is used to find the Floquet eigenstate \cite{reviewgeorges2},
since that method is efficient in finding isolated eigenvalues.
\subsection{The influence of the number of Fourier basis functions}
We give here some details about the stability of the results
as the number of Fourier basis functions $N$ is reduced in the CATM calculation.
To increase calculational speed and to decrease memory requirement it appears necessary
to use as small value of $N$ as possible.
Figure~\ref{stability} shows how the final populations obtained in the CATM calculation
varies as $N$ is increased. These calculations correspond to the STIRAP model (iv).
\begin{figure}[!ht]
\includegraphics[width=\linewidth]{fig10_CATM_ss.ps}
\caption{Stability of final populations $p_1$ (solid line), $p_2$ (dashed line) and $p_3$ (dotted line) functions of the Fourier time grid point number $N$,
in logarithmic scale.}
\label{stability}
\end{figure}
The values $p_3 \simeq 0.82$ and $p_1 \simeq 0.18$ are stable for $N \gtrsim 30$ but $p_2\simeq 10^{-3}$ is not obtained
accurately until $N$ is about $80$. We see that finding small probabilities in absolute values requires a more precise description of the temporal evolution;
however about 80 grid points appear to be ample for the calculations. The general principle is to choose an $N$ which is high enough to
follow the time variations in the Hamiltonian and to obtain accurate values for small probabilities.
\section{conclusion}
The optimum computational implementation of the CATM is still being actively researched but the basic principles behind the method are simple to follow. A static absorbing potential is often used in treating the time development of a wavefunction within the Floquet formalism. The novelty of our approach is that the absorbing potential is given a time-dependent form such that it actively constrains the wavefunction, both by imposing the correct boundary conditions on it and by modifying the spectrum so that the specific eigenvalue which is appropriate to describe the dynamical process is rendered relatively isolated from the other eigenvalues. The dynamical problem is then rendered into an eigenvalue problem in which the isolated eigenvalue is easier to find by techniques such as the Bloch wave operator method. That it is indeed possible to choose the time-dependent potential so as to produce the favourable features described above has been demonstrated for two small-scale systems for which accurate comparison results are available. For these small test systems the CATM gives accurate results, although it is clear that the eigenvalue problems which arise can involve strongly non-Hermitian matrices.
The CATM has some formal advantages for systems with a time-dependent Hamiltonian. A common approach for such systems is to use a step-by step propagation procedure with very small time steps. Many time steps are thus required to cover a given time interval and this leads to an accumulation of errors as the propagation proceeds. By contrast, in the CATM the solution is global over the full time interval and so there is no accumulation of errors; this feature is similar to that shown by the $(t,t')$ method \cite{peskin}.
As expected (and confirmed by the present study) the accuracy achievable within the CATM is governed by the ability to reproduce the initial conditions by suitably adjusting the time-dependent potential and by the use of a sufficiently dense Fourier time grid to describe any fast time variations contained in the Hamiltonian.
Our model calculations have also made clear the role of the time-dependent absorbing potential in dilating the Floquet spectrum so that the dynamical problem of propagation within a Hilbert space of a given dimension can be converted to that of locating an isolated eigenvalue of a non-Hermitian matrix of much larger dimension.
The difficulty of solving the dynamical problem is thus converted into the technical problem of devising efficient algorithms for large non-Hermitian matrices. In Ref. \cite{CATM}
a previous version of the CATM was successfully tested on a molecular system involving a few hundred states. At the moment we believe that the task of isolating and then calculating the important dynamically relevant complex eigenvalue is probably not possible within the CATM for systems which are much larger than those treated in \cite{CATM}; nevertheless, the method may be useful for some systems which cause difficulties for the usual propagation methods.
\begin{acknowledgments}
We acknowledge the support from the French Agence Nationale de la
Recherche (Project CoMoC), the European Marie Curie Initial Training
Network Grant No. CA-ITN-214962-FASTQUAST, and from the Conseil
R\'egional de Bourgogne.
\end{acknowledgments}
|
\section{Introduction.}
\setcounter{section}{0}
In this paper, curves are always
assumed to be regular (i.e. immersed).
The well-known Fabricius-Bjerre \cite{FB}
theorem asserts (see also \cite{H1})
that
\begin{equation}\label{eq:FB}
d_1(\gamma)-d_2(\gamma)=\#_\gamma+\frac{i_\gamma}2
\end{equation}
holds for closed curves $\gamma$ satisfying
suitable genericity assumptions,
where $d_1(\gamma)$ (resp. $d_2(\gamma)$) is
the number of double tangents
of same side (resp. opposite side) and
$\#_\gamma$ and $i_\gamma$ are the number of crossings
and the number of inflections on $\gamma$, respectively.
\begin{figure}[htb]
\begin{center}
\includegraphics[height=5.2cm]{newfive.eps}
\caption{A curve with $\#_\gamma=2$, $i_\gamma=2$, $d_1=4$ and $d_2=1$.}
\label{fig:fb}
\end{center}
\vspace{-2ex}
\end{figure}
However, for arbitrarily given
non-negative integers $d_1,d_2,n$ and $i$
satisfying $d_1-d_2=n+i/2$, there might not exist
a corresponding curve, in general.
As an affirmative answer to the Halpern conjecture in \cite{H2},
the second author \cite{O} proved the inequality
\begin{equation}\label{eq:Ozawa}
d_1(\gamma)+d_2(\gamma)\le \#_\gamma(2\#_\gamma-1)
\end{equation}
for closed curves without inflections
(see Remark \ref{rmk:halpern}).
It is then natural to expect that there might
be further obstructions for
the topology of closed planar curves without
inflections.
In this paper, we define a computable topological invariant
$\mu(\gamma)$ for closed planar curves.
By applying the Gauss-Bonnet formula, we show the
inequality $i_\gamma\ge \mu(\gamma)$,
which is sharp at least for closed curves
satisfying $\#_\gamma\le 4$. In fact, for such $\gamma$,
there exists a closed curve $\sigma$ which
has the same topological type
as $\gamma$ such that $i_\sigma= \mu(\sigma)$.
As an application, we classify the topological
types of closed planar curves
satisfying $i_\gamma=0$ and $\#_\gamma\le 5$.
Moreover, we discuss the relationship between
the number of double tangents of the curve and the invariant $I(\gamma)$
on a given $\gamma$.
\section{Preliminaries and main results}
We denote by $\R^2$ the affine plane, and by $S^2$
the unit sphere in $\R^3$.
A closed curve $\gamma$ in $\R^2$ or $S^2$
is called {\it generic} if
\begin{enumerate}
\item all crossings are transversal, and
\item the zeroes of curvature are nondegenerate.
\end{enumerate}
By stereographic projection,
we can recognize $S^2=\R^2\cup \{\infty\}$.
Two generic closed curves $\gamma_1$ and $\gamma_2$
in $\R^2$ (in $S^2$) are called {\it geotopic},
or said to have the {\it same topological type}, in $\R^2$
(resp. in $S^2$)
if there is an orientation
preserving diffeomorphism $\phi$ of $\R^2$ (resp. on $S^2$)
such that
$
\op{Im}\gamma_2=\phi(\op{Im}\gamma_1).
$
This induces an equivalence relation
on the set of closed curves.
We denote equivalence classes by
$\langle \gamma_1 \rangle$ (resp. $[\gamma_1]$).
We fix a closed regular curve $\gamma:S^1\to \R^2$.
A point $c\in S^1$ is called an {\it inflection point}
of $\gamma$ if $\op{det}(\dot\gamma(t),\ddot\gamma(t))$ vanishes
at $t=c$.
We denote by $i_\gamma$ the number of
inflection points on $\gamma$.
A closed curve $\gamma:[0,1]\to \R^2$
is called {\it locally convex}
if $i_\gamma=0$.
Whitney \cite{W} proved that
any two closed curves
are regularly homotopic if and only
if their rotation indices coincide.
So then one can ask if this regular homotopy
preserves the locally convexity when the given
two curves are both locally convex.
In fact, one can easily prove this
by a modification of Whitney's argument,
which has been pointed out in \cite[p35, Exercise 11]{O2}.
For the sake of the readers' convenience we shall
outline the proof:
\begin{proposition}
Let $\gamma_1$ and $\gamma_2$ be two
locally convex closed regular curves.
Suppose that $\gamma_1$ has the same rotation
index as $\gamma_2$.
Then there exists a family of closed
curves
$\{\Gamma_\epsilon\}_{\epsilon\in [0,1]}$
such that
\begin{enumerate}
\item $\Gamma_0=\gamma_1$ and $\Gamma_1=\gamma_2$,
\item each $\Gamma_\epsilon$ $(\epsilon\in [0,1])$ is a locally convex
closed regular curve.
\end{enumerate}
\end{proposition}
\begin{proof}
Suppose that the curves $\gamma_j$ $(j=1,2)$
are both positively
curved and have the same rotation indices, equal to $m$.
We change, if necessary, the parametrizations of $\gamma_j$
so that the tangent vectors $\dot\gamma_j(t)$ are
positive scalar multiples of $\pmt{\cos t\\ \sin t}$.
Define a homotopy $\Gamma_\epsilon$ between $\gamma_1$ and
$\gamma_2$ by
\[
\Gamma_\epsilon(t)
= (1-\epsilon)\gamma_1(t) + \epsilon\gamma_2(t).
\]
It is easy to verify that
$\{\Gamma_\epsilon\}_{\epsilon\in[0,1]}$ satisfies
the required conditions (1) and (2).
\end{proof}
For a given generic closed curve $\gamma$,
we set
\begin{equation}
I(\gamma):=\min_{\sigma\in \langle \gamma\rangle}i_\sigma.
\end{equation}
Since $i_\gamma$ is an even number, so is $I(\gamma)$.
Inflection points on curves
in $\R^2$ correspond to
singular points on their Gauss maps.
So it is natural to ask about the existence of
topological restrictions on closed curves
without inflections, in other words,
we are interested in the topological
type of closed curves $\gamma$ satisfying
$I(\gamma)=0$.
There are explicit combinatorial
procedures for determining generic
closed spherical curves with a
given number of crossings,
as in Carter \cite{S} and Cairns and Elton \cite{CE}
(see also Arnold \cite{A}).
We denote by $\#_\gamma$ the
number of crossings for a given generic
closed curve. In this paper, we use the table
of closed spherical curves
with $\#_\gamma\le 5$
given in the appendix of \cite{KU}.
For example, the table of closed spherical curves
with $\#_\gamma\le 2$ is given in Figure \ref{fig:curve2s},
where $1_2$ (resp. $2_2$)
means that the corresponding curve has 2-crossings and
appears in the table of curves in \cite{KU} with $\#_\gamma=2$ primary
(resp. secondary).
\begin{figure}[htb]
\begin{center}
\includegraphics[height=1.3cm]{curve2s.eps}
\caption{Spherical closed curves with $\#_\gamma\le 2$.}
\label{fig:curve2s}
\end{center}
\vspace{-2ex}
\end{figure}
Moving the position of $\infty$
via motions in $S^2=\R^2\cup \{\infty\}$,
we get the table of
closed planar curves with $\#_\gamma \le 2$
as in Figure \ref{fig:curve2a}.
For example $1_1$ and $1_1^b$
(resp. $2_2$, $2_2^b$ and $2_2^c$) are
equivalent to $[1_1]$ (resp. $[2_2]$)
as spherical curves. Here, only the curves of
type $1_1$ and $2_2$
can be drawn with no inflections.
Similarly, using the table of spherical
curves with $\#_\gamma\le 5$ given in \cite{KU},
we prove the following theorem.
The authors do not know of any reference for
such a classification of generic locally convex curves.
\begin{figure}[htb]
\begin{center}
\includegraphics[height=2.8cm]{curve2aa.eps}
\caption{Planar closed curves with $\#_{\gamma}\le 2$.}
\label{fig:curve2a}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{theorem}\label{thm:main}
For a given generic closed regular curve $\gamma$
in $\R^2$, the inequality
\begin{equation}\label{eq:affine}
I(\gamma)\ge \mu(\gamma)
\end{equation}
holds. Moreover,
$I(\gamma)=0$ if and only if $\mu(\gamma)=0$ under the assumption
that $\#_\gamma\le 5$.
\end{theorem}
\medskip
In particular, the number of equivalence classes of
closed locally convex curves with $\#_\gamma\le 5$
is $76$ (see Figure \ref{fig:curve2a} and Figures
\ref{fig:3}, \ref{fig:4} and \ref{fig:5} in Section 3).
For example, in Figure \ref{fig:curve2a}
the curves of type $1_0,1_1,1_2,2_2$ satisfy $I(\gamma)=0$,
and the remaining $1_1^b,1_2^b,2_2^b,2_2^c$ satisfy
$I(\gamma)=2$. In Figure \ref{fig:3}, the curve of type $6_3^a$
is of the same topological type as $6_3^b$ as a spherical
curve, which is obtained from the 6th curve in
the table of curves with
$\#_\gamma=2$ in the appendix of \cite{KU}.
\begin{corollary}\label{cor:main}
A generic closed curve $\gamma$
with $\#_\gamma\le 4$ satisfies
$I(\gamma)= \mu(\gamma)\le 2$,
unless the topological type of
$\gamma$ is as in Figure \ref{fig:i4}.
\end{corollary}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=1.2cm]{i4.eps}
\caption{A curve with $\#_\gamma=4$ and
$I(\gamma) =\mu(\gamma) =4$.}
\label{fig:i4}
\end{center}
\vspace{-2ex}
\end{figure}
\noindent
A table of closed curves $\gamma$ with $i_\gamma=\mu(\gamma)$
for $\#_\gamma \le 3$ is given in
Figures \ref{fig:curve2a}, \ref{fig:3} and \ref{fig:3g}.
For example, in Figure \ref{fig:3g},
the curves of type $6_3^c$
consideres as spherical curves
are of the same topological type as the curves
of type $6_3^a$ or $6_3^b$
in Figure \ref{fig:3}.
\section{Definition of the invariant $\mu(\gamma)$}
We fix a generic closed curve $\gamma:S^1\to \R^2$.
We set $\#_\gamma=m$.
We may suppose that $\gamma(0)=\gamma(1)$
is one of the crossings of $\gamma$.
Let
$$
0=c_1<\cdots < c_{2m}(<1)
$$
be the inverse image of the crossings of
$\gamma$, which consists of $2m$ points in $S^1=\R/\Z$.
We set
$$
S^1_{\gamma}:=S^1\setminus \{c_1,...,c_{2m}\}.
$$
To introduce the invariant $\mu(\gamma)$,
we define special subsets on the curves called
{\lq $n$-gons\rq}:
\begin{figure}[htb]
\begin{center}
\includegraphics[height=1.3cm]{shells.eps}
\caption{Shells.}
\label{fig:shells}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{definition}
Let $n(\ge 3)$ be an integer.
A disjoint union of $n$
proper closed intervals
$$
J:=[a_1,b_1]\cup \cdots \cup [a_n,b_n]
$$
on $S^1$ is called an {\it $n$-gon}
if $a_1,b_1,\cdots,a_n,b_n\in \{c_1,...c_{2m}\}$
and the image $\gamma(J)$ is
a piecewise smooth simple closed curve in $\R^2$.
The simply connected domain
bounded by $\gamma(J)$
is called the {\it interior domain} of the $n$-gon.
An $n$-gon is called {\it admissible} if
at most two of the $n$ interior
angles of $D$ are less than $\pi$.
\end{definition}
We denote by $\mathcal G_n(\gamma)$ the set of all
admissible $n$-gons, and set
$$
\mathcal G(\gamma):=\bigcup_{n=1}^\infty \mathcal G_n(\gamma).
$$
Each element of $\mathcal G(\gamma)$ is called
an {\it admissible polygon}.
A $1$-gon is called a {\it shell}
(cf. Figure \ref{fig:shells}).
A $2$-gon is called a {\it leaf}
(cf. Figure \ref{fig:leaves})
and a $3$-gon is called
a {\it triangle} (cf. Figure \ref{fig:triangles}).
All shells and all leaves are admissible.
However, a triangle whose interior angles are all
acute is not admissible.
\begin{figure}[htb]
\begin{center}
\includegraphics[height=1.3cm]{leaves.eps}
\caption{Leaves.}
\label{fig:leaves}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=2cm]{triangles.eps}
\caption{Admissible triangles.}
\label{fig:triangles}
\end{center}
\vspace{-2ex}
\end{figure}
We fix an admissible $n$-gon
$
J:=[a_1,b_1]\cup \cdots \cup [a_n,b_n].
$
Then $\gamma(J)$ is a piecewise smooth simple closed curve
in $\R^2$. We give an orientation of $\gamma(J)$ so that
the interior domain of $\gamma(J)$ is on the
left hand side of $\gamma(J)$.
This orientation induces an orientation on $[a_i,b_i]$ for each
$i=1,...,n$.
We call $[a_i,b_i]$ a {\it positive interval}
(resp. {\it negative interval})
if the orientation of the interval $[a_i,b_i]$
coincides with (resp. does not coincide with) the orientation of
$\gamma$.
Let $t$ be a point on $J\setminus\{a_1,b_1,...,a_n,b_n\}
=\bigcup_{i=1}^n (a_i,b_i)$.
Then $t$ belongs to an open interval $(a_i,b_i)$ for some
$i=1,...,n$. Then we set
$$
\epsilon_J(t):=
\begin{cases}
1 & \mbox{if $[a_i,b_i]$ is positive}, \\
-1 & \mbox{if $[a_i,b_i]$ is negative}.
\end{cases}
$$
\begin{definition}
An admissible $n$-gon is called {\it positive}
(resp. {\it negative})
if $\epsilon_J(t)>0$ (resp. $\epsilon_J(t)<0$)
holds for each
$t\in J\setminus\{a_1,b_1,...,a_n,b_n\}$.
\end{definition}
The notions of positivity and negativity of shells were used
in \cite{U} and \cite{KU} differently from here.
The following assertion is the key to proving the
inequality \eqref{eq:affine}:
\begin{lemma} \label{lem:key}\
Let
$
J:=[a_1,b_1]\cup \cdots \cup [a_n,b_n]
$
be an admissible $n$-gon.
Then there exists an interval $[a_i,b_i]$
{\rm ($1\le i\le n$)}
and a point $c\in (a_i,b_i)$ such that
$\op{sgn}(\kappa_\gamma(c))=\epsilon_J(c)$, where
$
\kappa_\gamma(t):={\op{det}(\dot\gamma(t),\ddot \gamma(t))}/
{|\dot\gamma|^3}
$
is the curvature of $\gamma$
and $\op{sgn}(\kappa_\gamma(c))$ is the sign of the
real number $\kappa_\gamma(c)$.
\end{lemma}
\begin{proof}
Let $A_1,A_2,\cdots, A_n$ be the interior angles of
the interior domain
of $J$, and set
$\Gamma=\gamma(J)$, which we regard as
an oriented piecewise smooth simple closed curve
with counterclockwise orientation.
In this proof,
$\R^2$ is considered as the Euclidean plane,
and we take the arclength parameter $s$ of $\Gamma$.
Let $s=s_1,...,s_{n}$ be the points where $d\Gamma/ds$
is discontinuous.
We denote by $\kappa_{\Gamma}(s)$
($s\ne s_1,...,s_{n}$)
the curvature of the curve $\Gamma$.
Then, the Gauss-Bonnet formula yields that
$$
\int_{J}\kappa_{\Gamma}(s) ds=-(n-2)\pi+\sum_{i=1}^nA_i.
$$
Since $J$ is admissible, we may assume that $A_1,...,A_{n-2}\ge \pi$,
and so $\int_{J}\kappa_\Gamma(s) ds>0$.
Then there exist an index $i$ $(1\le i \le n)$
and $c\in [a_i,b_i]$
such that
$
\kappa_{\Gamma}(c)>0.
$
We denote by $\kappa_{\gamma}(s)$ the Gaussian curvature of
$\gamma$ at $\Gamma(s)$.
Then we have that
$$
0<\kappa_{\Gamma}(c)=\epsilon_J(c)\kappa_{\gamma}(c),
$$
which proves the assertion.
\end{proof}
We now define the invariant $\mu(\gamma)$
mentioned in the introduction:
\begin{definition}
A function
$
\Phi:S^1_{\gamma}\to \{0,-1,1\}
$
is called an {\it admissible
function} of $\gamma$
if it satisfies the following conditions:
\begin{enumerate}
\item
$
\op{supp}(\Phi):=\Phi^{-1}(1)\cup \Phi^{-1}(-1)
$
is a finite set, and
\item for each $J\in \mathcal G(\gamma)$,
there exists $t\in \op{supp}(\Phi)$
such that $t\in J$ and $\epsilon_J(t)=\Phi(t)$.
\end{enumerate}
A point $t\in \op{supp}(\Phi)$ is called a {\it positive point}
(resp. {\it negative point})
if $\Phi(t)>0$ (resp. $\Phi(t)<0$).
We denote by $\mathcal A$ the set of all admissible functions
of $\gamma$.
\end{definition}
We fix an admissible function $\Phi\in \mathcal A$.
Then we have an expression
$
\op{supp}(\Phi)=\{u_1,...,u_{\ell}\}
$
such that
$
c_1\le u_1< u_2<\cdots < u_{\ell}\le c_{2m}.
$
Let $\mu(\Phi)$ be the
number of sign changes
of the sequence
$$
\Phi(u_1),\,\,\Phi(u_2),\,\,\cdots,\Phi(u_{\ell}),\,\,\Phi(u_1).
$$
Then we set
\begin{equation}
\mu(\gamma):=\min_{\Phi\in \mathcal A}\mu(\Phi).
\end{equation}
\medskip
\noindent
{\it Proof of the inequality \eqref{eq:affine}.}
Let $\gamma$ be a generic closed curve in $\R^2$.
We take a curve $\sigma\in \langle \gamma\rangle$
such that $i_\sigma=I(\gamma)$.
Without loss of generality, we may assume that
$i_\gamma=I(\gamma)$.
We can take a point $c_J\in [0,1]$ for each
$J\in \mathcal G$ in order that
$\op{sgn}(\kappa_\gamma(c_J))=\epsilon_J(c_J)$,
and that $c_{J}\neq c_K$ if $J\neq K$.
We define a function $\Phi:S^1_\gamma\to \{0,-1,1\}$
by
$$
\Phi(t):=
\begin{cases}
\op{sign}(\kappa_\gamma(c_J)) & \mbox{if $t=c_J$ for some
$J\in \mathcal G(\gamma)$},
\\
0 & \mbox{otherwise}.
\end{cases}
$$
Then, $\Phi$ is an admissible
function.
Since $i_\gamma=I(\gamma)$,
the curvature function of $\gamma$ changes sign at most
$I(\gamma)$ times. So we have that
$\mu(\Phi)\le I_\gamma$,
in particular, $\mu(\gamma)\le I_\gamma$.
\qed
Also we have the following assertion:
\begin{proposition}
\label{prop:trivial}
Let $\gamma$ be a generic closed curve in $\R^2$.
Then $\mu(\gamma)$ is a non-negative even integer,
as well as $I(\gamma)$. Moreover,
$\mu(\gamma)>0$ holds if and only if
$\mathcal G(\gamma)$ does not contain a positive
polygon and a negative polygon at the same time.
\end{proposition}
\begin{proof}
Since the number of sign changes of a cyclic sequence
of real numbers is always even, $\mu(\gamma)$ is also even.
Moreover,
if $\gamma$ has a positive (resp. negative)
polygon, each admissible
function
$\Phi$ must take a positive (resp. negative)
value.
So the existence of two distinct polygons of
opposite sign implies that $\mu(\gamma)\ge 2$.
Now, we prove the converse.
A closed curve which is not a simple closed curve $1_0$
has at least one shell, and a shell is necessarily
a positive or a negative polygon. Suppose that $\gamma$
has no negative polygons.
Then we can choose a point $c_J\in J\cap S^1_\gamma$
for each admissible polygon $J\in \mathcal G(\gamma)$
such that
$\epsilon_J(c_J)>0$.
If we set
$$
\Phi(t):=
\begin{cases}
\epsilon_J(c_J) &
\mbox{if $t=c_J$ for some $J\in \mathcal G(\gamma)$}, \\
0 & \mbox{otherwise},
\end{cases}
$$
then $\Phi$ is an admissible function, and
$\mu(\Phi)=0$. Thus we have $\mu(\gamma)=0$.
This proves the converse.
\end{proof}
To show the computability of the invariant $\mu(\gamma)$,
we fix $4m$ ($m=\#_\gamma$)
points $t_i,s_i$ ($i=1,...,2m$) satisfying
$$
0=c_1<t_1<s_1<c_2<\cdots < c_{2m}<t_{2m}<s_{2m}(<1),
$$
and show the following lemma:
\begin{lemma}\label{lemma:reduction}
For each admissible function $\Phi$ of $\gamma$,
there exists
an admissible function $\Psi$ satisfying the
following properties:
\begin{enumerate}
\item $\op{supp}(\Psi)\subset \{t_1,s_1,...,t_{2m},s_{2m}\}$,
\item $\mu(\Psi)\le \mu(\Phi)$.
\end{enumerate}
\end{lemma}
\begin{proof}
We fix an interval $U=(c_i,c_{i+1})$, where
$c_{2m+1}:=c_1$.
If $\Phi$ takes non-negative (resp. non-positive)
values on $U$,
then we set
$$
\Psi(t):=
\begin{cases}
1 \mbox{\,\,(resp. $-1$)} & \mbox{if $t=t_i$}, \\
0 & \mbox{if $t\in U\setminus\{t_i\}$}.
\end{cases}
$$
If $\op{supp}(\Phi)\cap U$ contains two points
$v_1,v_2$ such that $v_1<v_2$ and
$\Phi(v_1)=-\Phi(v_2)$, then we set
$$
\Psi(t):=
\begin{cases}
\Phi(v_1) & \mbox{if $t=t_i$}, \\
\Phi(v_2) & \mbox{if $t=s_i$}, \\
0 & \mbox{otherwise}.
\end{cases}
$$
Since $U$ is arbitrary, we get
a function $\Psi$ defined on $S^1_\gamma$.
Since $U \cap J$ coincides with
either $U$ or an empty set
for each $J\in \mathcal G(\gamma)$,
$\Psi$ is an admissible function, and one can easily verify
$\mu(\Psi)\le \mu(\Phi)$.
\end{proof}
\begin{remark}\label{rmk:computability}
(Computability of the invariant $\mu(\gamma)$)
The function $\Psi$ obtained in Lemma \ref{lemma:reduction}
is called a {\it reduction} of $\Phi$.
($\Psi$ may not be uniquely determined from $\Phi$,
since
$\op{supp}(\Phi)\cap U$ might consist of more than two points.)
By definition, there exists an admissible
function $\Phi$
such that $\mu(\Phi)=\mu(\gamma)$.
By Lemma \ref{lemma:reduction}, there is a
reduction of such a function $\Phi$.
Thus the invariant $\mu(\gamma)$ is attained by
a reduced function $\Psi$.
Since the number of reduced admissible
functions
is at most $3^{4m}$, the invariant $\mu(\gamma)$ can be
computed in a finite number of steps.
\end{remark}
\begin{remark}\label{rmk:flexibility}
(A flexibility of the reduced admissible
function)
In the above construction of the function $\Psi$ via
$\Phi$ we may set
$$
\Psi(t):=
\begin{cases}
\Phi(v_2) & \mbox{if $t=t_i$}, \\
\Phi(v_1) & \mbox{if $t=s_i$}, \\
0 & \mbox{otherwise},
\end{cases}
$$
when $\op{supp}(\Phi)\cap U$ contains two points
$v_1,v_2$ such that $v_1<v_2$ and
$\Phi(v_1)=-\Phi(v_2)$.
Then $\Psi$ is also an admissible
function.
This modification of $\Psi$ can be done for each
fixed interval $U=(c_i,c_{i+1})$.
However, after the operation, it might not hold that
$\mu(\Psi)\le \mu(\Phi)$.
\end{remark}
\begin{proposition}
\label{prop:crossings}
Let $\gamma$ be a generic closed curve in $\R^2$.
Then it satisfies
\begin{equation}\label{eq:hidden}
\mu(\gamma)\le 2\#_\gamma.
\end{equation}
\end{proposition}
\begin{proof}
As pointed out in Remark \ref{rmk:computability},
there exists a reduced admissible function $\Psi$.
Since $\mu(\Psi)\le 4m$, we get the estimate
$\mu(\gamma)\le 4\#_\gamma$.
However, we can improve it as follows:
As pointed out in Remark \ref{rmk:flexibility},
one can replace the values $\Psi(t_i),\Psi(s_i)$
by $-\Psi(t_i),-\Psi(s_i)$
whenever $\Psi(t_i)=-\Psi(s_i)$ for each $i=1,...,2m$.
So using this modification inductively for $i=1,...,2m$
if necessary,
we can modify $\Psi$ so that $\mu(\Psi)\le 2m$, and
then we have $\mu(\gamma)\le 2\#(\gamma)$.
\end{proof}
\begin{remark}
If $\gamma$ is a lemniscate $1^b_1$, then
$\mu(\gamma)= 2\#_\gamma=2$ holds.
However, the authors do not know of any other example
satisfying the equality $\mu(\gamma) = 2\#_\gamma$
(cf. Question 3).
\end{remark}
\begin{example}\label{ex:c2}(Curves with
a small number of intersections.)
Here, we demonstrate how to determine $\mu(\gamma)$
with $\#_\gamma\le 2$. In the eight classes of curves
in Figure \ref{fig:curve2a},
four classes have been drawn without inflections.
So $I(\gamma)=0$ holds for these four curves.
Each of the remaining four curves satisfies $\mu(\gamma)>0$,
by Proposition \ref{prop:trivial}.
On the other hand, these four curves
in Figure \ref{fig:curve2a}
have been drawn with exactly two inflections.
Thus we can conclude that they satisfy $\mu(\gamma)=2$.
Now, let $\gamma$ be a curve
as in Figure \ref{fig:i4}.
Then $\gamma$ has four disjoint shells,
two of which are positive, and the other two
are negative. So we can conclude that
$\mu(\gamma)\ge 4$. Since $\gamma$ as in
Figure \ref{fig:i4} has exactly four inflections,
we can conclude that $I(\gamma)=\mu(\gamma)=4$.
\end{example}
\begin{figure}[htb]
\begin{center}
\includegraphics[width=3.3cm]{long.eps}\qquad \quad
\includegraphics[width=2.8cm]{new_chain-like_curve.eps}
\caption{A chain-like curve and its realization with
$i_\gamma=2$.}
\label{fig:long}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{example}\label{ex:chain} (Chain-like curves.)
We consider a curve $\gamma$ with $\#_\gamma=n$
($n\ge 1$) as in Figure \ref{fig:long}, left.
This curve has two shells and $n-1$ leaves,
including a positive shell and a negative leaf,
which are disjoint.
Thus $(I(\gamma)\ge )\mu(\gamma) \ge 2$ holds.
As in Figure \ref{fig:long}, right,
this curve can be drawn along a spiral with
two inflections. So we can conclude that
$I(\gamma)=\mu(\gamma)=2$.
In this manner, drawing curves along a spiral
is often useful to reduce the number of inflection points.
Several useful techniques for drawing curves
with a restricted number of inflections
are mentioned in Halpern \cite[Section 4]{H2}.
\end{example}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=3.0cm]{new_curve_gamma_n_2.eps}
\caption{A figure of the curve $\gamma_n$.}
\label{fig:n}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{example}\label{ex:3}
(Curves with a negative $n$-gon.)
We consider a curve $\gamma_n$
as in
Figure \ref{fig:n},
which has several positive
polygons, but only
one negative admissible $n$-gon,
marked in gray in Figure \ref{fig:n}.
So we can conclude that $I(\gamma)=\mu(\gamma)=2$,
as in Figure \ref{fig:n}.
This example shows that an $n$-gon
$(n\ge 4)$ is needed
to find a curve that cannot be
locally convex.
Admissibility of a polygon is important for
the definition of the invariant $\mu$.
In fact, the curve
as in Figure \ref{fig:okey}, left, has
negative polygons which are not admissible, and
it can be realized
without inflections as in Figure \ref{fig:okey}, right.
\end{example}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=2.3cm]{ippan.eps}\qquad \qquad \quad
\includegraphics[height=2.5cm]{toppo.eps}
\caption{A curve satisfying $i_\gamma=0$.}
\label{fig:okey}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{example}\label{ex:ohno}
(A curve with an effective leaf which
is neither positive nor negative.)
We consider a curve $\gamma$ as
in Figure \ref{fig:ohno}.
This curve has $5$ crossings
and exactly two positive shells at $A$ and $B$.
It also has a negative shell at $C$.
Thus $\mu\ge 2$ by Proposition \ref{prop:trivial}.
We show that $\mu(\gamma)=4$ by way of contradiction:
There exist two positive points $p_1$ and $p_2$ on
the two positive shells at $A$ and $B$, respectively.
There is a unique simple closed arc $\Gamma$ bounded by
$A$ and $B$ which passes through $D$ and $E$.
Suppose that $\mu(\gamma)=2$. Then there are no
negative points on $\Gamma$.
Now we look at the negative leaf with vertices $A$ and $D$.
In Figure \ref{fig:ohno}, this leaf is marked in gray.
Since $\Gamma$ does not contain a negative point,
there must be a negative point $m_1$ between $A$
and $D$ on this leaf.
Since the curve has a symmetry,
applying the same argument to
the negative leaf at $B$ and $E$,
there is another negative point $m_2$
between $E$ and $C$.
Finally, we look at a leaf with vertices $D$ and $E$, which is
not positive nor negative.
Since $\Gamma$ has no negative point, there must have
a positive point $p_3$ on the arc on the
right-hand side of the leaf.
Since the sequence
$p_1,m_1,p_3,m_2,p_2$ changes sign four times,
this gives a contradiction. Thus
$\mu(\gamma)\ge 4$.
Since the curve can be drawn with exactly four inflections
as in Figure \ref{fig:ohno},
we can conclude that $I(\gamma)=\mu(\gamma)=4$.
In this example, an admissible polygon which
is neither positive nor negative plays
a crucial role to estimate the
invariant $I(\gamma)$
by using $\mu(\gamma)$.
\end{example}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=3.3cm]{figure5.eps}\qquad \qquad \quad
\caption{A curve satisfying $i_\gamma=4$ and $\#_\gamma=5$.}
\label{fig:ohno}
\end{center}
\vspace{-2ex}
\end{figure}
\medskip
\noindent
({\it Proof of the second assertion of Theorem \ref{thm:main}.})
The table of spherical curves up to $\#_\gamma\le 5$
is given in the appendix of \cite{KU}.
By moving the position of $\infty$ in $S^2=\R^2\cup\{\infty\}$,
we get the table of planar curves up to $\#_\gamma\le 5$
and can compute the invariant $\mu(\gamma)$.
So we can list the curves with $\mu(\gamma)=0$.
By Proposition \ref{prop:trivial}, it is sufficient
to check for
the existence of positive polygons and negative polygons.
When $\#_\gamma\le 2$, the number of topological
types of such curves is $3$.
If $\#_\gamma= 3,4,5$, then the number of
topological types of such curves is $6,16,50$, respectively.
After that we can draw the pictures of
the curves by hand.
If we are able to draw $76$ figures of the curves without inflections,
the proof is finished,
and this was accomplished in Figures \ref{fig:3},
\ref{fig:4} and \ref{fig:5}.
\qed
\medskip
\noindent
({\it Proof of Corollary \ref{cor:main}.})
For curves $\gamma$ with $\mu(\gamma)=0$,
we can show $I(\gamma)=0$ by drawing curves
without inflections.
On the other hand, when $\mu(\gamma)>0$,
we can show $I(\gamma)=2$ by drawing curves
with exactly two inflections, except for
the curve as in Figure \ref{fig:i4}.
\qed
\section{Double tangents and geotopical tightness}
In this section, we would like to give an application.
\begin{definition}
Let $\gamma$ be a generic planar curve.
We set
\begin{equation}
d(\gamma):=d_1(\gamma)+d_2(\gamma).
\end{equation}
Then we define
\begin{equation}
\delta(\gamma):=\min_{\sigma\in \langle \gamma\rangle}d(\sigma),
\end{equation}
which gives the minimum number of double tangents
of the curves in the equivalence class $\langle \gamma\rangle$.
(As in Figure \ref{fig:fb2}, $d(\sigma)$ might be different from
$d(\gamma)$ even if $\sigma\in \langle \gamma\rangle$ and $i_\gamma
=i_\sigma=I(\gamma)$.) A curve $\sigma\in \langle \gamma\rangle$ satisfying
$d(\gamma)=\delta(\gamma)$
is called {\it geotopically tight} or {\it $g$-tight}.
We call the integer $\delta(\gamma)$
the {\it $g$-tightness number}.
\end{definition}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=3.5cm]{fb2a.eps}\qquad \qquad \quad
\includegraphics[height=4.0cm]{fb2b.eps}
\caption{Curves of $2_2^c$ with $i_\gamma=2$
but with different $d$.}
\label{fig:fb2}
\end{center}
\vspace{-2ex}
\end{figure}
The following assertion holds:
\begin{proposition}\label{prop:d}
It holds that
$\delta(\gamma)\ge \#_\gamma+I(\gamma)/2$.
\end{proposition}
\begin{proof}
By \eqref{eq:FB}, we get that
$$
d(\sigma)=d_1(\sigma)+d_2(\sigma)
\ge d_1(\sigma)-d_2(\sigma)=\#_\sigma+\frac{I(\sigma)}2
$$
holds for $\sigma\in \langle \gamma\rangle$.
Taking the infimum, we get the assertion.
\end{proof}
We expect that any locally convex
generic closed curves might be $g$-tight
(see Question 4).
Relating this, we can prove the following
assertion:
Let $\gamma:\R\to\R^2$ be a periodic parametrization
by arclength of a locally convex curve with total
length $\ell$, that is,
$\gamma(t+\ell)=\gamma(t)$ holds for $t\in\R$.
Without loss of generality, we may assume
that the curvature of $\gamma$ is positive.
Let
$$
x=\gamma(s_1)=\gamma(s_2)\qquad (0<|s_1-s_2|<\ell)
$$
be a crossing of $\gamma$.
Replacing $s_1$ by $s_1+\ell$ if necessary,
we may assume that $s_1<s_2$ and
$\dot\gamma(s_2)$ is a positively rotated vector of
$\dot\gamma(s_1)$ through an angle $\alpha$ with $0<\alpha<\pi$.
When the parameter $t$ varies in the interval $[s_1,s_2]$,
the tangent vector $\dot\gamma(t)$
rotates through an angle $2\pi n_1+\alpha$,
where $n_1$ is a positive integer.
Similarly, for the interval $[s_2,s_1+\ell]$,
there exists a positive integer $n_2$
such that the rotation angle
of $\dot\gamma(t)$ is equal to $2\pi n_2-\alpha$.
The sum $n_1+n_2$ is the total rotation index of $\gamma$.
Denote by $W(x)$ the difference $n_1-n_2$.
We easily recognize that the sum of $W(x)$ for each crossing
$x$ of $\gamma$ is a geotopy invariant.
The following theorem can be proved using the equality
of $d_2$ in \cite[p7]{O} (we omit the details):
\begin{theorem}\label{prop:conj}
The number of double tangents for
any locally convex generic closed
curve $\gamma$ depends only on its
geotopy type. More precisely, the following identity
holds:
\begin{equation}\label{eq:new}
d_2(\gamma)
=\sum W(x),
\end{equation}
where the sum runs over all crossings $x$ of $\gamma$.
\end{theorem}
\begin{remark}\label{rmk:halpern}
The rotation index $R_\gamma$ of $\gamma$ which is
at each crossing equal to $n_1+n_2$, as mentioned above,
is less than or equal to
$\#_\gamma+1$ (cf. [{W}]).
Thus the formula \eqref{eq:new} implies
$$
d_2=\#_\gamma R_\gamma-2\sum W(x)\le
\#_\gamma (\#_\gamma+1)-2\#_\gamma=\#_\gamma (\#_\gamma-1),
$$
which reproves Halpern's conjecture in [{H2}].
\end{remark}
\begin{corollary}\label{cor:d}
$\delta(\gamma)=1,2,2,3,3$ for $\gamma$ of
type $1_1,1_1^b,1_2,2_2^b,2_2^c$ as in Figure \ref{fig:curve2a},
respectively.
\end{corollary}
\begin{proof}
By Proposition \ref{prop:d},
$\delta(\gamma)\ge 1,2,3,3$ for $\gamma$ of
type $1^b,2_2^b,2_2^c$, respectively.
On the other hand, the curves given
in Figure \ref{fig:curve2a} attain equality
in this inequality.
\end{proof}
In Figure \ref{fig:curve2a}, there are two remaining
curves of type $1_2^b$ and $2_2$,
whose g-tightness numbers have not been specified
by the authors.
The curve of type $2_2$ given in Figure \ref{fig:curve2a}
satisfies $d(\gamma)=4$, and we expect that $\delta(\gamma)=4$.
On the other hand,
the corresponding curve as in
the right of Figure \ref{fig:curve2a}
satisfies $d=7$, but one can realize the curve
with $d=5$ as in Figure \ref{fig:fb}.
We expect that it might be
$g$-tight. If true, $\delta(\gamma)=5$ holds.
\begin{figure}[htb]
\begin{center}
\includegraphics[height=3cm]{daruma.eps}\qquad \qquad \quad
\includegraphics[height=3cm]{seven.eps}
\caption{Curves of type $2_2$ and $1^b_2$.}
\label{fig:75}
\end{center}
\vspace{-2ex}
\end{figure}
\bigskip
A relationship between inflection points and
double tangents for simple closed curves
in the real projective plane
with a suitable convexity is given in \cite{TU}.
Finally, we leave several open questions on
the invariants $I(\gamma)$ and $\delta(\gamma)$:
\medskip
\noindent
{Question 1.}
{\it Does $\mu(\gamma)=0$ imply
$I(\gamma)=0$?}
\medskip
\noindent
{Question 2.}
{\it Is there a generic closed curve
satisfying $I(\gamma)>\mu(\gamma)$?}
\medskip
The authors do not know of any such examples.
If we suppose $I(\gamma)= \mu(\gamma)$,
then
\eqref{eq:hidden} yields the
inequality $I(\gamma)\le 2\#_\gamma$.
\medskip
\noindent
{Question 3.}
{\it Does
$I(\gamma)\le 2\#_\gamma$ hold
for any generic closed curve in $\R^2$?}
\medskip
\noindent
{Question 4.}
{\it Is an arbitrary locally convex curve
$g$-tight?}
\medskip
\noindent
{Question 5.}
{\it Find a criterion for $g$-tightness.
For example, can one determine
$\delta(\gamma)$ when
$\gamma$ is of type $2_2$ or $1^b_2$?
}
\begin{acknowledgements}
The authors thank Wayne Rossman
for careful reading of the first draft and for giving
valuable comments.
\end{acknowledgements}
\section{Tables of curves.}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=1.2cm]{cross3.eps}\\
\caption{Curves with $I_\gamma=0$ and $\#_\gamma= 3$
(6 curves in total).}
\label{fig:3}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=2.5cm]{cross3g.eps}\qquad \quad
\caption{Curves with $I(\gamma)>0$ and $\#_\gamma\le 3$ (13 curves
in total).}
\label{fig:3g}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=3.5cm]{cross4.eps}
\caption{Curves with $I(\gamma)=0$ and $\#_\gamma\le 4$ (16 curves
in total).}
\label{fig:4}
\end{center}
\vspace{-2ex}
\end{figure}
\begin{figure}[htb]
\begin{center}
\includegraphics[height=6.5cm]{cross5b.eps}\qquad \quad
\caption{Curves with $I(\gamma)=0$ and $\#_\gamma=5$ (50 curves
in total).}
\label{fig:5}
\end{center}
\vspace{-2ex}
\end{figure}
|
\section{Introduction}
\section{The discrete Riesz $s$-energy problem}
The {\em Riesz $s$-energy} (real $s\neq 0$) and the {\em logarithmic energy} ($s = 0$, by convention) of an $N$-point configuration $X_N$ with points $\PT{x}_1, \dots, \PT{x}_N$ in the Euclidean space $\mathbb{R}^p$ ($p \geq 1$)
are defined as
\begin{equation*}
E_s(X_N) {:=} \mathop{\sum_{j=1}^N \sum_{k=1}^N}_{j \neq k} \frac{1}{\left| \PT{x}_j - \PT{x}_k \right|^s}, \qquad E_0(X_N) {:=} \mathop{\sum_{j=1}^N \sum_{k=1}^N}_{j \neq k} \log \frac{1}{\left| \PT{x}_j - \PT{x}_k \right|}.
\end{equation*}
The logarithmic energy can be understood as the limiting case $s \to 0$. For large $s$ the nearest neighbor's interaction dominates and the other limiting case $s \to \infty$ gives the {\em best-packing problem} (or {\em Tammes problem} \cite{tammes:1930}). Most deeply studied is the classical ({\em Newtonian}) case $s = p - 2$ for the harmonic potential $1/r^{p-2}$, naturally with an abundance of literature in physics in the Coulomb case ($p = 3$), where the Riesz $s$-energy essentially is the potential energy of an ensemble of (positive) unit point charges placed at the points of the configuration $X_N$.
For this note of interest are {\em sums of distances} (that is $s = -1$) for points on the sphere $\mathbb{S}^d$ because of the close connection to the spherical cap $\mathbb{L}_2$-discrepancy defined in Definition~\ref{def:sph.cap.L.2.discr} by means of Stolarsky's invariance principle (Proposition~\ref{prop:Stolarsky.inv.principle}) and to the worst-case error for equal weight quadrature formula for functions in the unit ball in a certain reproducing kernel Hilbert space over $\mathbb{S}^d$ (cf. \cite{BrDi2011_pre} and \cite{BrWo20xx}).
The {\em $N$-point $s$-energy} of an infinite compact set $A \subseteq \mathbb{R}^p$ is defined as follows:
\begin{align*}
\mathcal{E}_s( A; N ) &{:=} \sup \left\{ E_s( \PT{x}_1, \dots, \PT{x}_N ) : \PT{x}_1, \dots, \PT{x}_N \in A \right\} &\text{if $s < 0$,} \\
\mathcal{E}_s( A; N ) &{:=} \inf \left\{ E_s( \PT{x}_1, \dots, \PT{x}_N ) : \PT{x}_1, \dots, \PT{x}_N \in A \right\} &\text{if $s \geq 0$.}
\end{align*}
It gives the optimal $s$-energy an $N$-point configuration $X_N$ in $A$ can assume; that is
\begin{equation*}
\mathcal{E}_s( A; N ) = E_s(X_N^{(s)}) \qquad \text{for an {\em optimal $s$-energy system $X_N^{(s)}$}.}
\end{equation*}
Let $d$ denote the Hausdorff dimension of $A$. When normalized appropriately (such that the total 'charge' of the $N$-point configuration is one) the quantity $\mathcal{E}_s( A; N ) / N^2$ remains finite (and has a limit as $N \to \infty$)\footnote{Indeed, for example, for $s > 0$ the sequence $N(N-1)/\mathcal{E}_s( A; N )$ is monotone and bounded and converges to the 'generalized transfinite diameter of $A$ of order $s$' defined by P{\'o}lya and Szeg{\H{o}}~\cite{PoSz1931}.} in the {\em potential-theoretical regime $s < d$} but grows beyond any bound in the {\em hypersingular case} $s \geq d$. In the regime $s<d$ tools and result from classical potential theory (see, for example, Bj{\"o}rck~\cite{Bj1956} ($s<0$), Saff and Totik~\cite{SaTo1997} (logarithmic case) and Landkof~\cite{La1972}) are used to treat the limit $N\to\infty$. The limit distribution of optimal $s$-energy configurations is given by the (unique if $-2<s<d$) equilibrium (or extremal) measure $\mu_A$ maximizing (if $s<0$) or minimizing (if $s\geq0$) the energy integral
\begin{equation*}
\mathcal{I}_s[\mu] {:=} \int \int k_s(\PT{x}, \PT{y}) \mathop{\mathrm{d}} \mu(\PT{x}) \mathop{\mathrm{d}} \mu(\PT{y}),
\end{equation*}
where $k_s(\PT{x}, \PT{y}) {:=} | \PT{x} - \PT{y} |^{-s}$ if $s\neq0$ and $k_0(\PT{x}, \PT{y}) {:=} - \log | \PT{x} - \PT{y} |$, among all Borel probability measures $\mu$ supported on $A$. The extremal quantity $V_s(A) = \mathcal{I}_s[\mu_A]$ is called the {\em $s$-energy of the set $A$}. If $s \geq d$, then $\mathcal{I}_s[\mu] = + \infty$ for any measure $\mu$ supported on $A$, that is $V_s(A) = +\infty$. In this case methods and results from geometrical measure theory are successfully used to gain more insight. One outcome is that other normalizations of the optimal $s$-energy would give a converging sequence for certain classes of compact sets (see, for example, Kuijlaars and Saff~\cite{KuSa1998} and Hardin and Saff~\cite{HaSa2005}). These results are the building blocks for the fundamental conjecture below.
The discrete Riesz energy problem is concerned with investigating properties of sequences of optimal (and nearly optimal) Riesz $s$-energy configurations. Questions concern
\begin{inparaenum}[\itshape (i)]
\item explicit computations of optimal configurations,
\item limit distribution,
\item asymptotic expansion of the optimal energy (cf. Smale's Problem \#7~\cite{Sm1998}),
\item geometric properties ('well-separation', 'mesh'-norm).
\end{inparaenum}
We refer the interested reader to the survey articles Saff and Kuijlaars~\cite{SaKu1997} and Hardin and Saff~\cite{HaSa2004}
Let $\mathbb{S}^d$ denote the unit sphere in $\mathbb{R}^{d+1}$. The following long-standing open fundamental conjecture for the asymptotic expansion of the optimal Riesz $s$-energy is known. (In the upcoming paper \cite{BrHaSa2009a_pre} this conjecture and its history will be discussed in more detail.) Let $\mathcal{H}_d$ denote the $d$-dimensional Hausdorff measure (normalized so that the $d$-dimensional unit cube in $\mathbb{R}^p$ has measure $1$).
\begin{conj}[Fundamental Conjecture] \label{conj:Riesz.s.energy.Sd}
Let $d \geq 2$ and $-2 < s < d + 2$ ($s \neq 0,d$).\footnote{For $s<-2$ the extremal distribution is concentrated in no more than $d+1$ points on $\mathbb{S}^d$ (\cite{Bj1956}). } Then
\begin{equation*}
\mathcal{E}_s( \mathbb{S}^d; N ) = V_s( \mathbb{S}^d ) \, N^2 + \frac{C_{s,d}}{\left[ \mathcal{H}_d( \mathbb{S}^d ) \right]^{s/d}} \, N^{1 + s / d} + \mathcal{R}_s( \mathbb{S}^d; N ),
\end{equation*}
where $\mathcal{R}_s( \mathbb{S}^d; N ) / N^{1 + s / d - \varepsilon} \to 0$ as $N \to \infty$ for some $\varepsilon > 0$ possibly depending on $d$ and $s$.
\end{conj}
This conjecture combines known results on the basis of the {\em principle of analytic continuation}:
\begin{itemize}
\item In the potential-theoretical regime $-2 < s < d$ ($s \neq 0$) the dominant term grows like $N^2$ and the coefficient is given by the $s$-energy of the unit sphere
\begin{equation} \label{eq:V.s.S.d}
V_s( \mathbb{S}^d ) = 2^{d-s-1} \frac{\gammafcn((d+1)/2) \gammafcn((d-s)/2)}{\sqrt{\pi}\gammafcn(d-s/2)}, \qquad -2 < s < d (s \neq 0).
\end{equation}
For other values of $s$ it is understood to be the analytic continuation of the right-hand side above to the complex $s$-plane except at the simple poles at $s = d, d + 2, \dots, 2 d - 2$ (finitely many) if $d$ is even and at $s = d, d + 2, d + 4, \dots$ (infinitely many) if $d$ is odd.
\item In the hyper-singular case $s > d$ the leading term behaves like $N^{1+s/d}$ and it is shown by Kuijlaars and Saff \cite{KuSa1998} that the limit $\mathcal{E}_s( \mathbb{S}^d; N ) / N^{1+s/d}$ exists.\footnote{The existence of this limit is proven for the class of rectifiable $d$-dimensional manifolds in \cite{HaSa2005} and for weighted Riesz $s$-energy for rectifiable $d$-dimensional sets in \cite{BoHaSa2008}.} It is believed that the constant $C_{s,d}$ can be analytically continued to the $s$-plane, which is supported by the conjecture for its value in case of $d=2$ also provided in \cite{KuSa1998}.
\item In the boundary case $s=d$ \cite{KuSa1998} gives that
\begin{equation*}
\mathcal{E}_d( \mathbb{S}^d; N ) \sim \frac{\mathcal{H}_d(\mathbb{B}^d)}{\mathcal{H}_d(\mathbb{S}^d)} \, N^2 \log N \qquad \text{as $N \to \infty$,}
\end{equation*}
which can be also understood by means of a limit process $s \to d$ assuming that the poles of $V_s(\mathbb{S}^d)$ and $C_{s,d}$ at $s=d$ cancel each other. (Obviously, one can also consider limit processes $s \to s^\prime$, where $s^\prime$ is a pole of $V_s$. In such a case one would gain information about the singularity at $s=s^\prime$ of the coefficient of the corresponding term in the asymptotic expansion of the optimal $s$-energy ($s$ near $s^\prime$) provided this lesser-order term exists. The limit process $s\to0$ connects the asymptotic expansion of the optimal logarithmic energy and optimal $s$-energy for $s$ near $0$.)
\end{itemize}
The constant $\mathcal{H}_d(\mathbb{B}^d) / \mathcal{H}_d(\mathbb{S}^d)$ is the ratio of the volume of the unit ball $\mathbb{B}^d$ in $\mathbb{R}^d$ and the surface area of $\mathbb{S}^d$ (denoted by $\omega_d$) and can be expressed in terms of a ratio of surface areas or a ratio of gamma functions
\begin{equation} \label{eq:ratio}
\frac{\mathcal{H}_d(\mathbb{B}^d)}{\mathcal{H}_d(\mathbb{S}^d)} = \frac{1}{d} \frac{\omega_{d-1}}{\omega_d} = \frac{1}{d} \, \frac{\gammafcn((d+1)/2)}{\sqrt{\pi} \gammafcn(d/2)} \sim \frac{1}{\sqrt{2\pi}} \, d^{-1/2} \quad \text{(as $d \to \infty$).}
\end{equation}
It is a long-standing open problem what the precise value of the constant
\begin{equation} \label{eq:C.s.d}
C_{s,d} = \lim_{N \to \infty} \mathcal{E}_s( [0,1]^d; N ) \big/ N^{1+s/d}
\end{equation}
is for $d \geq 2$. (In the case $d = 1$ Mart{\'{\i}}nez-Finkelshtein et al. showed that $C_{s,1}=2 \zeta(s)$, where $\zeta(s)$ is the Riemann zeta function.) The significance of the constant $C_{s,d}$ follows from the fact that the limit $\lim_{s\to\infty}[ C_{s,d} ]^{1/s}$ can be expressed in terms of the 'largest sphere packing density' in $\mathbb{R}^d$, which is only known in three cases: $d = 1,2$ and $d = 3$ (Kepler conjecture, rather recently proved by Hales~\cite{Ha2005}). For $d = 2$ Kuijlaars and Saff \cite{KuSa1998} obtained the estimate
\begin{equation*}
\limsup_{N \to \infty} \frac{\mathcal{E}_s( \mathbb{S}^2; N )}{N^{1+s/2}} \leq \frac{\left( \sqrt{3} / 2 \right)^{s/2} \zetafcn_{\Lambda}(s)}{\left( 4 \pi \right)^{s/2}}, \qquad s > 2,
\end{equation*}
and they conjecture that equality holds above; that is:
\begin{conj}[Kuijlaars and Saff \cite{KuSa1998}] \label{conj:C.s.2}
For $s > 2$, $C_{s,2} = \left( \sqrt{3} / 2 \right)^{s/2} \zetafcn_{\Lambda}(s)$.
\end{conj}
The function $\zetafcn_\Lambda(s)$ is the zeta function of the hexagonal lattice $\Lambda = \{ m \, (1, 0) + n (1/2, \sqrt{3}/2) : m, n \in \mathbb{Z} \}$, which was used in \cite{KuSa1998} to locally approximate a minimal $s$-energy configuration on the sphere in the hypersingular case $s>2$. For $\mathop{\mathrm{Re}} s > 2$ it is defined as
\begin{equation*}
\zetafcn_\Lambda(s) = \sum_{\PT{0} \neq \PT{a} \in \Lambda} \frac{1}{\left| \PT{a} \right|^s} = \sum_{(m,n) \in \mathbb{Z}^2\setminus \{ (0,0) \}} \frac{1}{\left( m^2 + m n + n^2 \right)^{s/2}}
\end{equation*}
and it is understood that $\zetafcn_\Lambda(s)$ is the meromorphic extension to $\mathbb{C}$ of the right-hand side above.
The function $\zetafcn_\Lambda(s)$ admits a factorization (cf., for example, \cite[Chapter~X, Section~7]{Co1980})
\begin{equation} \label{zeta.lambda.prod}
\zetafcn_\Lambda(s) = 6 \zetafcn( s / 2 ) \DirichletL_{-3}( s / 2 ), \qquad \mathop{\mathrm{Re}} s > 2,
\end{equation}
into a product of the Riemann zeta function $\zetafcn$ and the first negative primitive Dirichlet $\DirichletL$-Series
\begin{equation}
\DirichletL_{-3}(s) {:=} 1 - \frac{1}{2^s} + \frac{1}{4^s} - \frac{1}{5^s} + \frac{1}{7^s} - \cdots, \qquad \mathop{\mathrm{Re}} s > 1.
\end{equation}
Interestingly, it is assumed in physics that a hexagonal configuration has lowest potential energy but there seems to be no mathematical proof for this. However, it is known that the hexagonal lattice is (even universal) optimal among all lattices in $\mathbb{R}^2$, see Montgomery~\cite{Mo1988} and Cohn and Kumar~\cite{CoKu2007}.
\section{Uniform Distribution and Discrepancy}
\begin{defn} \label{def:asymptotically.uniformly.distributed}
A sequence $\{ X_N \}$ is {\em asymptotically uniformly distributed on $\mathbb{S}^d$} if
\begin{equation*}
\lim_{N\to\infty} \frac{\#\left\{ k : \PT{x}_{k,N} \in B \right\}}{N} = \sigma_d(B)
\end{equation*}
for every $\sigma_d$-measurable clopen set $B$ in $\mathbb{S}^d$.
\end{defn}
Informally speaking: a reasonable test set gets a fair share of points as $N$ becomes large.
The quality of a sequence $\{ X_N \}$ of $N$-point systems can be quantified using the discrepancy
\begin{equation*}
D(\mathcal{F};\PT{x}_1,\dots,\PT{x}_N) {:=} \sup_{B\in\mathcal{F}} \left| \frac{\#\left\{ k : \PT{x}_k\in B \right\}}{N} - \sigma_d(B) \right|
\end{equation*}
measuring the maximum deviation between the uniform measure (limit distribution of optimal configurations) and empirical point distribution with respect to a family $\mathcal{F}$ of test sets (for example spherical caps). The {\em spherical cap discrepancy} of an $N$-point configuration $X_N$ will be denoted by $\capDISCR( X_N )$.
A different approach to uniform distribution makes use of numerical integration and function spaces. Let
\begin{equation*}
\xctint[f] {:=} \int_{\mathbb{S}^d} f \mathop{\mathrm{d}} \sigma_d, \qquad \displaystyle Q_N[f] {:=} \frac{1}{N} \sum_{k=1}^N f(\PT{x}_k).
\end{equation*}
\begin{defn} \label{def:equi-distributed.cont.fcns}
$\{ X_N \}$ is {\em equi-distributed with respect to every function in the space of continuous functions} if \footnote{ That is, the discrete probability measure associated with $X_N$ tends to $\sigma_d$ as $N \to \infty$ (in the weak-$*$ limit). }
\begin{equation*}
\lim_{N \to \infty} \numint_N[f] = \xctint[f] \qquad \text{for every $f \in C(\mathbb{S}^d)$.}
\end{equation*}
\end{defn}
As it is well-known both Definitions \ref{def:asymptotically.uniformly.distributed} and \ref{def:equi-distributed.cont.fcns} are equivalent.
On the sphere one has no satisfactory analogue to the celebrated Koksma-Hlawka inequality in the unit cube. Cui and Freeden~\cite{CuFr1997} introduced the concept of {\em generalized discrepancy on $\mathbb{S}^2$} based on pseudo-differential operators. For a particular choice $\mathbf{D}$ they obtained a {\em Koksma-Hlawka} like inequality
\begin{equation*}
\big| \numint_N[f] - \xctint[f] \big| \leq \sqrt{6} \, \mathrm{D}_{\mathrm{CF}}(X_N) \, \| f \|_{\mathbb{H}^{3/2}(\mathbb{S}^2)},
\end{equation*}
where $f$ is from a certain Sobolev space $\mathbb{H}^{3/2}(\mathbb{S}^2)$ whose reproducing kernel is defined using $\mathbf{D}$. In this context, a sequence $\{X_N\}$ of $N$-point systems is called {\em $\mathbf{D}$-equidistributed with respect to all functions in $\mathbb{H}^{3/2}(\mathbb{S}^2)$} if $\lim_{N\to\infty} \mathrm{D}_{\mathrm{CF}}(X_N) = 0$. Moreover, the generalized discrepancy associated with $\mathbf{D}$ has a closed form expressible in terms of elementary functions
\begin{equation*}
4 \pi \left[ \mathrm{D}_{\mathrm{CF}}(X_N) \right]^2 = 1 - \frac{1}{N^2} \sum_{j, k = 1}^N \log \left( 1 + \left| \PT{x}_j - \PT{x}_k \right| / 2 \right)^2.
\end{equation*}
(Sloan and Womersley~\cite{SlWo2004} showed that $\left[ \mathrm{D}_{\mathrm{CF}}(X_N) \right]^2$ has a natural interpretation as the worst-case error for $Q_N$ for function from the unit ball in $\mathbb{H}^{3/2}(\mathbb{S}^2)$ provided with a norm which is equivalent to the one used by Cui and Freeden.) This approach is followed further in \cite{BrWo20xx} leading to the generalized discrepancy
\begin{equation} \label{eq:sum.dist.discr}
\left[ \mathrm{D}(X_N) \right]^2 = \frac{4}{3} - \frac{1}{N^2} \sum_{k, \ell = 1}^N \left| \PT{x}_\ell - \PT{x}_k \right|
\end{equation}
associated with the Sobolev space $\mathbb{H}^{3/2}(\mathbb{S}^2)$ with the reproducing kernel $K(\PT{x}, \PT{y}) = (8 / 3) - |\PT{x} - \PT{y} |$.
\section{The spherical cap $\ELLtwo$-discrepancy on $\mathbb{S}^d$}
A spherical cap on $\mathbb{S}^d$ centered at $\PT{x} \in \mathbb{S}^d$ is the set
\begin{equation*}
C(\PT{x}; t) {:=} \left\{ \PT{y} \in \mathbb{S}^d : \left\langle \PT{x}, \PT{y} \right\rangle \geq t \right\},
\end{equation*}
where $\langle \PT{\cdot}, \PT{\cdot} \rangle$ denotes the usual inner product. The family of all spherical caps on $\mathbb{S}^d$ forms a {\em discrepancy system} (see \cite{DrTi1997}) giving rise to the notion of {\em asymptotic uniform distribution}
\begin{defn} \label{def:sph.cap.L.2.discr}
The {\em spherical cap $\ELLtwo$-discrepancy} of an $N$-point configuration $X_N$ {on $\mathbb{S}^d$} is given by
\begin{equation*}
\ELLtwoDISCR( X_N ) {:=} \left[ \int_{-1}^1 \int_{\mathbb{S}^d} \left| \frac{\left| X_N \cap C( \PT{x}, t) \right|}{N} - \sigma_d( C( \PT{x}, t ) ) \right|^2 \mathop{\mathrm{d}} \sigma_d( \PT{x} ) \mathop{\mathrm{d}} t \right]^{1/2}.
\end{equation*}
The {\em optimal $\ELLtwo$-discrepancy} of $N$-point configurations on $\mathbb{S}^d$ is denoted by
\begin{equation*}
\ELLtwoDISCR( \mathbb{S}^d; N ) {:=} \inf \left\{ \ELLtwoDISCR( X_N ) : X_N \subseteq \mathbb{S}^d \right\}.
\end{equation*}
\end{defn}
An obvious upper bound of the $\ELLtwo$-discrepancy is in terms of the spherical cap discrepancy:
\begin{equation*}
\ELLtwoDISCR( X_N ) \leq \sqrt{2} \, \capDISCR( X_N ), \qquad X_N \subseteq \mathbb{S}^d.
\end{equation*}
The next result connects the sum of distances, the $\ELLtwo$-discrepancy and the distance integral.
\begin{prop}[Stolarsky's invariance principle~\cite{St1973}] \label{prop:Stolarsky.inv.principle} \footnote{The given version is a special case of the result in \cite{St1973} which is extended in \cite{Br2003}.} Let $d \geq 2$. Then
\begin{equation*}
\frac{1}{N^2} \sum_{j, k = 1}^N \left| \PT{x}_j - \PT{x}_k \right| + \frac{\mathcal{H}_d(\mathbb{S}^d)}{\mathcal{H}_d(\mathbb{B}^d)} \left[ \ELLtwoDISCR( X_N ) \right]^2 = \int \int \left| \PT{x} - \PT{x}^\prime \right| \mathop{\mathrm{d}} \sigma_d(\PT{x}) \mathop{\mathrm{d}} \sigma_d(\PT{x}^\prime), \qquad X_N \subseteq \mathbb{S}^d.
\end{equation*}
\end{prop}
Clearly, reduction of the $\ELLtwo$-discrepancy means increase of the sum of distances and vice versa.
Stolarsky's invariance principle provides a simple way to compute the $\ELLtwo$-discrepancy of a given $N$-point configuration on $\mathbb{S}^d$. In the language of the discrete energy problem we have
\begin{equation} \label{eq:ELLtwoDISCR}
\frac{\mathcal{H}_d(\mathbb{S}^d)}{\mathcal{H}_d(\mathbb{B}^d)} \left[ \ELLtwoDISCR( X_N ) \right]^2 = V_{-1}(\mathbb{S}^d) - \frac{1}{N^2} \sum_{j, k = 1}^N \left| \PT{x}_j - \PT{x}_k \right| = - \left[ E_{-1}(X_N) - V_{-1}(\mathbb{S}^d) \, N^2 \right] / N^2.
\end{equation}
Evidently, the square of the optimal $\ELLtwo$-disrepancy is closely related to the second term of the asymptotic expansion of the optimal Riesz $(-1)$-energy, cf. Conjecture~\ref{conj:Riesz.s.energy.Sd}. This connection between the spherical cap ($\mathbb{L}_2$-)discrepancy and the second term in the asymptotic expansion of the $s$-energy is exploited in \cite{Br2008}, Grabner and Damelin~\cite{DaGr2003}.
The $\ELLtwo$-discrepancy of an $N$-point configuration $X_N$ is minimial if and only if the sum of distances of $X_N$ is maximal. Thus, optimal $\ELLtwo$-discrepancy configurations are, in fact, maximal sum-of-distance configurations or, equivalently, maximal Riesz $(-1)$-energy configurations. Such configurations can be generated using numerical optimization which is a highly non-linear process.
Stolarsky used his invariance principle and discrepancy results of Schmidt~\cite{Sch1969} on the discrepancy of spherical caps to estimate the difference
\begin{equation*}
\int \int \left| \PT{x} - \PT{x}^\prime \right| \mathop{\mathrm{d}} \sigma_d(\PT{x}) \mathop{\mathrm{d}} \sigma_d(\PT{x}^\prime) - \frac{\mathcal{E}_{-1}( \mathbb{S}^d; N )}{N^2}
\end{equation*}
and obtained the correct order of $N$ for the upper bound. Harman~\cite{Ha1982} improved the lower bound (for the general invariance principle) and, finally, Beck~\cite{Be1984} obtained the correct order of $N$ for the Euclidean metric. By means of the invariance principle these bound for the energy difference translate into bounds for the spherical cap $\ELLtwo$-discrepancy.
\begin{prop} \label{prop:ELLtwoDISCR.bounds}
Let $d \geq 2$. There exist constants $c^\prime, C^\prime > 0$ such that for sufficiently large $N$
\begin{equation*}
c^\prime \, N^{-1/2 - 1/(2d)} \leq \ELLtwoDISCR( \mathbb{S}^d; N ) \leq C^\prime \, N^{-1/2 - 1/(2d)}.
\end{equation*}
\end{prop}
\begin{rmk}
By Proposition~\ref{prop:ELLtwoDISCR.bounds} the correct order of the decay of the optimal $\ELLtwo$-discrepancy on $\mathbb{S}^d$ in terms of powers of the number of points $N$ is $N^{-1/2 - 1/(2d)}$ which is the same rate (apart from the $\sqrt{\log N}$ term in the upper bound) as for the optimal spherical discrepancy on $\mathbb{S}^d$.
\end{rmk}
Proposition~\ref{prop:ELLtwoDISCR.bounds} rises the question if $\ELLtwoDISCR( \mathbb{S}^d; N ) \, N^{1/2 + 1/(2d)}$ has a limit as $N \to \infty$. By employing the connection to the discrete Riesz $s$-energy problem, where one has conjectures regarding the asymptotic expansion of the optimal $s$-energy, we derive the following conjectures for the leading term in the asymptotic expansion of the optimal $\ELLtwo$-discrepancy.
\begin{conj} \label{conj:L2discr.Sd}
Let $d \geq 2$. If Conjecture~\ref{conj:Riesz.s.energy.Sd} holds, then
\begin{equation} \label{eq:A.d}
\ELLtwoDISCR( \mathbb{S}^d; N ) \sim A_d \, N^{-1/2 - 1/(2d)} + \cdots \qquad \text{as $N \to \infty$,} \qquad A_d {:=} \sqrt{\frac{\mathcal{H}_d(\mathbb{B}^d)}{\mathcal{H}_d(\mathbb{S}^d)} \, \frac{-C_{-1,d}}{\left[ \mathcal{H}_d( \mathbb{S}^d ) \right]^{-1/d}}}.
\end{equation}
\end{conj}
\begin{proof}[Justification]
By Stolarsky's invariance principle (Prop.~\ref{prop:Stolarsky.inv.principle} and Eq.~\eqref{eq:ELLtwoDISCR})
\begin{equation*}
\frac{\mathcal{H}_d(\mathbb{S}^d)}{\mathcal{H}_d(\mathbb{B}^d)} \left[ \ELLtwoDISCR( \mathbb{S}^d; N ) \right]^2 = - \left[ \mathcal{E}_{-1}(\mathbb{S}^d; N) - V_{-1}( \mathbb{S}^d ) \, N^2 \right] / N^2.
\end{equation*}
Application of the fundamental Conjecture~\ref{conj:Riesz.s.energy.Sd} yields
\begin{equation*}
\frac{\mathcal{H}_d(\mathbb{S}^d)}{\mathcal{H}_d(\mathbb{B}^d)} \left[ \ELLtwoDISCR( \mathbb{S}^d; N ) \right]^2 = \frac{-C_{-1,d}}{\left[ \mathcal{H}_d( \mathbb{S}^d ) \right]^{-1/d}} \, N^{-1 - 1 / d} - \mathcal{R}_{-1}( \mathbb{S}^d; N ) / N^2.
\end{equation*}
The conjecture follows.
\end{proof}
For the unit sphere $\mathbb{S}^2$ one can make the conjecture more precise.
\begin{conj} \label{conj:L2discr.S2}
If Conjectures~\ref{conj:Riesz.s.energy.Sd} and \ref{conj:C.s.2} hold, then
\begin{equation*}
\ELLtwoDISCR( \mathbb{S}^2; N ) \sim A_2 \, N^{-1/2 - 1/(2d)} + \cdots \qquad \text{as $N \to \infty$,}
\end{equation*}
where
\begin{equation*}
A_2
= \sqrt{ \frac{3}{2} \left( \frac{8 \pi}{\sqrt{3}} \right)^{1/2} \left[ - \zetafcn(-1/2) \right] \DirichletL_{-3}(-1/2) }
= 0.44679728350408\dots
\end{equation*}
\end{conj}
\begin{proof}[Justification]
Since $\mathcal{H}_d( \mathbb{S}^d ) = \omega_2 = 4 \pi$ and (by \eqref{eq:ratio})
\begin{equation*}
\frac{\mathcal{H}_d(\mathbb{B}^d)}{\mathcal{H}_d(\mathbb{S}^d)} \Bigg|_{d=2} = \frac{1}{2} \, \frac{\gammafcn(3/2)}{\sqrt{\pi} \gammafcn(1)} = \frac{1}{4},
\end{equation*}
one obtains (using Conjecture~\ref{conj:C.s.2} and relations \eqref{zeta.lambda.prod} and \eqref{eq:A.d})
\begin{align*}
A_2
&= \sqrt{\frac{1}{4} \, \frac{-C_{-1,2}}{\left( 4 \pi \right)^{-1/2}} } = \sqrt{\frac{1}{4} \left( \sqrt{3} / 2 \right)^{-1/2} \frac{-6 \zetafcn( -1 / 2 ) \DirichletL_{-3}( -1 / 2 )}{\left( 4 \pi \right)^{-1/2}} } \\
&= \sqrt{ \frac{3}{2} \left( \frac{8 \pi}{\sqrt{3}} \right)^{1/2} \left[ - \zetafcn(-1/2) \right] \DirichletL_{-3}(-1/2) } = 0.44679728350408\dots
\end{align*}
\end{proof}
\subsection*{The case $d=1$}
Stolarsky' invariance principle also holds for $d=1$ as one can see from the proof given in \cite{BrDi2011_pre}. In this case one has (cf. Eq.s~\eqref{eq:V.s.S.d} and \eqref{eq:ratio})
\begin{equation*}
\frac{1}{N^2} \sum_{j, k = 1}^N \left| \PT{x}_j - \PT{x}_k \right| + \frac{1}{\pi} \left[ \ELLtwoDISCR( X_N ) \right]^2 = \frac{4}{\pi}, \qquad X_N \subseteq \mathbb{S}^1.
\end{equation*}
On the other hand (in joint work with Hardin and Saff)~\cite{BrHaSa2009} we obtained a complete asymptotic expansion of the Riesz $s$-energy $\mathcal{L}_{s}(N)$ of the $N$th roots of unity which represent optimal $N$-point configurations for $s > - 2$. In general, for $s \in \mathbb{C}$ with $s \neq 0, 1, 3, \dots$
\begin{equation*} \label{eq:cal.L.s.N}
\mathcal{L}_s(N) = V_s( \mathbb{S} ) \, N^2 + \frac{2\zetafcn(s)}{(2\pi)^s} N^{1+s} + \frac{2}{(2\pi)^{s}} \sum_{n=1}^{p} \alpha_n(s) \zetafcn(s-2n) \, N^{1+s-2n} + \mathcal{O}( N^{-1+ \mathop{\mathrm{Re}} s-2p} ).
\end{equation*}
The coefficients $\alpha_n(s)$, $n\geq0$, satisfy the generating function relation
\begin{equation*} \label{sinc.power.0}
\left( \frac{\sin \pi z}{\pi z} \right)^{-s} = \sum_{n=0}^\infty \alpha_n(s) z^{2n},
\quad |z|<1, \ s\in \mathbb{C}.
\end{equation*}
It follows that
\begin{equation*}
\frac{1}{\pi} \left[ \ELLtwoDISCR( X_N ) \right]^2 = \frac{2 [ - \zetafcn(-1) ]}{(2 \pi)^{-1}} \, N^{-2} + \frac{2}{(2 \pi)^{-1}} \sum_{n=1}^{p-1} \alpha_n(-1) \zetafcn(-1-2n) \, N^{-2-2n} + \mathcal{O}(N^{-2-2p}),
\end{equation*}
where ($B_0=1$, $B_1=-1/2$, \dots{} are the so-called {\em Bernoulli numbers} and $(-1)^{n+1} B_{2n}>0$)
\begin{equation*}
\alpha_n(-1) \zetafcn(-1-2n) = \frac{(-1)^{n+1} B_{2(n+1)}}{(2(n+1))!} \pi^{2n} < 0, \qquad n \geq 1.
\end{equation*}
In the last step it was used that the Riemann zeta function at negative integers can be expressed in terms of Bernoulli numbers.
\subsection*{Numeric Results}
Recently, in joint work with Josef Dick we investigated the properties of so-called digital nets (see Niederreiter~\cite{Ni1988} and Dick and Pillichshammer~\cite{DiPi2010}) lifted to the sphere by means of an area preserving map. Numerical experiments for Sobol' Sequences lifted to $\mathbb{S}^2$ in \cite{BrDi2011b_pre} suggest that the order of $N$ is optimal but the constant seems to be too large (in the range $[0.50,0.59]$ for $N = 2^m$, $m=1,\dots,20$). (Optimal $(-1)$-energy configurations are more difficult to come by and the number of points for which numerical optimization is feasible and the result could be trusted to be the global maximum is very limited. Cf. \cite{BaBr2009,BeCa2009,BeClDu2004} and references cited therein.)
\section{Estimating the spherical cap discrepancy}
The classical {\em Erd{\"o}s-Tur{\'a}n type inequality} (cf. Grabner~\cite{Gr1991}, also cf. Li and Vaaler~\cite{LiVa1999}) estimate the spherical cap discrepancy in terms of Weyl sums. Recently, by generalizing LeVeques result for the unit circle \cite{LeV1965}, Narcowich et al.~\cite{NaSuWa2010} obtained {\em LeVeque type inequalities} on the sphere:
\begin{equation*}
c_1 \left[ \sum_{\ell=1}^\infty a_\ell \sum_{m=1}^{Z(d,\ell)} \left| \frac{1}{N} \sum_{j=0}^{N-1} Y_{\ell,m}(\PT{x}_{j}) \right|^2 \right]^{1/2} \leq D_{\mathrm{C}}(X_{N}) \leq c_2 \left[ \sum_{\ell=1}^\infty b_\ell \sum_{m=1}^{Z(d,\ell)} \left| \frac{1}{N} \sum_{j=0}^{N-1} Y_{\ell,m}(\PT{x}_{j}) \right|^2 \right]^{1/(d+2)},
\end{equation*}
where $a_\ell {:=} \Gamma(\ell - 1/2) / \Gamma( \ell + d + 1/2) \asymp 1 / \ell^{d+1} {=:} b_\ell$ for some positive constant $c_1$ and $c_2$ and $Z(d,n)$ denotes the number of linearly independent real spherical harmonics $Y_{\ell,m}$ of degree $\ell$. In Sun and Chen~\cite{SuCh2008} 'spherical basis functions' (as a counter part to radial basis function on spheres utilized in \cite{NaSuWa2010}) are used to investigate uniform distribution on spheres.
G. Wagner obtained ground-breaking results concerning estimates of the Riesz $s$-energy and explored connections between $s$-energy and discrepancy (see \cite{Wa1990, Wa1992b, Wa1992}). In fact, the proofs of both the Erd{\"o}s-Tur{\'a}n type inequality and the LeVeque type inequality can be modified yielding estimates of the spherical cap discrepancy in terms of $s$-energy. A different approach exploits the connection between the error of numerical integration for polynomials (cf. Damelin and Grabner~\cite{DaGr2003}) and, by utilizing a result of Andrievskii, Blatt and G{\"o}tz~\cite{AnBlGo1999} (see this author~\cite{Br2008}).
We close this note by recalling that optimal Riesz $s$-energy configurations may not have optimal spherical cap discrepancy (essentially of order $\mathcal{O}(N^{-1/2-1/(2d)})$, see Beck~\cite{Be1984}) which is reflected in Korevaar's conjecture~\cite{Ko1996} claiming for the harmonic case $s=d-1$ that the spherical cap discrepancy is of order $\mathcal{O}(N^{-1/d})$. This was essentially proved by G{\"o}tz~\cite{Go2000}. He also showed that the bound is sharp for $d=2$. The conjecture is still open for other values of $s$ and $d$.
\vspace{10mm}
{\bf Acknowledgement:} The author is grateful to the School of Mathematics and Statistics at UNSW for their support.
\def\cprime{$'$} \def\polhk#1{\setbox0=\hbox{#1}{\ooalign{\hidewidth
\lower1.5ex\hbox{`}\hidewidth\crcr\unhbox0}}}
|
\section{Introduction}
Vertex connectivity is a classical topic in graph theory and motivated many problems in structural graph theory. One such problem is the study of constrained vertex separators. In particular, clique separators, stable separators, and balanced separators are the most popularly studied constrained vertex separator in the literature \cite{dirac,clique,stable,marx}. This line of study was initiated by Dirac \cite{dirac} with a structural characterization of all clique vertex separators graphs. More precisely, Dirac addressed a fundamental question of characterizing graphs in which every minimal vertex separator is a clique.\\
A graph is chordal (also known as triangulated), if there is no induced cycle of length at least 4. Alternately, every cycle of length at least 4 has a chord. In \cite{dirac} Dirac presented a structural characterization of chordal graphs with respect to vertex separators. The famous characterization says, a graph is chordal if and only if every minimal vertex separator is a clique. While chordal graphs and its structural properties have received much attention in the literature, the analogous question, characterize graphs such that every minimal vertex separator is an independent set has not received attention in the past. \\
Another motivation for the study of stable minimal vertex separator graphs is from the theory of {\em contractible edges}. In a $k$-connected graph, an edge is said to be contractible if its contraction does not decrease the connectivity below $k$. This problem and its structural study was initiated by Tutte \cite{tutte}. Subsequently, Saito \cite{saito} characterized contractible edges in 3-connected graphs and in \cite{saito} it is shown that there exists a contractible edge in every 3-connected graphs. Unlike 3-connected graphs, there are 4-connected graphs without any contractible edge and these graphs are called {\em critically} 4-connected graphs. Martinov in \cite{martinov} characterized critically 4-connected graphs. A natural question is to characterize critically $k$-connected graphs for any $k$. The following lemma relates vertex connectivity and edge contraction.
\begin{lemma}\cite{saito}
Let $G$ be a $k$-connected graph and $G.e$ denote the graph obtained from $G$ by contracting the edge $e=\{u,v\}$. $G.e$ is $k$-connected iff for any minimum vertex separator $S$ of $G$, $\{u,v\} \not\subseteq S$.
\end{lemma}
In the light of the above lemma, we pose a natural question, which is to characterize graphs such that every edge of it is contractible. Equivalently, characterize graphs in which every minimum vertex separator is a stable set. However, from the point of view of structural graph theory, it is appropriate to characterize graphs in which every minimal vertex separator is a stable set. With these two motivations the important contribution in this paper are the following:
\begin{itemize}
\item Graph $G$ does not have a cycle of length at least 4 with exactly one chord as an induced subgraph (1-chord subgraph) if and only if every minimal vertex separator in $G$ is a stable set. \\
\item We establish the fact that deciding whether a graph has a maximum 1-chord subgraph is NP-complete. \\
\item We have also looked at an analogous question in the edge connectivity setting. We show that the class of graphs in which every minimal edge separator induces a matching are precisely the class of trees.
\end{itemize}
\section{Graph Preliminaries}
\label{prelims}
Notation and definitions are as per \cite{west,golu}. Let $G =(V,E)$ be an undirected non weighted simple graph, where $V(G)$ is the set of vertices and $E(G) \subseteq \{\{u,v\}~|~ u,v \in V(G)$, $u \not= v \}$. A separating set or a vertex separator of a graph $G$ is a set $S \subseteq V(G)$ such that the induced subgraph, denoted by $G \setminus S$, on the vertex set $V(G) \setminus S$ has more than one connected component. The vertex connectivity of a graph $G$, written $\kappa(G)$, is the minimum cardinality of a vertex set $S$ (minimum vertex separator) such that $G \setminus S$ is disconnected or has only one vertex. A graph is $k$-connected if its vertex connectivity is $k$. A vertex separator $S$ is called a $a$-$b$ vertex separator iff $S$ disconnects $a$ and $b$. i.e. the graph $G \setminus S$ has at least two connected components such that $a$ and $b$ are in distinct components. A $a$-$b$ vertex separator is said to be a minimal vertex separator iff no proper subset of it is a $a$-$b$ vertex separator. A minimum $a$-$b$ vertex separator is a minimal $a$-$b$ vertex separator of least size. For $S \subset V(G)$, $G[S]$ denote the graph induced on the set $S$. For a $a$-$b$ vertex separator $S$ let $\{C_1,\ldots,C_r\}$ denote the set of connected components in $G \setminus S$. A chord of a cycle $C$ is an edge joining a pair of nonadjacent vertices in $C$. A graph is said to be chordal if every cycle of length at least 4 has a chord in it. A subgraph $H$ of $G$ is said to be {\em 1-chord} if $G[V(H)]$ is a cycle of length at least 4 with exactly one chord. $G$ is {\em 1-chord free} if $G$ contains no such $H$. An example is shown in Figure \ref{1chordfig}.
\begin{figure}
\begin{center}
\includegraphics[scale=0.6]{1-chord.pdf}
\caption{1-chord graph}\label{1chordfig}
\end{center}
\end{figure}
\section{Stable Vertex Separator graphs: Structural Characterization}
In this section, we characterize graphs in which all minimal $(a,b)$ vertex separators are stable sets. For simplicity, we use minimal vertex separator instead of minimal $(a,b)$ vertex separator and the pair $(a,b)$ will be clear from the context. The following well known lemma is a key lemma in the proof of our main theorem and we describe a proof of the same to make the paper self contained.
\begin{lemma}
\label{char-lemma}
Let $S$ be a $(a,b)$ vertex separator in $G$ and $C_1$ and $C_2$ are connected components in $G \setminus S$ such that $a \in C_1$ and $b \in C_2$. $S$ is minimal iff for each $v \in S$ there exists $u \in C_i, i \in \{1,2\}$ such that $\{u,v\} \in E(G)$.
\end{lemma}
\begin{proof}
Suppose there exists $v \in S$ for all $u \in C_1$ there is no edge $\{u,v\}$ in $G$. Consider the set $S'=S \setminus \{v\}$. Clearly in $G \setminus S'$ the pair $\{a,b\}$ are in distinct components. This implies that $S'$ is a $(a,b)$ vertex separator. Since $S' \subset S$, implies that $S$ is not a minimal $(a,b)$ vertex separator. This contradicts the given fact that $S$ is minimal $(a,b)$ vertex separator. Therefore, for each $v \in S$ there exists $u \in C_1$ such that $\{u,v\} \in E(G)$. A similar argument proves that for each $v \in S$ there exists $u \in C_2$ such that $\{u,v\} \in E(G)$. Conversely, assume there exists $S' \subset S$ in $G$ such that $S'$ is a minimal $(a,b)$ vertex separator. Let $C'_1$ and $C'_2$ are the connected components in $G \setminus S'$ such that $a \in C'_1$ and $b \in C'_2$. This implies that for each $z \in C'_1$ and for each $z' \in C'_2$ every path $P_{zz'}$ from $z$ to $z'$ contains an element from $S'$. Since $S' \subset S$ there must exist $w \in S \setminus S'$ such that $\{w,w'\} \notin E(G)$ for any $w' \in C_1$ or $\{w,w''\} \notin E(G)$ for any $w'' \in C_2$. However, this contradicts the given hypothesis that for each $v \in S$ there exists $u \in C_i, i \in \{1,2\}$ such that $\{u,v\} \in E(G)$. Therefore, the sufficiency follows. Hence the lemma. \qed
\end{proof}
\begin{theorem}
\label{1chordchar}
$G$ is 1-chord free if and only if every minimal $(a,b)$ vertex separator of $G$ is an independent set.
\end{theorem}
\begin{proof}
{\em Necessity:} Given that $G$ is 1-chord free we now show that every minimal $(a,b)$ vertex separator is an independent set. We present a proof by contradiction. Suppose there exists a minimal $(a,b)$ vertex separator $S$ such that $G[S]$ is not an independent set. This implies that there exists $u,v \in S$ such that $\{u,v\} \in E(G)$. Consider the connected components $\{C_1,\ldots,C_r\}$ in $G \setminus S$. Without loss of generality we assume that $a \in C_1$ and $b \in C_2$. Since $S$ is a minimal $(a,b)$ vertex separator, by Lemma \ref{char-lemma}, for each $w \in S$ there exists $z \in C_1$ such that $\{w,z\} \in E(G)$. Similarly there exists $z' \in C_2$ such that $\{w,z'\} \in E(G)$. Consider shortest paths $P_1=\{u,z_1,\ldots,z_k,v\}, k \geq 1, z_i \in C_1$ and $P_2=\{u,w_1,\ldots,w_l,v\}, l \geq 1, w_i \in C_2$. Since $P_1$ and $P_2$ are shortest paths between $u$ and $v$ and $\{u,v\} \in E(G)$ implies that $P_1$ and $P_2$ together form a cycle of length at least 4 with exactly one chord. In other words, $G$ contains an induced 1-chord graph as a subgraph. However, this is a contradiction to the given hypothesis. Therefore, our assumption that there exists a minimal $(a,b)$ vertex separator $S$ such that $G[S]$ is not an independent set is wrong. Hence the necessity follows. \\
{\em Sufficiency:} We prove that $G$ is 1-chord free by the method of contradiction. Assume that in $G$ there exists an induced 1-chord graph $H$ as a subgraph. We now construct a minimal $(a,b)$ vertex separator $S$ such that $G[S]$ is not an independent set. Let $V(H)=\{u,a,z_1,\ldots,z_l=v,y_1,\ldots,y_k=b\}, l \geq 1, k \geq 1$ and $E(H)=\{\{u,a\},\{a,z_1\},\{v,y_1\},\{b,u\},\{u,v\}\} \cup \{\{z_i,z_{i+1}\}, 1 \leq i \leq l-1 \} \cup \{ \{y_j,y_{j+1}\}, 1 \leq j \leq k-1\}$. Let $X=\{a\} \cup \{z_i ~|~ z_i \in V(H)\} \setminus \{v\}$ and $Y=\{b\} \cup \{y_j ~|~ y_j \in V(H)\}$. Note that any minimal vertex separator $S$ in $G$ separating $X$ and $Y$ must contain $u$ and $v$. Moreover, we observe that $S$ is also a minimal $(a,b)$ vertex separator in $G$. The reason this is true is due to the following: clearly, $u \in S$ and if $ v \notin S$ then there exists a path between $a$ and $b$ through $v$, contradicting the fact that $S$ is a vertex separator in $G$. Therefore, $S$ is a minimal $(a,b)$ vertex separator such that $\{u,v\} \subset S$. Since $\{u,v\} \in E(G)$ implies that $G[S]$ is not an independent set. A contradiction to the hypothesis. Therefore, $G$ is 1-chord free. Hence the theorem. \qed
\end{proof}
We now present two more combinatorial observations on 1-chord free graphs with respect to its vertex connectivity.
\begin{lemma}
\label{2connectobs}
Let $G$ be a non complete at least 2-connected graph. If $G$ is 1-chord free then $G$ is triangle free.
\end{lemma}
\begin{proof}
Suppose $G$ is not triangle free. Let $\{a,b,c\}$ induce a triangle in $G$. Since $G$ is at least 2-connected and non complete, there must exist a path between $b$ and $c$ avoiding $a$. Let $P_{bc}$ denote a shortest such path and $V(P_{bc})=\{b,z_1,\ldots,z_i,c\}, i \geq 1$. We now show that $G$ contains 1-chord graph as a subgraph by considering three cases. An illustration is given in Figure \ref{case}. {\em Case 1:} There is no edge $\{a,z_j\}, 1 \leq j \leq i$, for any $z_j \in V(P_{bc})$. Clearly, $\{a,b,c\}$ together with $P_{bc}$ induce a 1-chord subgraph in $G$. However, we know that $G$ is 1-chord free. A contradiction. Therefore, $G$ is triangle free. Note that for other two cases, $i \geq 2$ as $G$ is a non complete graph. {\em Case 2:} There exists an edge $\{a,z_j\}, 1 \leq j \leq i-1$, for some $z_j \in V(P_{bc})$. If $a$ is adjacent more than one $z_j$ then without loss of generality we choose the $z_j$ such that $j$ is the least. Now, the set $\{c,a,b,z_1,\ldots,z_j\}$ induces a 1-chord subgraph with $\{a,b\}$ as the chord. A contradiction in this case too. {\em Case 3:} $\{a,z_j\} \in E(G), z_j=z_i$. In this case the set $\{b,a,c,z_i\}$ is a 1-chord subgraph with $\{a,c\}$ as the chord. A contradiction. This completes our case analysis and we see that our assumption that $G$ is not triangle free is wrong. Therefore, the claim follows. \qed
\end{proof}
\begin{figure}
\begin{center}
\includegraphics[scale=0.75]{case-analysis.pdf}
\caption{An illustration for the proof of Lemma \ref{2connectobs}}\label{case}
\end{center}
\end{figure}
Note that the converse of the above lemma is not true. For example, trees are triangle free and 1-chord free, however, trees are not 2-connected. The following observation characterizes the connected components in 1-connected 1-chord free graphs.
\begin{lemma}
Let $G$ be an exactly 1-connected 1-chord free graph. For a cut-vertex $v$ of $G$, let $\{C_1,\ldots,C_r\}$ denote the connected components in $G \setminus \{v\}$. For each cut-vertex $v$, the subgraph induced on $V(C_i) \cup \{v\}$ is either a complete graph or a 1-chord free graph.
\end{lemma}
\begin{proof}
The claim follows from the contrapositive of Lemma \ref{2connectobs} and the fact that 1-chord freeness is a hereditary property. \qed
\end{proof}
The characterization in Theorem \ref{1chordchar} can be used to test whether a given graph has the property that every minimal vertex separator is an independent set. In particular, this calls for testing the existence of 1-chord subgraph in the given graph. Two of the closely related problems are finding minimum (maximum) sized 1-chord subgraph of a graph. In the next section, we show that decision version of maximum sized 1-chord subgraph is NP-complete by presenting a polynomial time reduction from Maximum induced cycle problem. This result is of independent interest and can be used in other combinatorial problems. The decision version of maximum 1-chord subgraph is given below.
\begin{center} {\em Maximum 1-chord subgraph problem}
\begin{tabular}{|p{14cm}|}
\hline \\
{\bf Instance:} Graph $G$, and an integer $l$\\
{\bf Question:} Is there a subgraph $H$ of $G$ such that $|V(H)| \geq l$ and $G[V(H)]$ is a 1-chord? \\
\hline
\end{tabular}
\end{center}
\begin{theorem}
\label{1chordredn}
Decision version of maximum 1-chord subgraph of a graph is NP-complete
\end{theorem}
\begin{proof}
{\bf 1-chord subgraph is in NP:} Given a certificate $C=(G,H,l)$, to witness the fact that this problem is in NP, we now present a deterministic polynomial time algorithm to verify the validity of $C$. Observe that 1-chord graph of size $l$ has the degree sequence $(3,3,2,\ldots,2)$ with exactly $l-2$ vertices of degree 2. Also, both degree 3 vertices are adjacent and deleting the corresponding edge between them results in a cycle. It is now clear that the above two crucial observations along with standard Depth First Search algorithm can verify whether $C$ is valid or not in time polynomial in the input size. Therefore, we conclude that 1-chord subgraph is in NP. \\
{\bf 1-chord subgraph is NP-hard:} We establish a polynomial time reduction from maximum induced cycle problem and its decision version is known to be NP-complete\cite{garey}. The decision version of the problem is described as follows:
\begin{center}{\em Maximum induced cycle problem}
\begin{tabular}{|p{14cm}|}
\hline \\
{\bf Instance:} Graph $G$, and an integer $l$\\
{\bf Question:} Is there a subgraph $H$ such that $|V(H)| \geq l$ and $G[V(H)]$ is a cycle ? \\
\hline
\end{tabular}
\end{center}
Given an instance $(G,H,k)$ of induced cycle problem, we construct an instance $(G',H',2k)$ of 1-chord subgraph as follows: $V(G')=V(G) ~ \cup ~\{ \{ v_{ij}^1,v_{ij}^2,\ldots,v_{ij}^k\} ~|~ \{v_i,v_j\} \in E(G)\}$ and $E(G')=E(G) ~ \cup ~ \{\{v_{ij}^p,v_{ij}^{p+1}\} ~|~ 1 \leq p \leq k-1\} ~\cup~ \{\{v_i,v_{ij}^1\}, \{v_{ij}^k,v_j\}\}$. An example is illustrated in Figure \ref{1chordredn}. We now show that $(G,H,k)$ has an induced cycle of size at least $k$ if and only if $(G',H',2k)$ has a 1-chord subgraph of size at least $2k$. For {\em only if} claim, $G$ contains an induced cycle $H$, $|V(H)| \geq k$. By our construction of $G'$, for any edge $e=\{v_i,v_j\} \in E(H)$ there is a path $P_{v_iv_j}$ using the vertex set $\{ v_{ij}^1,v_{ij}^2,\ldots,v_{ij}^k\}$. Clearly, in $G'$, $V(H)$ together with $\{ v_{ij}^1,v_{ij}^2,\ldots,v_{ij}^k\}$ induce a 1-chord subgraph with $e$ as the unique chord. Thus, we have constructed in $G'$, a 1-chord subgraph $H'$ such that $|V(H')| \geq 2k$. For {\em if} claim, $G'$ contains 1-chord subgraph $H'$ of size at least $2k$. Note that if such a $H'$ exists in $G'$ then either $V(H') \cap \{ v_{ij}^1,v_{ij}^2,\ldots,v_{ij}^k\} = \phi$ or $\{ v_{ij}^1,v_{ij}^2,\ldots,v_{ij}^k\} \subset V(H')$. Also, $H'$ has two induced cycles with at least one is of size at least $k$. This implies that, in either case, there exists an induced cycle $H$ in $G$ such that $|V(H)| \geq k$. Hence the claim. Note that $|V(G')|=|V(G)|+k|E(G)|$ and $|E(G')|=(k+2)|E(G)|$ and $G'$ can be constructed in $O(k|E(G)|)$ time. For non trivial cases, $k$ will always be less than or equal to $|E(G)|$, so time complexity will be $O(|E(G)|^2)$. Thus, we have established a polynomial time reduction from induced cycle problem to 1-chord subgraph problem. As a consequence, we conclude that deciding 1-chord subgraph is NP-hard. Therefore, 1-chord subgraph problem is NP-complete. Hence the theorem. \qed
\begin{figure}
\begin{center}
\includegraphics[scale=0.75]{1chordredn.pdf}
\caption{Reducing an instance of induced cycle problem to 1-chord graph problem}\label{1chordredn}
\end{center}
\end{figure}
\end{proof}
{\bf Remark:} It is interesting to compare our NP-hardness result with an open problem posed in \cite{spinrad-open} and it is the following: can we determine the longest cycle without crossing chords in polynomial time? Note that maximum induced cycle problem and maximum 1-chord subgraph are special cases of this problem and both are known to be NP-complete. With these observations we believe that the above problem may not have a polynomial time algorithm.\\
{\bf Other Observations:} As far as approximation algorithm is concerned for the above problem, there is no polynomial time approximation algorithm with approximation ratio $O(n^{1-\epsilon})$ for any $\epsilon \geq 0$. This is true due to the following observation. NP-hard reduction of induced cycle problem is from independent set problem and this reduction is an approximation ratio preserving reduction. Since independent set does not have an approximation algorithm with approximation ratio $O(n^{1-\epsilon})$ for any $\epsilon \geq 0$, it follows that induced cycle problem does not have an approximation algorithm with approximation ratio $O(n^{1-\epsilon})$ for any $\epsilon \geq 0$. Moreover, NP-hard reduction of Theorem \ref{1chordredn} is also an approximation ratio preserving reduction and hence we conclude that 1-chord subgraph problem does not have an approximation algorithm with approximation ratio $O(n^{1-\epsilon})$ for any $\epsilon \geq 0$. As far as parameterized complexity is concerned with parameter as the size of 1-chord subgraph, we observe that parameterized 1-chord subgraph is $W[1]$-hard. i.e. there is no parameterized algorithm with parameter as the size of 1-chord subgraph. More about parameterized complexity can be found in the book by Downey and Fellows \cite{downey}. The above result follows from the fact that parameterized independent set is $W[1]$-hard and NP-hard reduction of independent set to induced cycle and NP-hard reduction of Theorem \ref{1chordredn} are parameterized reduction. Therefore, parameterized 1-chord subgraph is $W[1]$-hard.
\section{Structural Characterization of Matching Edge Separators}
We shall now focus on an analogous question, which is to characterize the graph class such that any minimal edge separator induces a matching. An edge separator of a graph $G$ is a set $E' \subseteq E(G)$ such that the induced subgraph $G'$, $V(G')=V(G)$ and $E(G')=E(G)\setminus E'$ has two connected components. We denote the induced subgraph $G'$ by $G \setminus E'$. An edge separator $E'$ is called a $a$-$b$ edge separator if $E'$ disconnects $a$ and $b$. i.e. the graph $G \setminus E'$ has two connected components such that $a$ and $b$ are in distinct components. A $a$-$b$ edge separator is said to be a minimal $a$-$b$ edge separator iff no proper subset of it is a $a$-$b$ edge separator. A minimum $a$-$b$ edge separator is a minimal $a$-$b$ edge separator of least size.
\begin{theorem}
A graph $G$ is such that for all $a,b \in V(G)$, all minimal $a$-$b$ edge separators induce a matching if and only if $G$ is a tree
\end{theorem}
\begin{proof}
For {\em if claim}, since $G$ is a tree, it is a well known fact that between any pair $(a,b)$ of vertices there exists exactly one path between $a$ and $b$. This implies that any minimal $a$-$b$ edge separator contains exactly one edge, which is a matching. Hence the claim. For {\em only if claim}, we present a proof by the method of contradiction. Suppose $G$ is not a tree. This implies there exists a cycle $C$ in $G$. Now let us consider two vertices $a$ and $b$ in $C$ such that $\{a,b\} \notin E(G)$. It is a well known fact that for any two vertices in $C$ there exists two edge disjoint paths between them. In particular, this observation is true for $a$ and $b$. Let ${\cal P}=\{ P_{ab} ~|~ P_{ab}$ is a path between $a$ and $b$ not containing $a$ as an internal vertex $\}$ and $X= \{x ~|~ x \in V(P_{ab}), P_{ab} \in {\cal P}\}$. In other words, the set $X$ is the set of vertices which lie on at least one path $P_{ab}$. Consider the set $Y=N_G(a) \cap X$, that is $Y$ is set of all vertices which are adjacent to $a$ and lie on a path from $a$ to $b$ which does not contain $a$ as an internal vertex. Now we will prove that the set of edges $E'=\{\{a,x\}~|~ x \in Y \} $ form a minimal $a$-$b$ edge separator in $G$. Since $a$ and $b$ lie on $C$ it is clear that $E'$ must contain the edges $\{a,y\}$ and $\{a,z\}$ where $y,z \in V(C)$. This is true due to the fact that $\{a,y\}$ and $\{a,z\}$ lie on two edge disjoint paths between $a$ and $b$. What follows is that $|E'| \geq 2$ and $E'$ is not
a matching, because any two edges in $E'$ has $a$ as the common vertex. We now show that $E'$ is a $a$-$b$ edge separator. Suppose $E'$ is not a $a$-$b$ edge separator, then there is a path $P'_{ab}$ from $a$ to $b$ in $G$ not containing any edge from $E'$. Without loss of generality we assume that $P'_{ab}$ does not contain $a$ as an internal vertex. If $P'_{ab}$ contains $a$ as an internal vertex
then we can remove all the vertices until the last $a$ in the path and still get a path from $a$ to $b$. Let $u$ be the vertex adjacent
to $a$ in $P'_{ab}$, then clearly $u \in X$ and since $\{a,u\} \in E(G)$, $u \in Y$ as well. This implies $\{a,u\} \in E'$, which is a contradiction to our assumption that $P'_{ab}$ does not contain any edges from $E'$. Therefore, $E'$ is a $a$-$b$ edge separator. Now to prove the minimality of $E'$, suppose $E'$ is not minimal then there is at least one edge $e=\{a,x\} \in E'$ such that
$E' \setminus \{e\} $ is still a $a$-$b$ edge separator in $G$. Since $e \in E'$, we know that $x \in X$ and there is at least one path $P_{ab} \in {\cal P}$, containing $x$ as an internal vertex. Now consider the sub path $P'$ of $P_{ab}$ from $x$ to $b$. It is clear that $a\notin P'$ as $P_{ab}$ by definition does not contain $a$ as an internal vertex. So there does not exist $y \in V(G)$ such that $\{a,y\} \in E(G)$ and $\{a,y\}$ is an edge of the path $P'$. In what follows, the edge $\{a,x\}$ together with the path $P'$ yields a path from $a$ to $b$, which does not contain any edge from $E' \setminus \{e\}$, giving rise to a contradiction that $E \setminus \{e\} $ is a $a$-$b$ edge separator in $G$. Hence we conclude that $E'$ is a minimal $a$-$b$ edge separator which is not a matching. A contradiction to the hypothesis. Therefore our assumption that $G$ is not a tree is wrong. Hence the theorem. \qed
\end{proof}
\bibliographystyle{splncs}
|
\section{Introduction}
\label{sec:Intro}
The purpose of this work is to study properties of propagating
electromagnetic fields in linear medium. We will work in a
relativistic setting where Maxwell's equations are written on a
$4$-manifold and the electromagnetic medium is represented by an
antisymmetric $2\choose 2$-tensor $\kappa$. Pointwise, such medium is determined by
$36$ parameters.
To understand the propagation of an electromagnetic wave in this setting, the key object
is the \emph{Fresnel surface}, which can be seen a
generalisation of the light-cone
\cite{Rubilar2002, Obu:2003, PunziEtAl:2009}.
For a Lorentz metric, the light-cone is always a polynomial surface of second order
in each cotangent space.
The Fresnel surface, in turn, is a polynomial
surface of fourth order. For example, the
Fresnel surface can be the union of two light-cones. This allows the
Fresnel surface to model propagation also in birefringent medium. That
is, in medium where differently polarised electromagnetic waves can
propagate with different wave speeds.
The Fresnel surface is determined by the \emph{Tamm-Rubilar tensor
density} which is a symmetric $4\choose 0$-tensor density, which,
in turn, is determined by the medium $2\choose 2$-tensor $\kappa$.
This dependence is illustrated in the diagram below:
\begin{eqnarray*}
\label{ill:twoArrow}
\quad\mbox{Medium}\,\, \kappa\quad\to\quad \mbox{Tamm-Rubilar tensor density}\quad\to\quad \mbox{Fresnel surface}.
\end{eqnarray*}
In Lorentz geometry, we know the the light cone of a Lorentz metric
$g$ uniquely determine $g$ up to a conformal factor
\cite{Ehrlich:1991}. In this work we will study the analogue relation
between a general electromagnetic medium tensor $\kappa$ and its
Fresnel surface; Can one reconstruct an electromagnetic medium from
its Fresnel surface? In general, a unique reconstruction is not
possible. For example, the Fresnel surface is invariant under a
conformal change in the medium. Hence the Fresnel surface can, at
best, determine $\kappa$ up to a conformal factor. One would then
like to understand the following question:
\begin{question}
\label{mainProblem}
Under what assumptions does the Fresnel surface at a point $p\in N$
determine the electromagnetic medium $\kappa\vert_p$ up to a conformal
factor?
\end{question}
In terms of physics, Question \ref{mainProblem} asks when we can
reconstruct $\kappa\vert_p$ (up to a conformal factor) using only
wavespeed information about the medium at $p$. A proper understanding
of this question is not only of theoretical interest, but also of
interest in engineering applications like electromagnetic tomography.
Question \ref{mainProblem} is also similar is spirit to a question in
general relativity, where one would like to understand when the the
conformal class of a Lorentz metric can be determined by the five
dimensional manifold of null-geodesics \cite{Low:2005}.
Favaro and Bergamin have recently proven the following result of
positive nature \cite{FavaroBergamin:2011}: If $\kappa$ has only a
principal part and if the Fresnel surface of $\kappa$ coincides with the
light cone for a Lorentz metric $g$, then $\kappa$ is proportional to
the Hodge star operator of $g$. That is, in a restricted class of medium,
the Fresnel surface of $\kappa$ determines the conformal class of
$\kappa$. An important corollary
is the following: If $\kappa$ has only a principal part and its
Fresnel surface coincides with the light cone for a Lorentz metric,
then $\kappa$ satisfies the \emph{closure condition} $\kappa^2=-f
\operatorname{Id}$ for a function $f\colon N\to (0,\infty)$.
This resolves a conjecture on whether the closure condition
characterises non-birefringent medium in skewon-free medium
\cite{ObuFukRub:00, Obu:2003}. That the closure condition is sufficient, was
already proven in \cite{ObukhovHehl:1999, ObuFukRub:00}, but before
\cite{FavaroBergamin:2011} sufficiency was only known under additional
assumptions; a proof assuming that $\cC=0$ (see Section
\ref{sec:ABCDtransRules} for the definition of $\cC$ in terms on
$\kappa$) is given in \cite{ObuFukRub:00}, and a proof in a special
class of non-linear medium is given in \cite{ObuRub:2002}.
For additional positive results to Question \ref{mainProblem}, see
\cite{LamHeh:2004, Itin:2005, Schuller:2010, FavaroBergamin:2011}.
The main contribution of this paper is twofold. First, we give a new
proof of the result quoted above from \cite{FavaroBergamin:2011}.
This is formulated as implication \ref{coIII} $\Rightarrow$ \ref{coII}
in Theorem \ref{thm:mainResult}. While the original proof in
\cite{FavaroBergamin:2011} relies on the classification of skewon-free
$2\choose 2$-tensors into 23 normal forms by Schuller, Witte, and
Wohlfarth \cite{Schuller:2010}, we will use Gr\"obner bases to prove
Theorem \ref{thm:mainResult}. Essentially, Gr\"obner bases is a
computer algebra technique for simplifying a system polynomial
equations without changing the solution set. See Appendix
\ref{app:Groebner}.
The second contribution of this paper is given in Section
\ref{sec:uni} which contains a number of cases, where the Fresnel
surface does not determine $\kappa$. In Theorem \ref{thm:FkappaInvB}
\ref{thm:FkappaInvB:iv} we show that if $\kappa$ is invertible, then
$\kappa$ and $\kappa^{-1}$ have the same Fresnel surfaces. Also, in
Example \ref{ex:complexExample} we construct a $\kappa$ with complex
coefficients on $\setR^4$. At each $p\in \setR^4$, this medium is
determined by one arbitrary complex number, and hence the medium can
depend on both time and space. However, at each point, the Fresnel
surface of $\kappa$ coincides with the usual light cone of the flat
Minkowski metric $g=\operatorname{diag}(-1,1,1,1)$.
The paper is organised as follows. In Section \ref{mainSec} we review
Maxwell's equations and linear electromagnetic medium on a
$4$-manifold. In Section \ref{sec:GOS} we describe how the
Tamm-Rubilar tensor density and Fresnel surface is related to wave
propagation. To derive these objects we use the approach of geometric
optics.
As described in Section \ref{sec:GOS}, this can be seen as a step
towards a relativistic theory of electromagnetic Gaussian beams (if
such a theory exists). In general, Gaussian beams is an asymptotic
technique for studying propagation of waves in hyperbolic systems.
These solutions behave as wave packets; at each time instant, the
entire energy of the solution is concentrated around one point in
space. When time moves forward, the beam propagates along a curve, but
always retains its shape of a Gaussian bell curve. Electromagnetic
Gaussian beams are also known as quasi-photons \cite{Kachalov:2002,
Kachalov:2004, Kachalov:2005, DahlPIER:2006}. For the wave
equation, see \cite{Ralston:1982, KKL:2001}. For the history of
Gaussian beams, see \cite{Ralston:1982, Popov:2002}.
In Section \ref{sec:Closure} we prove the main result Theorem
\ref{thm:mainResult}, and in Section \ref{sec:uni} we describe a
number of cases where Question \ref{mainProblem} has a negative
answer.
This paper relies on a number of computations done with computer algebra.
Further information about these can be found on the author's homepage.
\section{Maxwell's equations}
\label{mainSec}
By a \emph{manifold} $M$ we mean a second countable topological Hausdorff
space that is locally homeomorphic to $\setR^n$ with $C^\infty$-smooth
transition maps. All objects are assumed to be smooth where defined.
Let $TM$ and $T^\ast M$ be the tangent and cotangent bundles,
respectively, and for $k\ge 1$, let $\Lambda^k(M)$ be the set of
$p$-covectors, so that $\Lambda^1(N)=T^\ast N$. Let $\Omega^k_l(M)$
be $k\choose l$-tensors that are antisymmetric in their $k$ upper
indices and $l$ lower indices. In particular, let $\Omega^k(M)$ be the
set of $k$-forms. Let also $\vfield{M}$ be the set of vector fields,
and let $C^\infty(M)$ be the set of functions. By $\Omega^k(M)\times
\setR$ we denote the set of $k$-forms that depend smoothly on a
parameter $t\in \setR$.
By $T(M,\setC)$, $T^\ast(M,\setC)$, $\Lambda^p(M,\setC)$,
$\Omega^k_l(M,\setC)$ and $\vfield{M,\setC}$ we denote the
complexification of the above spaces where component may also
take complex values. Smooth complex valued functions are denoted
by $C^\infty(M,\setC)$.
The Einstein summing convention is used throughout. When writing tensors
in local coordinates we assume that the components satisfy the same symmetries as
the tensor.
We will use differential forms to write Maxwell's equations. On a
$3$-manifold $M$, \emph{Maxwell equations} then read \cite{BH1996,
Obu:2003}
\begin{eqnarray}
\label{max1}
dE &=& - \pd{B}{t} , \\
\label{max2}
dH &=& \pd{D}{t} + J, \\
\label{max3}
dD &=& \rho, \\
\label{max4}
dB &=& 0,
\end{eqnarray}
for field quantities $E,H\in \Omega^1 (M)\times \setR$, $D,B\in
\Omega^2 (M)\times \setR$ and sources $J\in \Omega^2(M)\times \setR$
and $\rho \in \Omega^3(M)\times \setR$. Let us emphasise that
equations \eqref{max1}--\eqref{max4} are completely
differential-topological and do not depend on any additional
structure. (To be precise, the exterior derivative does depend on the
smooth structure of $M$. However, for a manifold $M$ of dimension
$1,2,3$ one can show that all smooth structures for $M$ are
diffeomorphic.
For higher dimensions the analogue result is not true. Even for
$\setR^4$ there are uncountably many non-diffeomorphic smooth
structures \cite[p.~255]{Scorpan:2005}.)
\subsection{Maxwell's equations on a $4$-manifold}
\label{sec:MaxOn4}
Suppose $E,D,B,H$ are time dependent forms $E,H\in \Omega^1(M)\times
\setR$ and $D,B\in \Omega^2(M)\times \setR$ and $N$ is the
$4$-manifold $N=\setR\times M$. Then we can define forms $F,G \in
\Omega^2(N)$ and $j\in \Omega^3(N)$,
\begin{eqnarray}
\label{Fdef}
F &=& B + E\wedge dt, \\
\label{Gdef}
G &=& D - H\wedge dt,\\
j &=& \rho-J\wedge dt.
\end{eqnarray}
Now fields $E,D,B,H$ solve Maxwell's equations equations
\eqref{max1}--\eqref{max4} if and only if
\begin{eqnarray}
\label{max4A}
dF &=& 0, \\
\label{max4B}
dG &=& j,
\end{eqnarray}
where $d$ is the exterior derivative on $N$. More generally, if $N$ is
a $4$-manifold and $F,G, j$ are forms $F,G\in \Omega^2(N)$ and $j\in
\Omega^3(N)$ we say that $F,G$ solve \emph{Maxwell's equations} (for
a source $j$) when equations \eqref{max4A}--\eqref{max4B} hold.
By an \emph{electromagnetic medium} on $N$
we mean a map
\begin{eqnarray*}
\kappa \colon \Omega^2(N) &\to& \Omega^2(N).
\end{eqnarray*}
We then say that $2$-forms $F,G\in \Omega^2(N)$ \emph{solve Maxwell's
equations in medium $\kappa$} if $F$ and $G$ satisfy equations
\eqref{max4A}--\eqref{max4B} and
\begin{eqnarray}
\label{FGchi}
G &=& \kappa(F).
\end{eqnarray}
Equation \eqref{FGchi} is known as the \emph{constitutive equation}.
If $\kappa$ is invertible,
it follows that one can eliminate half of the free variables in
Maxwell's equations \eqref{max4A}--\eqref{max4B}.
We assume that $\kappa$ is linear and local so that we can represent
$\kappa$ by an antisymmetric $2\choose 2$-tensor $\kappa \in
\Omega^2_2(N)$. If in coordinates $\{x^i\}_{i=0}^3$ for $N$ we have
\begin{eqnarray}
\label{eq:kappaLocal}
\kappa &=& \frac 1 2 \kappa^{ij}_{lm} dx^l\otimes dx^m\otimes \pd{}{x^i}\otimes \pd{}{x^j}
\end{eqnarray}
and $F = F_{ij} dx^i \otimes dx^j$ and $G = G_{ij} dx^i \otimes dx^j$,
then constitutive equation \eqref{FGchi} reads
\begin{eqnarray}
\label{FGeq_loc}
G_{ij} &=& \frac 1 2 \kappa_{ij}^{rs} F_{rs}.
\end{eqnarray}
\subsection{Decomposition of electromagnetic medium}
\label{media:decomp}
Let $N$ be a $4$-manifold. Then
at each point on $N$, a general antisymmetric $2\choose
2$-tensor depends on $36$ parameters. Such tensors canonically decompose
into three linear subspaces. The motivation for this decomposition is
that different components in the decomposition enter in different
parts of electromagnetics. See \cite[Section D.1.3]{Obu:2003}.
The below formulation is taken from \cite{Dahl:2009}.
If $\kappa\in \Omega^2_2(N)$ we define the
\emph{trace} of $\kappa$ as the smooth function $N\to \setR$ given by
\begin{eqnarray*}
\operatorname{trace} \kappa &=& \frac 1 2 \kappa_{ij}^{ij}
\end{eqnarray*}
when $\kappa$ is locally given by equation \eqref{eq:kappaLocal}.
Writing
$\operatorname{Id}$ as in equation \eqref{eq:kappaLocal} gives
$\operatorname{Id}^{ij}_{rs}= \delta^i_r\delta^j_s-\delta^i_s\delta^j_r$, so
$\operatorname{trace}\operatorname{Id} = 6$ when $\dim N=4$.
\begin{proposition}[Decomposition of a $2\choose 2$-tensors]
\label{theorem:Decomp}
Let $N$ be a $4$-manifold, and let
\begin{eqnarray*}
Z &=& \{ \kappa \in \Omega^2_2(N) : u\wedge \kappa(v) = \kappa(u)\wedge v \,\,\mbox{for all}\,\, u,v\in \Omega^2(N),\\
& & \quad\quad\quad\quad\quad\quad \operatorname{trace} \kappa = 0\},\\
W &=& \{ \kappa \in \Omega^2_2(N) : u\wedge \kappa(v) = -\kappa(u)\wedge v \,\,\mbox{for all}\,\, u,v\in \Omega^2(N)\} \\
&=& \{ \kappa \in \Omega^2_2(N) :
u\wedge \kappa(v) = -\kappa(u)\wedge v \,\,\mbox{for all}\,\, u,v\in \Omega^2(N), \\
& & \quad\quad\quad\quad\quad\quad \operatorname{trace} \kappa = 0 \}, \\
U &=& \{ f \operatorname{Id}\in \Omega^2_2(N) : f\in C^\infty(N) \}.
\end{eqnarray*}
Then
\begin{eqnarray}
\label{AdecompSet}
\Omega^2_2(N) &=& Z\,\,\oplus\,\, W \,\,\oplus\,\, U,
\end{eqnarray}
and pointwise, $\dim Z = 20$, $\dim W = 15$ and $\dim U = 1$.
\end{proposition}
If we write a $\kappa\in \Omega^2_2(N)$ as
\begin{eqnarray*}
\kappa &=& \kappaI \,\,+ \,\,\kappaII\,\,+\,\,\kappaIII
\end{eqnarray*}
with $\kappaI\in Z$, $\kappaII\in W$, $\kappaIII\in U$, then we say that
$\kappaI$ is the \emph{principal part},
$\kappaII$ is the \emph{skewon part},
$\kappaIII$ is the \emph{axion part} of $\kappa$.
\subsection{The Hodge star operator}
\label{sec:Hodge}
By a \emph{pseudo-Riemann metric} on a manifold $M$ we mean a symmetric real
$0\choose 2$-tensor $g$ that is non-degenerate. If $M$ is not
connected we also assume that $g$ has constant signature.
If $g$ is positive definite, we say that $g$ is a \emph{Riemann metric}.
By $\sharp$
and $\flat$ we denote the isomorphisms $\sharp\colon T^\ast M\to TM$
and $\flat\colon TM\to T^\ast M$. By $\setR$-linearity we extend $g$,
$\sharp$ and $\flat$ to complex arguments. Moreover, we extend $g$
also to covectors by setting $g(\xi,\eta)=g(\xi^\sharp,\eta^\sharp)$
when $\xi,\eta\in \Lambda^1_p(N,\setC)$.
Suppose $g$ is a pseudo-Riemann metric on a
orientable manifold $M$ with $n=\dim M\ge 2$. For $p\in\{0,\ldots,
n\}$, the \emph{Hodge star operator} $\ast$ is the map $\ast\colon
\Omega^p(M)\to \Omega^{n-p}(M)$ defined as
\cite[p. 413]{AbrahamMarsdenRatiu:1988}
\begin{eqnarray*}
\label{hodgedef}
\ast(dx^{i_1} \wedge \cdots \wedge dx^{i_p}) &=& \frac{\sqrt{|\det g|}}{(n-p)!} g^{i_1 l_1}\cdots g^{i_p l_p} \varepsilon_{l_1 \cdots l_p\, l_{p+1} \cdots l_n} dx^{l_{p+1}}\wedge \cdots \wedge dx^{l_{n}},
\end{eqnarray*}
where $x^i$ are local coordinates in an oriented atlas,
$g=g_{ij}dx^i\otimes dx^j$, $\det g= \det g_{ij} $, $g^{ij}$ is the
$ij$th entry of $(g_{ij})^{-1}$, and $\varepsilon_{l_1\cdots l_n}$ is
the \emph{Levi-Civita permutation symbol}. We treat
$\varepsilon_{l_1\cdots l_n}$ as a purely combinatorial object (and
not as a tensor density). We also define $\varepsilon^{l_1\cdots l_n}=
\varepsilon_{l_1\cdots l_n}$.
If $g$ is a pseudo-Riemann metric on an oriented
$4$-manifold $N$, then the Hodge star operator for $g$ induces a
$2\choose 2$-tensor $\kappa=\ast_g\in \Omega^2_2(N)$. If $\kappa$ is
written as in equation \eqref{eq:kappaLocal} for local coordinates
$x^i$ then
\begin{eqnarray}
\label{eq:hodgeKappaLocal}
\kappa^{ij}_{rs} &=& \sqrt{\vert g\vert} g^{ia}g^{jb} \varepsilon_{abrs}.
\end{eqnarray}
\begin{proposition}
\label{prop:HodgeHasOnlyPrincipalPart}
Suppose $g$ is a pseudo-Riemann metric on an orientable $4$-manifold $N$.
Then $\ast_g$ defines a $2\choose 2$-tensor with only a principal part.
\end{proposition}
\begin{proof} Let $\kappa$ be the $2\choose 2$-tensor induced by $\ast_g$.
Then $u\wedge \kappa (v) = \kappa(u)\wedge v$ for
all $u,v\in \Omega^2(N)$ \cite[p. 412]{AbrahamMarsdenRatiu:1988}. By Theorem \ref{theorem:Decomp}
it therefore suffices to prove that $\operatorname{trace} \kappa =
0$. Let us fix $p\in N$ and let $x^i$ are local coordinates for $N$ such
that $g\vert_p$ is diagonal. If $\kappa$ is written as in equation
\eqref{eq:kappaLocal} then equation \eqref{eq:hodgeKappaLocal} implies that
$\operatorname{trace}\kappa=\frac 1 2 \kappa^{ij}_{ij}=0$ since $g^{ij}$ is diagonal and
$\varepsilon_{ijkl}$ is non-zero only when $ijkl$ are distinct.
\end{proof}
A pseudo-Riemann metric $g$ is a \emph{Lorentz metric} if $M$ is
$4$-dimensional and $g$ has signature $(+---)$ or $(-+++)$.
For a Lorentz metric, we define the \emph{null cone} at $p$ as the set
$
\{\xi\in \Lambda^1_p(M,\setR) : g(\xi,\xi)=0\}.
$
Usually, the null cone is defined as a subset in the tangent bundle. The
motivation for treating the null-cone in the cotangent bundle is given by
equation \eqref{eq:FresIsNullCone}.
The next theorem shows that the conformal class of a Lorentz metric $g$ can
be represented either using the $2\choose 2$-tensor $\ast_g$ or the
null cone of $g$.
\begin{theorem}
\label{prop:Principle}
Suppose $g,h$ are Lorentz metrics on an orientable $4$-manifold $N$. Then
the following are equivalent:
\begin{enumerate}
\item
\label{eq:xx1}
There exists a non-vanishing function $\lambda\in C^\infty(N)$ such that $h=\lambda g$.
\item
\label{eq:xx2}
$\ast_g = \ast_h$, where $\ast_g$ and $\ast_h$ are the $2\choose 2$-tensors defined by $g$ and $h$, respectively.
\item
\label{eq:xx3}
$g$ and $h$ have the same null cones.
\end{enumerate}
\end{theorem}
\begin{proof}
Implications \ref{eq:xx1} $\Rightarrow$ \ref{eq:xx2} and
\ref{eq:xx1} $\Rightarrow$ \ref{eq:xx3} are clear. Implication
\ref{eq:xx2} $\Rightarrow$ \ref{eq:xx1} is proven in
\cite[Theorem 1]{Dray:1989}, and implication \ref{eq:xx3} $\Rightarrow$
\ref{eq:xx1} is proven in \cite[Theorem 3]{Ehrlich:1991}.
See also \cite{MingSan:2008}.
\end{proof}
\subsection{Decomposition of $\kappa$ into four $3\times 3$ matrices}
\label{sec:ABCDtransRules}
Suppose $(x^0, x^1, x^2, x^3)$ are local coordinates for
$N=\setR\times M$ such that $x^0$ is the coordinate for $\setR$ and
$(x^1, x^2, x^3)$ are coordinates for $M$. If forms $F,G$ are given by
equations \eqref{Fdef}--\eqref{Gdef}, then
$$
F_{i0} = E_i, \quad
F_{ij} = B_{ij}, \quad
G_{i0} = -H_i, \quad
G_{ij} = D_{ij}
$$
for all $i,j=1,2,3$ and equation \eqref{FGeq_loc} then reads
\begin{eqnarray}
\label{eq:kappaLocal_I}
H_i &=& -\kappa_{i0}^{r0} E_r - \frac 1 2 \kappa_{i0}^{rs} B_{rs}, \\
\label{eq:kappaLocal_II}
D_{ij} &=& \kappa_{ij}^{r0} E_r +\frac 1 2 \kappa_{ij}^{rs} B_{rs},
\end{eqnarray}
where $i,j=1,2,3$ and $r,s$ are summed over $1,2,3$.
Next we show that in coordinates $(x^0, x^1,
x^2, x^3)$ the tensor $\kappa$ is represented by four $3\times
3$-matrices. To do this, let $\ast$ is the Hodge star operator
induced by the Euclidean metric on $x^1, x^2, x^3$ so that $\ast dx^i
= \frac 1 2 \sum_{a,b=1}^3 \varepsilon^{iab} dx^a \wedge dx^b$.
Thus $B=
\sum_{i=1}^3 B^i \ast\!dx^i$
where $B^i =\frac 1 2 \varepsilon^{ijk} B_{jk}$ and $B_{mn} =
\varepsilon_{imn} B^i$. In the same way we define $D^1, D^2, D^3$. Now
components $D^i$ and $B^i$ represent $2$-forms $D$ and $B$ in the
basis $\{\ast dx^i \}_{i=1}^3$, and by equations
\eqref{eq:kappaLocal_I}--\eqref{eq:kappaLocal_II},
\begin{eqnarray}
\label{eq:kappaLocal_Ix}
H_i &=& \cC^r\, _i (-E_r) + \cB_{ri} B^{r}, \\
\label{eq:kappaLocal_IIx}
D^{i} &=& \cA^{ri} (-E_r) +\cD_r\,^i B^{r},
\end{eqnarray}
where $i\in \{1,2,3\}$, $r$ is summed over $1,2,3$, and
$$
\cC^r\,_i = \kappa^{r0}_{i0}, \quad
\cB_{ri} = - \frac 1 2 \varepsilon_{rab} \kappa^{ab}_{i0}, \quad
\cA^{ri} = -\frac 1 2 \varepsilon^{iab} \kappa^{r0}_{ab}, \quad
\cD_r\,\!^i = \frac 1 4 \varepsilon_{rmn} \varepsilon^{iab} \kappa^{mn}_{ab}.
$$
Here $r$ index rows and $i$ index columns in $3\times 3$ matrices
$\cA, \cB, \cC, \cD$.
Inverting the relations gives
$$
\kappa^{0r}_{0i} = \cC^r\,\! _i, \quad
\kappa^{ij}_{0r} = \varepsilon^{kij} \cB_{kr}, \quad
\kappa^{0i}_{rs} = \varepsilon_{krs} \cA^{ik}, \quad
\kappa^{ij}_{rs} = \varepsilon_{krs} \varepsilon^{lij} \cD_l\,\!^k,
$$
where $i,j,r,s\in \{1,2,3\}$ and $k, l$ are summed over $1,2,3$.
The above matrices $\cA, \cB, \cC, \cD$ coincide with the matrices
$\cA, \cB, \cC, \cD$ defined in \cite[Section D.1.6]{Obu:2003} and
\cite{Rubilar2002}. Since these matrices are only part of tensor
$\kappa$, they do not transform in a simple way under a general
coordinate transformation in $N$ (see equations D.5.28--D.5.30 in
\cite{Obu:2003}). However, if $\{x^i\}_{i=0}^3$ and $\{\widetilde
x^i\}_{i=0}^3$ are overlapping coordinates such that
\begin{eqnarray*}
\widetilde x^0 &=& x^0,\\
\widetilde x^i &=& \widetilde x^i(x^1, x^2, x^3), \quad i\in \{1,2,3\}.
\end{eqnarray*}
Then we have transformation rules
\begin{eqnarray}
\widetilde \cC^r\,_i &=&
\label{eq:transRuleC}
\cC^a\,_b \pd{ x^b}{\widetilde x^i}\pd{\widetilde x^r}{ x^a}, \\
\widetilde \cB_{ri} &=&
\label{eq:transRuleB}
\det\left( \pd{\widetilde x^m}{ x^n}\right) \cB_{ab} \pd{x^a}{\widetilde x^r}\pd{x^b}{\widetilde x^i}, \\
{\widetilde \cA}^{ri} &=&
\label{eq:transRuleA}
\det\left( \pd{ x^m}{ \widetilde x^n}\right) \cA^{ab} \pd{\widetilde x^r}{ x^a}\pd{\widetilde x^i}{ x^b}, \\
\widetilde \cD_r\,\!^i &=&
\label{eq:transRuleD}
\cD_a\,\!^b \pd{\widetilde x^i}{ x^b}\pd{ x^a}{\widetilde x^r}.
\end{eqnarray}
If $\kappaII=0$ then Proposition \ref{theorem:Decomp} implies that
$\kappa$ is pointwise determined by $21$ coefficients. The next
proposition shows that these coefficients can pointwise be reduced to
$18$ when the coordinates are chosen suitably.
\begin{proposition}
\label{prop:localReduction}
Suppose $N$ is a $4$-manifold and $\kappa\in \Omega^2_2(N)$.
Then
\begin{enumerate}
\item
\label{cc:AA}
$\kappa$ has no skewon component if and only if locally
$$
\cA = \cA^T, \quad
\cB = \cB^T, \quad
\cC=\cD^T,
$$
where $^T$ is the matrix transpose, and $\cA, \cB, \cC, \cD$ are defined as above.
\item
\label{cc:BB}
Let $p\in N$. If $\kappa$ has no skewon component, then there are
local coordinates around $p$ such that $\cA$ is diagonal at $p$.
\item
\label{cc:CC}
Let $p\in N$. If $\kappa$ has no skewon component and $g$ is a Lorentz
metric on $N$ there are local coordinates around $p$ such that
$\cA\vert_p$ is diagonal and for some $k\in \{\pm 1\}$ we have
$g\vert_p = k\operatorname{diag}(-1,1,1,1)$.
\end{enumerate}
\end{proposition}
\begin{proof}
Part \ref{cc:AA} follows by \cite[Equation
D.1.100]{Obu:2003}.
Since we can always introduce a Lorentz metric in local coordinates for $N$, part
\ref{cc:BB} will follow from part \ref{cc:CC}.
For part \ref{cc:CC}, let $x^i$ be coordinates around $p$ such that
$g\vert_p = k\operatorname{diag}(-1,1,1,1)$ for $k\in \{\pm 1\}$. By
\ref{cc:AA}, matrix $\cA\vert_p$ is symmetric, so we can find an
orthogonal $3\times 3$ matrix $P=(P^i\,\,_j)_{ij}$ such that $P \cA
P^T$ is diagonal and $\det P=1$. A suitable coordinate system is
given by $\widetilde x^0 = x^0$ and $\widetilde x^i =\sum_{j=1}^3
P^i\,_j x^j.$
\proofread{\textbf{Theorem:} If $A$ is a symmetric square matrix, then there exists a
orthogonal matrix $Q$ such that $A = Q^T D Q = Q^{-1} D Q$ where $D$ is the diagonal matrix with
the eigenvalues of $A$ on the diagonal.
}
\end{proof}
\section{Geometric optics solutions}
\label{sec:GOS}
Let $\kappa\in \Omega^2_2(N)$ on a $4$-manifold $N$, and let $F$ and $G$ be asymptotic sums
\begin{eqnarray}
\label{eq:FGtrial}
F = \REAL\left\{e^{iP \Phi} \sum_{k=0}^\infty \frac{A_k}{(iP)^k} \right\},\quad
G = \REAL\left\{e^{iP \Phi} \sum_{k=0}^\infty \frac{B_k}{(iP)^k} \right\},
\end{eqnarray}
where $P>0$ is a constant, $\Phi \in C^\infty(N,\setC)$ and $A_k, B_k \in \Omega^2(N,\setC)$.
Substituting $F$ and $G$ into the sourceless Maxwell equations and differentiating termwise shows that
$F$ and $G$ form an asymptotic solution provided that
\begin{eqnarray}
\label{eq:As1}
d\Phi \wedge A_0 &=&0,\\
\label{eq:As2}
d\Phi \wedge B_0 &=&0,\\
\label{eq:kConst}
B_k &=& \kappa A_k, \\
\label{eq:Trans1}
d\Phi \wedge A_{k+1} + dA_{k} &=&0,\\
\label{eq:Trans2}
d\Phi \wedge B_{k+1} + dB_{k} &=&0,
\quad k=0,1,\ldots.
\end{eqnarray}
In equation \eqref{eq:kConst} we treat $\kappa$ as a linear map $\kappa\colon \Omega^2(N,\setC)\to\Omega^2(N,\setC)$.
In equation \eqref{eq:FGtrial} function $\Phi$ is called a \emph{phase function}, and
forms $A_k, B_k$ are called \emph{amplitudes}.
We will assume that $\IMAG \Phi\ge 0$, so
that $F$ and $G$ remain
bounded even if we take $P\to \infty$.
\begin{lemma}
\label{lemma:solv}
Suppose $N$ is a smooth manifold, and let $q$ be a $1$-form $q\in
\Omega^1(N,\setC)$ that is nowhere zero.
\begin{enumerate}
\item
\label{lemmaI}
If $q\wedge A = 0$ for some $A\in \Omega^k(N,\setC)$ where $k\ge 1$,
then there exists a $(k-1)$-form $a\in \Omega^{k-1}(N, \setC)$ such that $A =
q\wedge a$.
\item
\label{lemmaII}
If $q\wedge a = q\wedge a'$ for some $a,a' \in \Omega^1(N, \setC)$, then $a = a' + f q$
for some $f\in C^\infty(N,\setC)$.
\end{enumerate}
\end{lemma}
\begin{proof}
Let $\sharp$ be the isomorphism $T^\ast N\to TN$ induced by an
auxiliary (positive definite) Riemann metric on $N$, and let $\Vert
\cdot \Vert$ be the induced norms on $TN$ and $T^\ast N$. Let also $q = \alpha +
i\beta$, where $\alpha = \REAL q$ and $\beta = \IMAG q$. Then vector
field $X\in \vfield{N, \setC}$ given by
\begin{eqnarray*}
X &=&\frac{ \alpha^\sharp - i \beta^\sharp}{ \Vert \alpha\Vert^2 + \Vert \beta\Vert^2}
\end{eqnarray*}
satisfies $q(X)=1$. Contracting $q\wedge A=0$ by $X$
gives part \ref{lemmaI}.
Part \ref{lemmaII} follows by taking $A=a-a'$ in part \ref{lemmaI}.
\end{proof}
In this work we will only analyse the leading amplitudes $A_0$ and
$B_0$. However, since $B_0 = \kappa(A_0)$, it suffices to study $A_0$
in more detail. Let us assume that $\Phi$, $A_0$ and $B_0$ solve
equations \eqref{eq:As1}--\eqref{eq:kConst}. Then Lemma
\ref{lemma:solv} \ref{lemmaI} implies that there exists a $1$-form
$a_0 \in \Omega^1(N,\setC)$ such that $ A_0 = d\Phi\wedge a_0, $
whence
\begin{eqnarray}
\label{eq:NeqA0xx}
d\Phi \wedge \kappa( d\Phi \wedge a_0) &=& 0.
\end{eqnarray}
For $N=\setR\times M$ where $M$ is a $3$-manifold and for special
choices for $\kappa$, $\Phi$ and amplitudes $A_k, B_k$, equation
\eqref{eq:FGtrial} define an electromagnetic \emph{Gaussian beam} (see
Section \ref {sec:Intro}).
In this setting, $\Phi\vert_p$ and $d\Phi\vert_p$ are both
real when $p$ is at a centre of a Gaussian beam.
With the above as motivation we will hereafter only study equation
\eqref{eq:NeqA0xx} at a point $p\in N$ where $d\Phi$ is real. From
equation \eqref{eq:FGtrial} we then see that $d\Phi\vert_p$
is the direction of most rapid oscillation
(or direction of propagation) for $F$. Since $A_0 = d\Phi\wedge a_0$, the $1$-form $a_0$, in turn,
determines the polarisation of the solution in equation
\eqref{eq:FGtrial}. Equation \eqref{eq:NeqA0xx} is thus a condition
that constrains possible polarisations once the direction of
propagation is known. Since equation \eqref{eq:NeqA0xx} is a linear
in $a_0$, we may study the dimension of the the solution space for
$a_0$.
To do this, let $\xi\in \Lambda^1_p(N)$ for some $p\in N$ and
for $\xi$ let $L_\xi$ be the linear map
$L_\xi \colon \Lambda^1_p(N)\to \Lambda^3_p(N)$,
\begin{eqnarray}
\label{Ldef}
L_\xi(\alpha) &=& \xi \wedge \kappa( \xi \wedge \alpha), \quad \alpha \in \Lambda_p^{1}(N).
\end{eqnarray}
We have $\xi \in \operatorname{ker} L_\xi$. For all $\xi \in \Lambda^1_p N\slaz$ we can then
find a (non-unique) vector subspace $V_{\xi}\subset \Lambda^1_pN$ such that
\begin{eqnarray}
\label{eq:directSumKer}
\kerOp L_{\xi} &=& V_{\xi} \, \oplus \, \operatorname{span} \xi.
\end{eqnarray}
Let $\xi = d\Phi\vert_p$ be nonzero. Then $V_{\xi}\slaz$ parameterises
possible $a_0$ that solve equation \eqref{eq:NeqA0xx} and for which
$A_0 = d\Phi\wedge a_0$ is nonzero. For a general $\kappa\in
\Omega^2_2(N)$ and $\xi\in \Lambda^1 (N)\slaz$ we can have $\dim
V_{\xi}\in\{0,1,2, 3\}$: Proposition \ref{prop:FresnelForRiemann} will
show that $\dim V_\xi$ can be $0$ or $2$, Example \ref{ex:dimVxi=1}
shows that $\dim V_\xi$ can be $1$, and the next proposition
characterise $\kappa\vert_p$ when $\dim V_\xi=3$ for all $\xi\in
\Lambda_p^1 (N)\slaz$.
\begin{proposition}
Let $\kappa\in \Omega^2_2(N)$ on a $4$-manifold $N$ and let $p\in N$.
Then the following are equivalent:
\begin{enumerate}
\item $\kappa\vert_p$ is of axion type.
\label{le:I}
\item $\dim V_\xi=3$ for all $\xi\in \Lambda_p^1(N)\slaz$.
\label{le:II}
\end{enumerate}
\end{proposition}
\begin{proof}
Implication \ref{le:I} $\Rightarrow$ \ref{le:II} is clear. For the
converse direction suppose that \ref{le:II} holds and $x^i$ are
local coordinates around $p$. It follows that
\begin{eqnarray*}
\zeta\wedge \xi\wedge \kappa(\xi\wedge \alpha)&=&0, \quad \alpha,\xi,\zeta\in \Lambda^1_p(N).
\end{eqnarray*}
If locally $\xi=\xi_i dx^i\vert_p$ then
$
\xi_i \xi_j \kappa^{ir}_{ab} \varepsilon^{jsab}=0.
$
Differentiating with respect to ${\xi_c}$ and ${\xi_d}$ gives
\begin{eqnarray*}
\kappa^{cr}_{ab} \varepsilon^{dsab} + \kappa^{dr}_{ab} \varepsilon^{csab} &=&0.
\end{eqnarray*}
With computer algebra it follows that
$\kappa =\frac 1 6 \operatorname{trace} \kappa\,\, \operatorname{Id}$
and \ref{le:I} follows.
\end{proof}
\subsection{Fresnel surface}
Let
$\kappa\in \Omega^2_2(N)$ on a $4$-manifold $N$.
If $\kappa$ is locally given by equation \eqref{eq:kappaLocal} in coordinates $x^i$, let
\begin{eqnarray*}
\cG^{ijkl}_0 &=& \frac 1 {48}
\kappa^{a_1 a_2}_{b_1 b_2}
\kappa^{a_3 i}_{b_3 b_4}
\kappa^{a_4 j}_{b_5 b_6}
\varepsilon^{b_1 b_2 b_5 k}
\varepsilon^{b_3 b_4 b_6 l}
\varepsilon_{a_1 a_2 a_3 a_4}.
\end{eqnarray*}
In overlapping coordinates $\{\widetilde x^i\}$, these coefficients
transform as
\begin{eqnarray}
\label{eq:TRtrans}
\widetilde \cG_0^{ijkl} &=& \det \left(\pd{x^r}{\widetilde x^s}\right)\, \cG_0^{abcd} \pd{\widetilde x^i}{x^a} \pd{\widetilde x^j}{x^b}\pd{\widetilde x^k}{x^c}\pd{\widetilde x^l}{x^d}.
\end{eqnarray}
Thus components $\cG^{ijkl}_0$ define a tensor density $\cG_0$ on $N$ of
weight $1$. The \emph{Tamm-Rubilar tensor density} \cite{Rubilar2002,
Obu:2003} is the symmetric part of $\cG_0$ and we denote this tensor density by $\cG$.
In coordinates,
$\cG^{ijkl} = \cG^{(ijkl)}_0$, where parenthesis indicate that
indices $ijkl$ are symmetrised with scaling $1/4!$. Using tensor
density $\cG$, the \emph{Fresnel surface} at a point $p\in N$
is defined as
\begin{eqnarray}
\label{eq:Fr}
F_p &=& \{\xi\in \Lambda^1_p(N) : \cG^{ijkl} \xi_i \xi_j \xi_k \xi_l = 0\}.
\end{eqnarray}
By equation \eqref{eq:TRtrans}, the definition of $F_p$ does not
depend on local coordinates. Let $F$ be the disjoint union of all
Fresnel surfaces, $F=\coprod_{p\in N} F_p$. To indicate that $F_p$ and
$F$ depend on $\kappa$ we also write $F_p(\kappa)$ and $F(\kappa)$.
If $\xi\in F_p$ then $\lambda \xi\in F_p$ for all $\lambda\in \setR$. In
particular $0\in F_p$ for each $p\in N$.
When $\cG\vert_p$ is non-zero, equation \eqref{eq:Fr} shows that $F_p$
is a fourth order surface in $\Lambda^1_p(N)$, so $F_p$ may contain
non-smooth self intersections.
\begin{theorem}
\label{eq:ThmDimTR}
Suppose $N$ is a $4$-manifold and $\kappa\in \Omega^2_2(N)$. If
$\xi\in \Lambda^1_p(N)$ is non-zero, then the following are
equivalent:
\begin{enumerate}
\item
\label{thm:TR_i}
$\dim V_\xi\ge 1$ where $V_\xi$ are defined as in equation \eqref{eq:directSumKer}.
\item
\label{thm:TR_ii}
$\xi$ belongs to the Fresnel surface $F_p\subset \Lambda^1_p(N)$.
\end{enumerate}
\end{theorem}
\begin{proof}
Let $\{x^i\}_{i=0}^3$ be coordinates around $p$ such that $dx^0\vert_p=\xi$ and
let $\ast$ be the Hodge star operator induced by the Euclidean Riemann metric $g_{ij} =\delta_{ij}$
in these coordinates. Let $P\colon \Lambda_p^1(N)\to
\Lambda_p^1(N)$ be the map $P=2\ast\circ L_\xi$. Then locally
\begin{eqnarray*}
P(\alpha) &=& \sum_{j=0}^3 \alpha_i \varepsilon^{0abj} \kappa^{0i}_{ab} dx^j,
\end{eqnarray*}
where $\alpha = \alpha_i dx^i\vert_p$ and $\kappa^{ij}_{ab}$ are defined as in equation
\eqref{eq:kappaLocal}. It follows that in the
basis $\{dx^i\vert_p\}_{i=0}^3$, the map $P$ is represented by the $4\times 4$ matrix $\operatorname{diag}(0, Q)$,
where $Q$ is the $3\times 3$ matrix $Q^{ij} = \varepsilon^{0abj} \kappa^{0i}_{ab}$, $i,j\in \{1,2,3\}$.
Now $\dim V_\xi\ge 1$ is equivalent with $\dim \operatorname{ker} P\ge 2$ which
is equivalent with $\det Q = 0$.
Writing out $\det Q=0$ using
\begin{eqnarray*}
\det Q &=& \frac 1 {3!} \varepsilon_{abc} \varepsilon_{ijk} Q^{ai} Q^{bj} Q^{ck}
\end{eqnarray*}
gives $\cG^{ijkl}\xi_i\xi_j\xi_k\xi_l=0$. We omit the proof of the
last step which can be found in \cite{Rubilar2002} and
\cite[p. 267 -- 268]{Obu:2003}.
\proofread{ Let us check this: First define
$\chi^{abcd} = \frac 1 2 \varepsilon^{abpq} \kappa_{pq}^{cd}$. Then
$Q^{ij} = 2 \chi^{0j0i}$ and $\chi^{abcd}$ is antisymmetric in $ab$
and $cd$. By anti-symmetry,
$$
\varepsilon_{0bcd} \chi^{0b0r} =
\frac 1 2
\varepsilon_{\alpha \beta cd} \chi^{\alpha \beta 0r}.
$$
Since $\varepsilon_{abc} = \varepsilon_{0abc}$ it follows that
\begin{eqnarray*}
\det Q
&=& \varepsilon_{0bcd} \varepsilon_{0rst} \chi^{0b0r}
\chi^{0c0s} \chi^{0d0t} \\
&=& \frac 1 2
\varepsilon_{\alpha \beta cd} \varepsilon_{r s0t} \chi^{\alpha \beta0r} \chi^{0c0s} \chi^{0d0t}\\
&=& \frac 1 4
\varepsilon_{\alpha \beta cd} \varepsilon_{r s\mu \tau } \chi^{\alpha \beta0r} \chi^{0c0s} \chi^{0d\mu \tau }\\
&=& \frac 1 4
\varepsilon_{\alpha \beta \gamma\delta} \varepsilon_{\rho \sigma\mu \tau } \chi^{\alpha \beta0\rho} \chi^{0\gamma 0\sigma} \chi^{0\delta \mu \tau }\\
\end{eqnarray*}
}
\end{proof}
Suppose $g$ is a pseudo-Riemann metric on an orientable $4$-manifold $N$ and
$\kappa \in \Omega^2_2(N)$. Then $g$ and $\kappa$ define a symmetric
$4\choose 0$-tensor on $N$ by
\begin{eqnarray}
\label{eq:Tg}
\cG_{g,\kappa} &=& \frac{1}{\sqrt{\vert \operatorname{det} g\vert}}\,\, \cG^{ijkl} \pd{}{x^i}\otimes \pd{}{x^j}\otimes \pd{}{x^k}\otimes \pd{}{x^l},
\end{eqnarray}
where $\cG^{ijkl}$ are local components of the Tamm-Rubilar tensor
density for $\kappa$, and $x^i$ are coordinates in an oriented
atlas for $N$.
A key property of symmetric $p\choose 0$-tensors is that they are
completely determined by their values on the diagonal
\cite{Mujica:2006, PunziEtAl:2009}. For symmetric $4\choose
0$-tensors on a $4$-manifold (like $\cG_{g,\kappa}$), the precise
statement is contained in the following polarisation
identity
\begin{proposition}
\label{prop:polarId}
Suppose $L$ is a symmetric $4\choose 0$-tensor on a $4$-manifold $N$. If
$x_1, x_2, x_3, x_4\in \Lambda^1_p(N)$ then
\begin{eqnarray*}
L(x_1, x_2, x_3, x_4) &=& \frac {1}{4! 2^4} \sum_{\theta_i\in \{\pm1\}}
\theta_1 \theta_2 \theta_3 \theta_4
\, L(\sum_{i=1}^4\theta_i x_i, \ldots, \sum_{i=1}^4\theta_i x_i).
\end{eqnarray*}
\end{proposition}
\subsection{Electromagnetic medium induced by a Hodge star operator}
In Proposition \ref{prop:HodgeHasOnlyPrincipalPart} we saw that a
pseudo-Riemann metric on a $4$-manifold induces a $2\choose 2$-tensor
$\kappa$ with only a principal part. The next example shows how
standard isotropic electromagnetic medium can be modelled using a
Lorentz metric on $\setR^4$.
\begin{example}
\label{eq:ABCDminkowski}
On $N=\setR\times \setR^3$ let $\kappa$ be the $2\choose 2$-tensor determined
$3\times 3$ matrices
$$
\cA = -\epsilon \operatorname{Id}, \quad
\cB = \mu^{-1}\operatorname{Id}, \quad
\cC =\cD = 0,
$$
where $\epsilon,\mu\colon \setR^3\to (0,\infty)$.
Then constitutive equations \eqref{eq:kappaLocal_Ix}--\eqref{eq:kappaLocal_IIx} are
equivalent with the isotropic constitutive equations
\begin{eqnarray}
\label{eq:isotropicEq1}
D &=& \epsilon \ast_0 E, \\
\label{eq:isotropicEq2}
B &=& \mu \ast_0 H,
\end{eqnarray}
where $\epsilon$ is the \emph{permittivity} and $\mu$ is the
\emph{permeability} of the medium and $\ast_0$ is the Hodge star operator
induced by the Euclidean metric on $\setR^3$.
If $\kappa$ is the $2\choose 2$-tensor defined as
$\kappa = \sqrt{\frac{\epsilon}{\mu}} \ast_g$ where $g$ is the Lorentz metric
$
g = \operatorname{diag}(-\frac{1}{\epsilon \mu},1,1,1)
$, then equations \eqref{eq:isotropicEq1}--\eqref{eq:isotropicEq2} are equivalent
with equation \eqref{FGchi}.
\proofBox
\end{example}
The next proposition shows that if $g$ is a pseudo-Riemann metric with
signature $(++++)$ or $(----)$ then the medium with $\kappa =
\ast_g$ has no asymptotic solutions. That is, if $d\Phi\vert_p$ is
non-zero, then equation \eqref{eq:NeqA0xx} implies that
$A_0\vert_p=0$. The proposition also shows that if $\kappa = \ast_g$
for an indefinite metric $g$, then $A_0$ can be non-zero only when
$d\Phi\vert_p$ is a \emph{null covector}, that is, when
$g(d\Phi\vert_p,d\Phi\vert_p)=0$.
Let $\operatorname{sgn}\colon \setR\to \{-1,+1\}$ be the \emph{sign function},
$\operatorname{sgn} x= -1$ for $x<0$,
$\operatorname{sgn} x= 0$ for $x=0$ and
$\operatorname{sgn} x= 1$ for $x>0$.
\begin{proposition}
\label{eq:ThgExp}
Let $g$ and $h$ be pseudo-Riemann metrics on $N$ on an orientable
$4$-manifold $N$. Then
\begin{eqnarray*}
\cG_{h,\ast g}(\xi,\xi,\xi,\xi) &=& \operatorname{sgn}(\det g)\, \sqrt{\frac{\vert \det g\vert }{\vert \det h\vert} }\left( g(\xi,\xi)\right)^2, \quad \xi\in \Lambda^1( N).
\end{eqnarray*}
Thus the Fresnel surface induced by the $2\choose 2$-tensor $\ast_g$ is given by
\begin{eqnarray*}
\label{eq:FresIsNullCone}
F(\ast_g) &=& \{ \xi\in \Lambda^1(N) : g(\xi,\xi)=0\}.
\end{eqnarray*}
\end{proposition}
\begin{proof}
Let $\cG^{ijkl}$ be components for the Tamm-Rubilar tensor density for $\ast_g$.
Computer algebra then gives
\begin{eqnarray*}
\cG^{abcd} \xi_a \xi_b \xi_c \xi_d &=& \operatorname{sgn}(\det g)\,
\sqrt{\vert \det g\vert} \left( g(\xi,\xi)\right)^2,
\end{eqnarray*}
where $\xi=\xi_a dx^a$ and the claim follows by equation \eqref{eq:Tg}.
\end{proof}
We know that a general plane wave in homogeneous isotropic medium in
$\setR^3$ can be written as a sum of two circularly polarised plane
waves with opposite handedness.
The \emph{Bohren decomposition} generalise this classical result to
electromagnetic fields in homogeneous isotropic chiral medium
\cite{LiSiTrVi:1994}. The \emph{Moses decomposition}, or
\emph{helicity decomposition}, further generalise this decomposition
to arbitrary vector fields on $\setR^3$. For a decomposition of
Maxwell's equations using this last decomposition, see
\cite{Moses:1971, Dahl2004}. In all of these cases,
an electromagnetic wave can be polarised in two
different ways. Part \ref{prop:FrLoI} in the next proposition shows
that this is also the case for asymptotic solutions as defined above
when the medium is given by the Hodge star operator of a indefinite metric.
\begin{proposition}
\label{prop:FresnelForRiemann}
Let $N$ be an orientable $4$-dimensional manifold, and let $\kappa \in \Omega^2_2(N)$
the $2\choose 2$-tensor $\kappa =\ast_g$ induced by a pseudo-Riemann metric $g$ on $N$.
\begin{enumerate}
\item
\label{prop:FrLoI}
If $\xi \in \Lambda^1(N)$ is non-zero, and $V_\xi$ is as in equation \eqref{eq:directSumKer}, then
\begin{eqnarray*}
\dim V_{\xi} &=& \begin{cases} 2, &\mbox{when} \, \, \xi\in F(\kappa), \\
0, &\mbox{when} \, \, \xi\notin F(\kappa).
\end{cases}
\end{eqnarray*}
\item
\label{prop:FrLoII}
If $\xi \in F(\kappa)$ is non-zero, and $L_\xi$ is as in equation \eqref{Ldef} then
\begin{eqnarray*}
\operatorname{ker}L_\xi &=& \xi^\perp,
\end{eqnarray*}
where $\xi^\perp = \{ \alpha\in \Lambda^1(N) : g(\alpha, \xi)=0\}$. Thus,
for any choice of $V_\xi$ in equation \eqref{eq:directSumKer} we have $V_\xi\subset \xi^\perp$.
\end{enumerate}
\end{proposition}
\begin{proof}
Let $p$ be the basepoint of $\xi$ and let
$\{x^i\}_{i=0}^3$ are local coordinates for $N$ around $p$ such that
$g=g_{ij} dx^i\otimes dx^j$ and $g_{ij}\vert_p$ is diagonal with
entries $\pm 1$. We know that $\kappa^2 = \ast_g^2 = (-1)^\sigma
\operatorname{Id}$, where $\sigma$ is the
\emph{index} of $g$ \cite[p. 412] {AbrahamMarsdenRatiu:1988}. If
$\alpha \in \Lambda^1_p(N)$, equations \eqref{Ldef} and
\eqref{eq:hodgeKappaLocal} imply that
\begin{eqnarray}
\nonumber
L_\xi(\alpha)&=& \frac 1 2 \xi_r \xi_s \alpha_i g^{r a} g^{i b} \varepsilon_{abcd} dx^s \wedge dx^{c}\wedge dx^{d} \\
\nonumber
&=& \det g \, (-1)^{\sigma} \alpha_i H^{ir} g_{rs} \ast dx^s,
\label{eq:transformToLocal}
\end{eqnarray}
where $\xi = \xi_i dx^i\vert_p$ and $\alpha = \alpha_i dx^i\vert_p$ and
\begin{eqnarray}
H^{ir} &=& g(\xi,\xi) g^{ir} - \xi_a g^{ai}\xi_b g^{br}.
\label{eq:Cik}
\end{eqnarray}
For part \ref{prop:FrLoI}, equations \eqref{eq:transformToLocal} and \eqref{eq:directSumKer}
imply that $\dim V_\xi = \dim \kerOp H-1$ where $H$ is the $4\times 4$
matrix with entries $H^{ij}$. Let $\sigma(H)$ denote the spectrum of
$H$ with eigenvalues repeated according to their algebraic
multiplicity. With computer algebra we find that
\begin{eqnarray*}
\sigma(H) &=& \left(0, C_1 g(\xi,\xi), C_2 g(\xi,\xi), C_3 \sum_{i=0}^3 \xi_i^2\right),
\end{eqnarray*}
where $C_i\in \{\pm 1\}$ are constants that depend only the signature
of $g$. Now part \ref{prop:FrLoI} follows by Proposition
\ref{eq:ThgExp} and since algebraic and geometric multiplicity of an
eigenvalue coincide for symmetric matrices \cite[p. 260]{Szabo:2002}.
For part \ref{prop:FrLoII}, equality $\operatorname{ker}
L_\xi=\xi^\perp$ follows from the local representation of $L_\xi$ in
equation \eqref{eq:Cik}.
\end{proof}
The next example shows that the case
$\operatorname{dim} V_\xi = 1$ is possible in equation \eqref{eq:directSumKer}.
The medium defined by equations \eqref{eq:biaxial}
is called
a \emph{biaxial crystal} \cite[Section 15.3.3]{BornWolf}.
\begin{example}
\label{ex:dimVxi=1}
On $M=\setR\times \setR^3$, let $\kappa\in \Omega^2_2(M)$ be defined by
\begin{eqnarray}
\label{eq:biaxial}
\cA = -\operatorname{diag}(1,2,3), \quad
\cB = \operatorname{Id}, \quad
\cC=\cD=0.
\end{eqnarray}
Then the Fresnel equation reads
\begin{eqnarray}
\label{eq:FresnelEqXXY}
6 \xi_0^4
- \xi_0^2 (5 \xi_1^2 + 8 \xi_2^2 + 9 \xi_3^2)
+ (\xi_1^2 + \xi_2^2 + \xi_3^2) (\xi_1^2 + 2 \xi_2^2 + 3 \xi_3^2) &=& 0.
\end{eqnarray}
Let $S$ be the solution set in $\setR^3$ to the above equation when
$\xi_0=1$. By equation \eqref{eq:FresnelEqXXY}, it is clear that $S$
is mirror symmetric about the $\xi_1\xi_2$, $\xi_1\xi_3$ and
$\xi_2\xi_3$ coordinate planes. Figure \ref{fig:singular} below
illustrates $S$ in the quadrant $\xi_1\ge 0, \xi_2\ge 0, \xi_3\ge 0$,
and in this quadrant we see that $S$ has one singular point
$\xi_{\operatorname{sing}}\in S$.
\begin{center}
\begin{figure}[!ht]
\includegraphics[width= 0.65\textwidth]{SingularSurface2.pdf}
\caption{One quadrant in $\setR^3$ of a Fresnel surface with a singular point illustrated by a dot.}
\label{fig:singular}
\end{figure}
\end{center}
Surface $S$ is defined implicitly by $f(\xi_1, \xi_2,\xi_3)=0$ and
singular points are characterised by $\nabla f=0$. This yields
$\xi_{\operatorname{sing}} = (
\sqrt{\frac{3}{2}},0,\frac{1}{\sqrt{2}})$. (For an alternative way to
solve this point, see \cite[Lemma 4.2 \emph{(iii)}]{Dahl2004}.) Using
computer algebra and the arguments used to prove Theorem
\ref{eq:ThmDimTR} we may compute $\dim V_\xi$ when $\xi_0=1$ and $S$
intersects one of the coordinate planes $\{\xi_i=0\}_{i=1}^3$. In
these intersections we obtain $\dim V_\xi=1$ except at the singular
point $\xi_{\operatorname{sing}}$ where $\dim V_\xi=2$. \proofBox
\end{example}
\section{Determining the medium from the Fresnel surface}
\label{sec:Closure}
As described in the introduction, the new proof of implication
\ref{coIII} $\Rightarrow$ \ref{coII} in the next theorem is the first
main result of this paper.
Regarding the other implications let us make a few remarks.
Implication \ref{coII} $\Rightarrow$ \ref{coI} is a standard result
for the Hodge star operator on a $4$-manifold. The converse implication
\ref{coI} $\Rightarrow$ \ref{coII} is less well known. The result was
first derived by Sch\"onberg \cite{Shoenberg:1971}. For further derivations and
discussions, see \cite{Jadczyk:1979, Rubilar2002, Obu:2003}.
Below we will give yet another proof using computer algebra. The proof
follows \cite{Obu:2003} and we use a Sch\"onberg-Urbantke-like formula
(see equations \eqref{eq:definitionOfGlobalG}--\eqref{eq:SU}) to
define a metric $g$ from $\kappa$. However, the below argument that
$g$ transforms as a tensor seems to be new. For a different argument,
see \cite[Section D.5.4]{Obu:2003}.
When a general $2\choose 2$-tensor $\kappa$ on a $4$-manifold satisfies
$\kappa^2 = -f \operatorname{Id}$ as in condition \ref{coI} one says that
$\kappa$ satisfies the \emph{closure condition}. For physical
motivation, see \cite[Section D.3.1]{Obu:2003}.
Let us emphasise that Theorem \ref{thm:mainResult} is a global result.
The result gives criteria for the existence of a Lorentz metric on a
$4$-manifold. In general, we know that a connected manifold $M$ has a
Lorentz metric if and only if $M$ is non-compact, or if $M$ is compact
and the Euler number $\chi(N)$ is zero \cite[Theorem
2.4]{MingSan:2008}. Let us also note that if $J$ is an \emph{almost
complex structure} on a manifold $M$, that is, $J$ is a $1\choose
1$-tensor on $M$ with $J^2=-\operatorname{Id}$ and $\dim M\ge 2$, then
$M$ is orientable \cite[p. 77]{Hsiung:1995}. It does not seem to be
known if the analogous result also holds for $2\choose 2$-tensors, that is,
if the closure condition on a $4$-manifold implies orientability.
\begin{theorem}
\label{thm:mainResult}
Suppose $N$ is an orientable $4$-manifold. If $\kappa\in
\Omega^2_2(N)$ satisfies $\kappaII=0$, then the following conditions
are equivalent:
\begin{enumerate}
\item $\kappa^2 = -f \operatorname{Id}$ for some function $f\in C^\infty(N)$ with $f>0$.
\label{coI}
\item
\label{coII}
There exists a Lorentz metric $g$ and a nonvanishing function $f\in
C^\infty(N)$ such that
\begin{eqnarray}
\kappa &=& f \ast_g.
\end{eqnarray}
\item
\label{coIII}
$\kappaIII=0$ and there exists a Lorentz metric $g$ such that
\begin{eqnarray*}
F(\kappa) &=& F(\ast_g),
\end{eqnarray*}
where $F(\kappa)$
is the Fresnel surface for $\kappa$ and $F(\ast_g)$
is the Fresnel surface for
the $2\choose 2$-tensor $\ast_g$.
\end{enumerate}
Moreover, when equivalence holds, then metrics $g$ in conditions \emph{\ref{coII}} and
\emph{\ref{coIII}} are conformally related.
\end{theorem}
\begin
{proof
For implication \ref{coI} $\Rightarrow$ \ref{coII} let $\eta = f^{-1/2}
\kappa$
whence $\eta^2=-\operatorname{Id}$, and
let $h$ be an auxiliary positive definite Riemann metric on $N$.
Let $\mathscr{T}$ be an atlas given by applying Lemma \ref{lemma:AinvLocally}
to $\eta$. For the local claim, let $(U,x^i)$ be a chart in $\mathscr{T}$,
and in this chart let $\eta$ be
represented by $3\times 3$ matrices $\cA$ and $\cK$.
With computer algebra we then obtain
\begin{eqnarray}
\label{eq:ThExp}
\cG_{h,\eta}(\xi,\xi,\xi,\xi) &=& \operatorname{sgn} \det \cA\ (G^{ab} \xi_a \xi_b)^2, \quad \xi\in \Lambda^1(U),
\end{eqnarray}
where $G=(G^{ab})$ is the $4\times 4$ matrix
\begin{eqnarray}
\label{eq:SU}
G &=& \frac 1 {(\det h)^{1/4}} \ \frac{1} {\vert \det \cA\vert^{1/2}}
\left( \begin{array}{c|c}
\det \cA & k^i \\
\hline
k^j & -\cA^{ij} + (\det \cA)^{-1} k^i k^j,
\end{array} \right),
\end{eqnarray}
and $k^i = \cA^{ib} \frac 1 2 \varepsilon_{bcd} \cK^{cd}$ for $i\in
\{1,2,3\}$.
Using a Shur complement \cite[Theorem 3.1.1]{Prasolov:1994} we find that
\begin{eqnarray*}
\det G &=&
\label{eq:Gdet}
-\frac{1}{\det h}.
\end{eqnarray*}
Hence $\det G<0$, so matrix $G$ is invertible and has constant signature $(-+++)$ or
$(+---)$ in $U$. Let $G_{ij}$ be the $ij$th entry of the inverse of
$G$. In $U$ we define
\begin{eqnarray}
\label{eq:definitionOfGlobalG}
g &=& \sigma_U G_{ij} dx^i \otimes dx^j,
\end{eqnarray}
where constant $\sigma_U\in \{-1,1\}$ is chosen such that $g$ has signature
$(-+++)$. Then $g$ defines a smooth symmetric $0\choose 2$-tensor
in $U$ with signature $(-+++)$, and by computer algebra we have
\begin{eqnarray}
\label{eq:localClaimEta}
\eta\vert_U &=& -\operatorname{sgn} \operatorname{det} \cA \, \ast_g.
\end{eqnarray}
This completes the local claim in \ref{coI} $\Rightarrow$ \ref{coII}.
For the global claim, let $(U, x^i)$ and $(\widetilde U,\widetilde
x^i)$ be overlapping charts in $\mathscr{T}$, and in these charts let
$G^{ij}$ and $\widetilde G^{ij}$ be defined as above. Since
$\cG_{h,\eta}$ is a tensor, equation \eqref{eq:ThExp} implies that
\begin{eqnarray}
\label{eq:coordChange}
\operatorname{sgn} \det \cA\ (G^{ij} \xi_i \xi_j)^2
&=&
\operatorname{sgn} \det \widetilde \cA\ (\widetilde G^{ij} \pd{x^r}{\widetilde x^i}\pd{x^s}{\widetilde x^j}
\xi_r \xi_s)^2
\end{eqnarray}
for all $\xi=\xi_i dx^i\in \Lambda^1(U\cap \widetilde U)$. Since $G^{ab}$ is
non-degenerate we can find a $\xi$ such that the left hand side is
non-zero. Thus $\operatorname{sgn} \det \cA=\operatorname{sgn} \det
\widetilde \cA$ in $U\cap \widetilde U$ and $\operatorname{sgn} \det
\cA$ in equation \eqref{eq:localClaimEta} defines a smooth function
$N\to \setR$.
By Theorem \ref{prop:Principle} \ref{eq:xx3} $\Rightarrow$ \ref{eq:xx1}
there exists a smooth nonvanishing function $\lambda\colon U\cap \widetilde U\to \setR$ such that
\begin{eqnarray*}
G^{ij} &=& \lambda \widetilde G^{rs} \pd{x^i}{\widetilde x^r}\pd{x^j}{\widetilde x^s}.
\end{eqnarray*}
Equation \eqref{eq:coordChange} implies that function $\lambda$ can only take values $\{-1,+1\}$.
Thus
\begin{eqnarray*}
G_{ij} &=& \lambda \widetilde G_{rs} \pd{\widetilde x^r}{ x^i}\pd{\widetilde x^s}{ x^j}.
\end{eqnarray*}
Since $\sigma_U G_{ij}$ and $\sigma_{\widetilde U} \widetilde G_{ij}$ both have signature $(-+++)$.
It follows that $\lambda \sigma_U = \sigma_{\widetilde U}$ in $U\cap \widetilde U$, and equation
\eqref{eq:definitionOfGlobalG} defines a tensor on $N$.
This completes the proof of implication \ref{coI} $\Rightarrow$ \ref{coII}.
Implication \ref{coII} $\Rightarrow$ \ref{coIII}
follows by Propositions \ref{prop:HodgeHasOnlyPrincipalPart} and \ref{eq:ThgExp}.
For the proof of implication \ref{coIII} $\Rightarrow$ \ref{coI} we first establish two subclaims:
\textbf{Claim $1$.} The $4\choose 0$-tensor $\cG_{g,\ast g}$ is pointwise proportional to $\cG_{g,\kappa}$
by a non-zero constant.
Let $p\in N$. By Proposition \ref{prop:polarId} we only need to show that there exists a $\lambda\in \setR$ such that
\begin{eqnarray*}
\cG_{g,\ast g}(\xi,\xi,\xi,\xi)
&=& \lambda \cG_{g,\kappa}(\xi,\xi,\xi,\xi), \quad \xi\in \Lambda^1_p(N).
\end{eqnarray*}
Let $x^i$ be coordinates around $p$ such that $g\vert_p=k
\operatorname{diag}(1,-1,-1,-1)$ for $k\in \{\pm 1\}$. In these
coordinates, let $\cG^{ijkl}_{g,\ast g}$ and $\cG^{ijkl}_{g,\kappa}$
be components for the symmetric $4\choose 0$-tensors $\cG_{g,\ast
g}\vert_p$ and $\cG_{g,\kappa}\vert_p$, so that
$$
\cG_{g,\ast g}(\xi,\xi,\xi,\xi) =\cG^{ijkl}_{g,\ast_g} \xi_i \xi_j \xi_k \xi_l, \quad
\cG_{g,\kappa}(\xi,\xi,\xi,\xi) =\cG^{ijkl}_{g,\kappa} \xi_i \xi_j \xi_k \xi_l
$$
for $\xi = \xi_i dx^i\vert_p$. Using these components, let
$P,Q$ be the polynomials $P,Q\colon \setR^4\to \setR$,
$$
P(\xi_0,\vq) = \cG^{ijkl}_{g,\ast_g} \xi_i \xi_j \xi_k \xi_l, \quad
Q(\xi_0, \vq) =\cG^{ijkl}_{g,\kappa} \xi_i \xi_j \xi_k \xi_l,
$$
where $\xi_0\in \setR$, $\vq=(\xi_1,\xi_2,\xi_3) \in \setR^3$.
By
Proposition \ref{eq:ThgExp},
\begin{eqnarray*}
P(\xi_0, \vq) &=& -(\xi_0^2 - \vert\vq\vert^2)^2\\
&=& -\left(\xi_0-\vert\vq\vert \right)^2 \left(\xi_0+\vert\vq\vert \right)^2,
\end{eqnarray*}
for all $(\xi_0, \vq)\in \setR^4$ when $\vert \vq\vert$ is the
Euclidean norm of $\vq$. Thus $P(1,0,0,0)\neq 0$ so $dx^0\vert_p\notin
F_p(\ast_g)=F_p(\kappa)$ whence $\cG_{g,\kappa}^{0000}\neq 0$. For each
$\vq\in \setR^3$, $Q(\xi_0,\vq)$ is then a fourth order polynomial in
$\xi_0$ with coefficients determined by $\vq\in \setR^3$. Hence there
exists continuous maps
$$
r_i\colon \setR^3\to \setC, \quad i\in \{1,2,3,4\}
$$
so that for all $\vq\in \setR^3$, $\{r_i(\vq)\}_{i=1}^4$ are the roots
of $Q(\cdot, \vq)$ \cite{NaulinPabst:1994}. For each $\vq\in \setR^3$
there exists a $\alpha(\vq)\in \setR$ such that
\begin{eqnarray}
\label{eq:QxiPol}
Q(\xi_0, \vq) &=& \alpha(\vq) \prod_{i=1}^4 (\xi_0 - r_i(\vq)), \quad \xi_0\in \setR.
\end{eqnarray}
Applying $\partial^4/\partial\xi_0^4$ to both sides implies that
$\alpha(\vq)=\cG^{0000}_{g,\kappa}$. In particular, the map
$\vq\mapsto \alpha(\vq)$ is constant and non-zero. Let $\mu =
\cG^{0000}_{g,\kappa}$.
Since $P$ and $Q$ have the same zero set, there exists functions
$s_i\colon \setR^3\to \{-1,1\}$ such that
\begin{eqnarray*}
r_i(\vq) &=& s_i(\vq) \vert \vq\vert, \quad \vq \in \setR^3, \, i\in \{1,2,3,4\}.
\end{eqnarray*}
We know that $\setR^3\slaz$ is path connected. Hence $\setR^3\slaz$ is
connected. For a contradiction, suppose that
$s_i(\setR^3\slaz)=\{-1,+1\}$ for some $i\in \{1,2,3,4\}$. Then
$\setR^3\slaz = U_+\cup U_-$ for open, non-empty and disjoint sets
$U_\pm$ defined as
\begin{eqnarray*}
U_\pm &=& \{\vq\in \setR^3\slaz: \pm r_i(\vq) > 0\}.
\end{eqnarray*}
It follows that there are constants $s_1,s_2, s_3,s_4\in \{-1,+1\}$ such that
\begin{eqnarray}
\label{eq:ridef}
r_i(\vq) &=& s_i \vert \vq\vert, \quad \vq \in \setR^3, \, i\in \{1,2,3,4\}.
\end{eqnarray}
Let $\sigma$ be the number of $s_i$ with $s_i=1$. If $\vq\in
\setR^3\slaz$, then polynomial $P(\cdot, \vq)$ has two distinct roots
$\pm \vert \vq\vert$. Hence $\sigma=0$ or $\sigma = 4$ are not
possible, so $\sigma\in \{1,2,3\}$ and by equation \eqref{eq:QxiPol},
\begin{eqnarray*}
\label{eq:Qexp}
Q(\xi_0, \vq) &=& \mu \left(\xi_0-\vert\vq\vert \right)^{\sigma} \left(\xi_0+\vert\vq\vert \right)^{4-\sigma}\\
&=& \begin{cases}
\mu \left( \xi_0^4 - \vert \vq\vert^4 + 2\xi_0 \vert \vq\vert (\xi_0^2 -\vert\vq\vert^2)\right), &\mbox{if}\, \sigma=1,\\
\mu\, (\xi_0^2 - \vert \vq\vert^2)^2, &\mbox{if}\, \sigma=2,\\
\mu \left( \xi_0^4 - \vert \vq\vert^4 - 2\xi_0 \vert \vq\vert (\xi_0^2 -\vert\vq\vert^2)\right) , &\mbox{if}\, \sigma=3,
\end{cases}
\end{eqnarray*}
for all $(\xi_0, \vq)\in \setR^4$. Since $Q$ is a polynomial, we know
that $t\mapsto Q(1,t,0,0)$ is smooth near $0$. This is only possible when $\sigma=2$, and Claim $1$
follows.
\textbf{Claim $2$.} At each $p\in N$ there exists a non-zero
$\lambda\in \setR$ such that $\kappa\vert_p = \lambda \ast_g\vert_p$.
Let $p\in N$. By Proposition \ref{prop:localReduction} \ref{cc:CC}
there are coordinates $x^i$ around $p$ such that $g\vert_p=k
\operatorname{diag}(1,-1,-1,-1)$ for some $k\in \{\pm 1\}$ and
$\cA\vert_p$ is diagonal. For $\xi = \xi_i dx^i\vert_p$ we then have
$$
\cG_{g,\ast_g}(\xi,\xi,\xi,\xi) = \cG^{ijkl}_{g,\ast_g} \xi_i \xi_j \xi_k \xi_l, \quad
\cG_{g,\kappa}(\xi,\xi,\xi,\xi) = \cG^{ijkl}_{g,\kappa} \xi_i \xi_j \xi_k \xi_l,
$$
where $\cG^{ijkl}_{g,\ast g}$ and $\cG^{ijkl}_{g,\kappa}$ are components for
$\cG_{g,\ast_g}$ and $\cG_{g,\kappa}$ in coordinates $x^i$.
By Claim $1$ there exists a $\lambda\in \setR\slaz$ such that
\begin{eqnarray*}
\cG^{ijkl}_{g,\ast g} &=& \lambda \cG^{ijkl}_{g,\kappa}, \quad 0\le i\le j\le k\le l\le 3.
\end{eqnarray*}
Moreover, $\kappaIII=0$. We then have $36$ polynomial equations for
$\kappa$. Using the Gr\"obner basis (see Appendix \ref{app:Groebner})
for these equations we find that the equations have a unique real
solution for $\kappa$ and this solution is given by $\kappa\vert_p =
\lambda^{-1/3} \ast_g\vert_p$. This completes the proof of Claim $2$.
By Claim $2$, there exists a map $\lambda \colon N\to \setR\slaz$ such that
$\kappa = \lambda \ast_g$ whence $\kappa^2 = -\lambda^2 \operatorname{Id}$. To
see that $\lambda^2$ is smooth it suffices to note that $\lambda^2 = -\frac 1 6
\operatorname{trace} \kappa^2$. This completes the proof of
implication \ref{coIII} $\Rightarrow$ \ref{coI}.
When equivalence holds,
Proposition \ref{eq:ThgExp} and Theorem \ref{prop:Principle} imply that the
Lorentz metrics in conditions \ref{coII} and \ref{coIII} are conformally related.
\end{proof}
The next lemma was used to prove implication \ref{coI} $\Rightarrow$
\ref{coII} in Theorem \ref{thm:mainResult}. In the proof of the lemma, Claim 1
is based on \cite[Sections D.4--D.5]{Obu:2003}.
\begin{lemma}
\label{lemma:AinvLocally}
Suppose $N$ is an orientable $4$-manifold and $\kappa\in
\Omega^2_2(N)$. If $\kappa$ has no skewon component and $\kappa^2 =
-\operatorname{Id}$, then $N$ has an oriented atlas $\mathscr{T}$ with
the following property:
Each $p\in N$ can be covered with a connected chart $(U,x^i)\in
\mathscr{T}$ such that if $\cA, \cB, \cC, \cD$ represent $\kappa$ in
$U$, then
\begin{enumerate}
\item $\cA$ is invertible in $U$.
\label{lemma:collectB}
\item In $U$ there exists a smoothly varying antisymmetric $3\times 3$ matrix $\cK$ such that
$$
\quad\quad
\cB= -\cA^{-1} \left( \operatorname{Id}+ \left(\cK \cA^{-1}\right)^2\right),\quad
\cC = \cK\cA^{-1},\quad
\cD = -\cA^{-1} \cK.
$$
\label{lemma:collectD}
\end{enumerate}
\end{lemma}
\begin{proof}
Let us first make an observations: Suppose $\{x^i\}_{i=0}^3$ are
arbitrary coordinates for $N$ and $\cA$, $\cB$, $\cC, \cD$ are
$3\times 3$ matrices that represent $\kappa$ in these
coordinates. Then Proposition \ref{prop:localReduction} \ref{cc:AA}
implies that $\kappa^2=-\operatorname{Id}$ is equivalent with
\begin{eqnarray}
\cC^2 + \cA\cB
\label{eq:abc1}
&=& -\operatorname{Id}, \\
\cB \cC + \cC^T \cB
\label{eq:abc2}&=& 0, \\
\cC \cA + \cA\cC^T
\label{eq:abc3}
&=& 0.
\end{eqnarray}
Let $\mathscr{T}_0$ is a maximal oriented atlas for $N$.
The proof is divided into two subclaims, Claim $1$ and Claim $2$.
\textbf{Claim 1.} For each $p\in N$ there exists a connected chart
$(U, x^i)$ that satisfy condition \ref{lemma:collectB} and there
exists a chart $(W, y^i)\in \mathscr{T}_0$ with $U\cap W\neq
\emptyset$ such that the transition map $x^i\mapsto y^i$ is
orientation preserving.
By Proposition \ref{prop:localReduction} \ref{cc:BB} we can find a
connected chart $(U, x^i)$ that contains $p$ and where matrix $\cA$ for
$\kappa$ is diagonal at $p$. The rest of Claim $1$ is divided into
four cases depending on the eigenvalues of $\cA\vert_p$.
\textbf{Case A.} Suppose all three eigenvalues of $\cA\vert_p$ are
non-zero. Since eigenvalues depend continuously on the matrix entries
\cite{NaulinPabst:1994}, we can shrink $U$ and
part \ref{lemma:collectB} follows. Claim $1$ follows by possibly
reflecting the $x^1$-coordinate.
\textbf{Case B.} Suppose $\cA\vert_p$ has two non-zero eigenvalues. By
permutating the coordinates (see equation \eqref{eq:transRuleA}) we
may assume that $\cA\vert_p=\operatorname{diag}(a_1, a_2,0)$ for some
$a_1, a_2\neq 0$. Writing out equations
\eqref{eq:abc1}--\eqref{eq:abc3} with computer algebra gives
$$
\cC^1\,_1= \cC^2\,_2= \cC^3\,_1=\cC^3\,_2=0,\quad (\cC^3\,_3)^2=-1
$$
at $p$. The last equation contradicts that $\cC$ is real. Case B is
therefore not possible.
\textbf{Case C.} Suppose $\cA\vert_p$ has one non-zero eigenvalue. As
in Case B, we can find a chart $(U,x^i)$ for which $\cA\vert_p =
\operatorname{diag}(a_1,0,0)$ for some $a_1\neq 0$. Writing out
equations \eqref{eq:abc1}--\eqref{eq:abc3} as in Case B gives
$$
\cC^1\,_1= \cC^2\,_1= \cC^3\,_1=0,\quad \cB_{11}\neq 0, \quad \cC^2\,\,_3\neq 0,\quad \cC^3\,_2\neq 0
$$
at $p$. Let $\{\widetilde x^i\}_{i=0}^3$ be coordinates around $p$ defined as
\begin{eqnarray*}
\widetilde x^0 &=& x^0 + x^3,\\
\widetilde x^i &=& x^i, \quad i\in \{1,2,3\}.
\end{eqnarray*}
In these coordinates, matrix $\widetilde \cA\vert_p$
has determinant $-\cB_{11} (\cC^3\,_2)^2$, which is
non-zero, and Claim $1$ follows as in Case $A$.
\textbf{Case D.} Suppose all eigenvalues of $\cA\vert_p$ are zero. Then
$\cA\vert_p=0$ and equation \eqref{eq:abc1} implies that
$(\det\cC\vert_p)^2=-1$. This contradicts that $\cC\vert_p$ is a real matrix.
Case D is therefore not possible.
\textbf{Claim 2.} Let $\mathscr{T}$ be the collection of all charts
$(U, x^i)$ as in Step $1$ when $p$ ranges over all points in $N$. Then
$\mathscr{T}$ satisfies the sought properties.
Let $(U, x^i)$ and $(\widetilde U, \widetilde x^i)$ be overlapping
charts in $\mathscr{T}$. By Claim 1 and \cite[Lemma
13.9]{Trofimov:1994}, each chart $U$ and $\widetilde U$ is compatible
with all charts in $\mathscr{T_0}_0$. Hence the transition map $x^i
\mapsto \widetilde x^i$ is orientation preserving, and $\mathscr{T}$
is oriented.
We know that each chart in $\mathscr{T}$ satisfies property
\ref{lemma:collectB}, and property \ref{lemma:collectD} follows by
defining $\cK = \cC \cA$. Indeed, $\cK$ is antisymmetric by equation
\eqref{eq:abc3}, and the expression for $\cB$ follows by equation
\eqref{eq:abc1}.
\end{proof}
\section{Non-injectivity results}
\label{sec:uni}
Implication \ref{coIII} $\Rightarrow$ \ref{coII} in Theorem
\ref{thm:mainResult} shows that for a special class of medium, the
Fresnel surface determines the medium up to a conformal factor. In
this section we will describe results and examples where the opposite
is true. In the below we will see that there are various
non-uniquenesses that prevents us from determining $\kappa$ (or even
the conformal class of $\kappa$) from only the Fresnel surface
$F(\kappa)$.
Let us study the non-injectivity of the two maps in the diagram below:
\begin{eqnarray}
\label{dia:diag2}
\kappa \quad \to \quad \cG(\kappa) \quad \to \quad F(\kappa),\quad \quad\kappa\in \Omega^2_2(N),
\end{eqnarray}
where $\cG(\kappa)$ is the Tamm-Rubilar $4\choose 0$-tensor density
induced by $\kappa$.
\subsection{Non-injectivity of leftmost map}
\label{sec:TRinjective1}
Let us first study the non-injectivity of the leftmost map in diagram
\eqref{dia:diag2}, that is, the map
\begin{eqnarray}
\label{eq:LdefRmap}
\kappa\quad \to\quad \cG(\kappa), \quad\quad \kappa\in \Omega^2_2(N).
\end{eqnarray}
Parts \ref{thm:FkappaInvB:ii}--\ref{thm:FkappaInvB:iv} in the next
theorem describe three invariances that make the map in \eqref{eq:LdefRmap}
non-injective. The first two parts are well known \cite[Section
2.2]{Obu:2003}. However, let us make three remarks regarding part
\ref{thm:FkappaInvB:iv}.
First, an interpretation of part \ref{thm:FkappaInvB:iv} is as follows:
If $F,G$ solve the sourceless Maxwell equations in medium $\kappa$,
then $G,F$ solve the sourceless Maxwell equations in medium $\kappa^{-1}$.
In this setting, part \ref{thm:FkappaInvB:iv} states that both media have the
same Fresnel surfaces.
Second, suppose $\ast_g$ is the $2\choose 2$-tensor induced by a
pseudo-Riemann metric $g$. Then $\ast_g^2= \pm
\operatorname{Id}$, so $\ast_g^{-1}= \pm \ast_g$, whence
$\cG(\ast_g)$ and $\cG(\ast_g^{-1})$ are always conformally
related. Part \ref{thm:FkappaInvB:iv} states that this is
not only a result for Hodge star operators, but a general result
for \emph{all} $2\choose 2$-tensors.
Third, the proof of part \ref{thm:FkappaInvB:iv} is based on computer
algebra. Of all the proofs in this paper, this computation is
algebraicly most involved. For example, if we write out equation
\eqref{eq:bigEquation} as a text string, it requires almost $13$
megabytes of memory.
\begin{theorem}
\label{thm:FkappaInvB}
Suppose $\kappa\in \Omega^2_2(N)$ where $N$ is a $4$-manifold. Then
\begin{enumerate}
\item
\label{thm:FkappaInvB:i}
$\cG (f\kappa)=f^3 \cG(\kappa)$ for all $f\in C^\infty(N)$,
\item
\label{thm:FkappaInvB:ii}
$\cG(\kappaII)=0$,
\item
\label{thm:FkappaInvB:iii}
$
\cG(\kappa) = \cG(\kappa + f \operatorname{Id})
$ for all $f\in C^\infty(N)$,
\item
\label{thm:FkappaInvB:iv}
$\cG(\kappa^{-1})=\cG(-(\det\kappa)^{-1/3} \kappa)$ when
$\kappa$ is invertible.
\end{enumerate}
\end{theorem}
\begin{proof}
Part \ref{thm:FkappaInvB:i} follows by the definition, and parts
\ref{thm:FkappaInvB:ii}--\ref{thm:FkappaInvB:iii} are proven in \cite[Section 2.2]{Obu:2003}.
Therefore we only need to prove part \ref{thm:FkappaInvB:iv}.
Let $\operatorname{adj} \kappa = \det \kappa \, \kappa^{-1}$ be the \emph{adjugate} of $\kappa$.
By part \ref{thm:FkappaInvB:i} it suffices to show that
\begin{eqnarray}
\label{eq:bigEquation}
( \det \kappa)^2 \cG^{ijkl}_\kappa + \cG^{ijkl}_{\operatorname{adj} \kappa} &=& 0,
\quad 0\le i\le j\le k \le l \le 3,
\end{eqnarray}
where $\cG^{ijkl}_\kappa$ and $\cG^{ijkl}_{\operatorname{adj} \kappa}$
are components of the Tamm-Rubilar tensor densities of $\kappa$ and
$\operatorname{adj} \kappa$, respectively. The motivation for rewriting
the claim as in equation \eqref{eq:bigEquation} is that now both terms are
polynomials. By using the method described in Appendix
\ref{app:veryLarge} we obtain that equations \eqref{eq:bigEquation}
hold, and part \ref{thm:FkappaInvB:iv} follows.
\end{proof}
Theorem \ref{thm:FkappaInvB} \ref{thm:FkappaInvB:ii}
shows that if we restrict the map in equation \eqref{eq:LdefRmap} to purely
skewon tensors, we do not obtain an injection. The next example shows
that the same map is neither an injection when restricted to tensors of
purely principal type.
\begin{example}
On $N=\setR\times \setR^3$ with coordinates $\{x^i\}_{i=0}^3$, let $\kappa$
be the $2\choose 2$-tensor defined by $3\times 3$-matrices
\begin{eqnarray*}
\cA= 0_{3\times 3},\,\,\,
\cB= \begin{pmatrix}
0 & 0 & \lambda_1 \\
0 & 0 & \lambda_2 \\
\lambda_1 & \lambda_2 & \lambda_3
\end{pmatrix},\,\,\,
\cC= \begin{pmatrix}
-2^{-1/3} & 0 & \lambda_4 \\
0 & -2^{-1/3} & \lambda_5 \\
0 & 0& 2^{2/3}
\end{pmatrix},\,\,\,
\cD = \cC^T,
\end{eqnarray*}
where parameters $\lambda_1,\ldots, \lambda_5\in \setR$ are arbitrary.
Then $\kappa$ has only a principal part, $\det \kappa = 1$, and
\begin{eqnarray*}
\cG_{h,\kappa}(\xi,\xi,\xi,\xi) &=& 0, \quad \xi\in \Lambda^1(N)
\end{eqnarray*}
for any pseudo-Riemann metric $h$ on $N$.
\proofBox
\end{example}
When proving implication \ref{coIII} $\Rightarrow$ \ref{coI} in
Theorem \ref{thm:mainResult} we need to assume that $\kappa$ has real
coefficients. In fact, for $2\choose 2$-tensors with complex
coefficients a decomposition into principal-, skewon-, and axion
components does not seem to have been developed. The next example
shows that there are non-trivial complex tensors whose Fresnel surface
everywhere coincides with the Fresnel surface for the standard
Minkowski metric $g_0 = \operatorname{diag}(1,-1,-1,-1)$. (For
$\kappa\in \Omega^2_2(N,\setC)$ we define the Fresnel surface using
the same formulas as for real $\kappa$.)
\begin{example}
\label{ex:complexExample}
On $N=\setR \times \setR^3$ with coordinates $\{x^i\}_{i=0}^3$, let $\kappa$
be the $2\choose 2$-tensor with complex coefficients defined by $3\times 3$-matrices
\begin{eqnarray*}
\cA&=&
-\begin{pmatrix}
\frac{1}{2 z^2} & 0 & 0 \\
0 & 2z & 0 \\
0 & 0& z
\end{pmatrix},\,\,\,
\quad\quad\quad\quad\quad\quad\quad\,\,
\cB= -\cA, \quad\\
\cC&=&
i \begin{pmatrix}
\frac 1 {3z^2} -z & 0 & 0 \\
0 & -\frac {1}{6z^2} + z & 0 \\
0 & 0& -\frac{1}{6z^2}
\end{pmatrix},\,\,\,\quad\quad
\cD = \cC,
\end{eqnarray*}
where $z$ is an arbitrary function $z\colon N \to \setC\slaz$ and $i$ is the complex unit.
At each $p\in N$ the Fresnel surface is then determined by
\begin{eqnarray*}
\xi_0^2 - \xi_1^2-\xi_2^2-\xi_3^2&=&0,
\end{eqnarray*}
where $\xi_i dx^i \in \Lambda^1_p(N)$, and
\begin{eqnarray*}
\operatorname{trace} \kappa = 0, \quad
\det \kappa = \frac{\left(1+6z^3\right)^3 \, \left( 5 - 126 z^3 +684 z^6 - 648 z^9 \right) }{ 46 656\,\, z^{12}}.
\end{eqnarray*}
From the latter equation we see that for specific values of $z$, tensor $\kappa$ can be non-invertible as a linear map
\proofBox
\end{example}
\subsection{Non-injectivity of rightmost map}
\label{sec:TRinjective2}
The next example shows that there are skewon-free $2\choose 2$-tensors
$\kappa_1$ and $\kappa_2$ that have the same Fresnel surfaces, but
their Tamm-Rubilar tensors are not proportional to each other. This
shows that the rightmost map in equation \eqref{dia:diag2} is not
injective. Let us point out that this contradicts the first
proposition in \cite{PunziEtAl:2009} whose proof does not analyse
multiplicities of roots to the Fresnel equation.
\begin{example}
On $N=\setR\times \setR^3$ with coordinates $\{x^i\}_{i=0}^3$, let $\kappa_1$
be the $2\choose 2$-tensor defined by $3\times 3$-matrices
\begin{eqnarray*}
\cA_1= \begin{pmatrix}
0 & -1 & 1 \\
-1 & -2 & 1 \\
1 & 1 & -1
\end{pmatrix},\,\,\,
\cB_1= \begin{pmatrix}
0 & \frac 1 2 & 0 \\
\frac 1 2 & 0 & 0 \\
0 & 0 & 0
\end{pmatrix},\,\,\,
\cC_1= \begin{pmatrix}
0 & 0 & 0 \\
0 & 2 & 1 \\
\frac 1 2 & -\frac 1 2 & 1
\end{pmatrix},\,\,\,
\cD_1 = \cC_1^T.
\end{eqnarray*}
For the Euclidean metric $g_0$ on $N$ we then have
\begin{eqnarray*}
\cG_{g_0,\kappa_1}(\xi,\xi,\xi,\xi) &=& (\xi_0-\xi_1)(\xi_0-\xi_2)^3, \quad \xi\in \Lambda^1(N).
\end{eqnarray*}
To exchange the role of $\xi_1$ and $\xi_2$, we perform a coordinate
change $x_0\mapsto x_0$, $x_1\mapsto x_2$, $x_2\mapsto x_1$,
$x_3\mapsto x_3$. With this as motivation we define $\kappa_2$ as the
$2\choose 2$-tensor defined by $3\times 3$-matrices
\begin{eqnarray*}
\cA_2= \begin{pmatrix}
2 & 1 &-1 \\
1 & 0 & -1 \\
-1 & -1 & 1
\end{pmatrix},\,\,\,
\cB_2= \begin{pmatrix}
0 & -\frac 1 2 & 0 \\
-\frac 1 2 & 0 & 0 \\
0 & 0 & 0
\end{pmatrix},\,\,\,
\cC_2= \begin{pmatrix}
2 & 0 & 1 \\
0 & 0 & 0 \\
-\frac 1 2 & \frac 1 2 & 1
\end{pmatrix},\,\,\,
\cD_2 = \cC_2^T.
\end{eqnarray*}
Then
\begin{eqnarray*}
\cG_{g_0,\kappa_2}(\xi,\xi,\xi,\xi) &=& -(\xi_0-\xi_1)^3(\xi_0-\xi_2), \quad \xi\in \Lambda^1(N).
\end{eqnarray*}
Here $\kappa_1$ and $\kappa_2$ are not proportional,
their Tamm-Rubilar tensor densities are not proportional, but their Fresnel surfaces coincide.
Both $\kappa_1$ and $\kappa_2$ have $1$ has an eigenvalue of algebraic multiplicity $6$.
Hence
$$
\det \kappa_1 = \det \kappa_2 = 1, \quad \operatorname{trace}\kappa_1 = \operatorname{trace}\kappa_2 = 6,
$$
and for the trace-free components $\widetilde \kappa_i = \kappa_i-\operatorname{Id}$ we have $\det \widetilde \kappa_i=0$.
\proofBox
\end{example}
\textbf{Acknowledgements.} I would like to thank Luzi Bergamin and
Alberto Favaro for helpful discussions by email regarding
\cite{FavaroBergamin:2011}.
The author gratefully appreciates financial
support by the Academy of Finland (project 13132527 and Centre of
Excellence in Inverse Problems Research), and by the Institute of
Mathematics at Aalto University.
|
\section{Introduction}
The first run of the LHC aims to collect few fb$^{-1}$ of data in the $pp$ collisions at a center of mass energy $\sqrt{s}=7$ TeV. This is the first possibility to explore the Fermi scale beyond the reach of the Tevatron ($\sqrt{s}=1.96$ TeV). It is therefore very important to understand what we can expect during and at the end of this run. It was recently pointed out that with about $5$ fb$^{-1}$ of integrated luminosity, the combination of the data collected from both the ATLAS and CMS experiments can suffice to exclude, or even to discover, the Standard Model (SM) Higgs boson down to the LEP limit of $114$ GeV \cite{CMS:2010Hig,ATLAS:2011Hig}. Even if the discovery of the SM Higgs boson is certainly the most important task of the early LHC, we think that an open attitude in the expectation for possible signals of new physics is necessary at this stage. Model building prejudices normally play an important role in the determination of the experimental strategies. However, we think that these prejudices should be avoided as much as possible in order to develop model independent search strategies and to maximize the sensitivity to a large range of new physics models. For these reasons, we want to discuss the possible signals of new physics that we can expect at the early LHC in a model independent way. We base our considerations on simple phenomenological Lagrangians fulfilling reasonable consistency conditions without aiming to explain the underlying theory from which the states under discussion can arise.\\
\indent The single production of a relatively narrow resonance is the most obvious candidate for a large rate of new physics events. The case of the $q\bar{q}$ vector resonances, either neutral or charged, has been widely studied (see, e.g., Refs.~\cite{Langacker:2008yv, Salvioni:2010p010,Grojean:2011vu}). However, for parton luminosity reasons, we find that at the LHC, in competition with the Tevatron, the $q\bar{q}$-channel is definitely less favorable with respect to the $gg$-, $qg$- and $qq$-channels. Considering for simplicity the lower spin and SM representations there are only few states that can be produced in the channels mentioned above \cite{Barbieri:2011p2759}:
\begin{itemize}
\item {\bf $gg$-channel}: a spinless totally neutral scalar $S$;
\item {\bf $qg$-channel}: a $J=1/2$ color-triplet "heavy quark", either "U" or "D";
\item {\bf $qq$-channel}: a spinless color-triplet or color-sextet $\phi$, with various possible charges.
\end{itemize}
\indent In our phenomenological study we always assume that only a single new particle is available at a time, which therefore can only decay into SM particles. Moreover, we concentrate our attention on the first two cases, since the scalar triplet or sextet can only decay into pair of jets \cite{Bauer:2010p280}, consistenly with known constraints, whereas we find relatively more promising the final states containing at least one photon. The resonaces in the $qq$-channel also suffer of problems with flavor physics \cite{Barbieri:2011p2759}. For an estensive analysis of all the possible colored resonances at the early LHC see Ref.~\cite{Han:2010p2696}.
\section{Neutral scalar singlet $S$}
A reasonable effective Lagrangian describing the interactions of a neutral scalar singlet $S$ of mass $M_{S}$ with the SM particles is
\begin{equation}\label{eq1}
\mathcal{L}_{S}=c_{3}\frac{g_{S}^{2}}{\Lambda}G_{\mu\nu}^{a}G^{\mu\nu\,a}S+c_{2}\frac{g^{2}}{\Lambda}W_{\mu\nu}^{i}W^{\mu\nu\,i}S+c_{1}\frac{g^{\prime 2}}{\Lambda}B_{\mu\nu}B^{\mu\nu}S+\sum_{f}c_{f}\frac{m_{f}}{\Lambda}\bar{f}fS\,,
\end{equation}
where $\Lambda$ is an energy scale, $c_{i}$, $i=1,2,3,f$ are dimensionless constants and $f$ is any SM fermion, of mass $m_{f}$. \\
\indent The Branching Ratios (BR) of $S$ are independent on the scale $\Lambda$.
The total width of $S$ depends on the parameters $M_{S}$, $\Lambda$ and $c_{i}$ and scales as $1/\Lambda^{2}$. We take as reference values $\Lambda= 3$ TeV, reminiscent of a new strong interaction possibly responsible for ElectroWeak Symmetry Breaking (EWSB) at $\Lambda \approx 4\pi v$ and $c_{i}=1$. If $S$ were a composite particle generated by a new strong dynamics, the Naive Dimensional Analysis (NDA) \cite{Georgi:1984p3322} would suggest $c_{i}=1/4\pi$. However, larger values could arise from large $N$ and/or from more drastic assumptions about the nature of the gauge bosons. For the reference values of the parameters we are considering, $\Gamma_{S}$ ranges from about $10$ GeV at $M_{S}=500$ GeV to about $250$ GeV for $M_{S}=1.5$ TeV.\\
\indent We have made a preliminary study of the sensitivity to the search for $S$ in the di-$jet$ and $\gamma\gamma$ final states (see Ref.~\cite{Barbieri:2011p2759} for more details).
We have found that, while the large systematic uncertainties and the low signal over background ratio seem to disfavor an emerging signal in the di-$jet$ channel, at least with the first run integrated luminosity ($1-5$ fb$^{-1}$), a discovery in the $\gamma\gamma$-channel looks possible with a modest integrated luminosity of about $10-100$ pb$^{-1}$.
All these considerations can be trivially extended to arbitrary values of $\Lambda$ and $c_{i}$. In particular, assuming the dominance of the $S\to gg$ channel and the validity of the Narrow Width Approximation\footnote{The NWA allows one to write the total cross section as the product of the production rate times the BR: \mbox{$\sigma\(pp\to X\to 2x\)\approx \sigma\(pp\to X\)\times BR\(X\to 2x\)$}.} (NWA), an absence of signal for a given integrated luminosity is easily translated in a lower bound on $\Lambda/c_{3}$ for the di-$jet$ channel and $\Lambda/\(c_{1}+c_{2}\)$ for the $\gamma\gamma$ channel.\\
\indent We have used the recent search for a narrow resonance in the di-$jet$ channel performed by the CMS experiment\footnote{We could not use in this case the corresponding search of ATLAS \cite{ATLASCollaboration:2010p1746} since it was made assuming the $qg$ final state that gives rise to a different quantity of QCD radiation.} \cite{CMSCollaboration:2010p2595} to set an upper bound on the coupling $c_{3}$ for fixed $\Lambda=3$ TeV as a function of $M_{S}$. The upper bound on the total cross section compared with the prediction for our reference values of the parameters and the corresponding bound on the coupling $c_{3}$ are depicted in Fig.~\ref{Scalar_3}.%
\begin{figure}[t]
\centering
\includegraphics[scale=0.24]{Figures/Exclusion_scalar_dijet_1.pdf}\hspace{3mm}
\includegraphics[scale=0.24]{Figures/Exclusion_scalar_dijet_2.pdf}
\caption{Left panel: experimental upper limit at $95\%$ C.L. on the total cross section for a $gg$ resonance measured by the CMS experiment and total cross section predicted with the Lagrangian \eqref{eq1} for the scalar $S$, as functions of the resonance mass; Right panel: upper bound at $95\%$ C.L. on the coupling $c_{3}$ as a function of the resonance mass for $\Lambda=3$ TeV, assuming $BR\(S\to gg\)\approx 1$.}
\label{Scalar_3}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[scale=0.26]{Figures/Exclusion_scalar_gammagamma.pdf}
\caption{Upper limits on $c_{1}+c_{2}$ as functions of the scalar mass $M_{S}$ for $\Lambda=3$ TeV assuming $BR\(S\to \gamma\g\)\ll BR\(S\to gg\)$.}
\label{Scalar_4}
\end{figure}%
In Fig.~\ref{Scalar_4} we also show a preliminary estimation of the senstivity of the early LHC to the couplings combination $c_{1}+c_{2}$ relevant for the $\gamma\gamma$ final state. From this figure we see that an integrated luminosity of $1$ to $10$ fb$^{-1}$ is required to approach the NDA limit $1/4\pi$ for the couplings $c_{1}$ and $c_{2}$. Note that the upper limits in Figs.~\ref{Scalar_3} and \ref{Scalar_4} were obtained assuming the NWA. However, when the couplings become large ($g_{i}\gg1$), the resonance can become significantly broad, making the NWA no longer reliable. Moreover, threshold effects can spoil the NWA in the large resonance mass region \cite{Grojean:2011vu}. For these reasons the limit in the high mass region can differ significantly from those in Figs.~\ref{Scalar_3} and \ref{Scalar_4}.
\section{Heavy quark $U$}
The same pheomenological study we have made for the scalar singlet can be done for an heavy quark $U$ trasforming as a $\(3,1\)_{2/3}$ of the SM gauge group and coupling to the $qg$ initial state. A reasonalble phenomenological Lagrangian is
\begin{equation}\label{eq2}
\mathcal{L}_{U}=c_{G}\frac{g_{S}}{\Lambda}\bar{U}_{L}\sigma^{\mu\nu}T^{a}u_{R}G_{\mu\nu}^{a}+c_{B}\frac{g^{\prime}}{\Lambda}\bar{U}_{L}\sigma^{\mu\nu}u_{R}B_{\mu\nu}\,,
\end{equation}
where $\sigma^{\mu\nu}=i/2 [\gamma^{\mu},\gamma^{\nu}]$ and $T^{a}$ are the generators of the fundamental representation of $SU\(3\)$ ($T^{a}=\lambda^{a}/2$). As in the case of $S$, $U$ could be a composite state generated by a strong dynamics responsible for EWSB, in which case NDA suggests $\Lambda\approx 4\pi v\approx 3$ TeV and $c_{G}\approx c_{B}\approx 1/4\pi$. Problems with flavor can arise from the Lagrangian \eqref{eq2} and can be solved by making further assumptions on the new physics \cite{Barbieri:2011p2759}. We don't discuss them here since they don't affect the phenomenology we are interested in.
For the reference values of the parameters $\Lambda=3$ TeV and $c_{G}=c_{B}=1$, the width of the heavy quark $\Gamma_{U}$ ranges from about $2$ GeV for $M_{U}=500$ GeV to about $30$ GeV for $M_{U}=1.5$ TeV.\\
\indent We have made a preliminary study of the sensitivity to the search for $U$ in the di-$jet$ and $\gamma+jet$ final states (see Ref.~\cite{Barbieri:2011p2759} for more details).
We have found that the di-$jet$ channel is again strongly disfavored with respect to the one containing a photon. A discovery in the latter channel looks possible with a modest integrated luminosity of about $5-10$ pb$^{-1}$. Assuming the dominance of the $U\to qg$ channel and the validity of the NWA, an absence of signal for a given integrated luminosity is easily translated in a lower bound on $\Lambda/c_{G}$ for the di-$jet$ channel and $\Lambda/c_{B}$ for the $\gamma+jet$ channel.\\
\indent In this case we have used both the searches of the CMS \cite{CMSCollaboration:2010p2595} and the ATLAS \cite{ATLASCollaboration:2010p1746} experiments in the di-$jet$ channel to set an upper bound on the coupling $c_{G}$ (for fixed $\Lambda=3$ TeV) as a function of $M_{U}$. The upper bounds on the total cross section compared with the prediction for our reference values of the parameters and the bounds on the corresponding coupling $c_{G}$ are depicted in Fig.~\ref{Fermion_3}.%
\begin{figure}[t]
\centering
\includegraphics[scale=0.24]{Figures/Exclusion_fermion_dijet_1.pdf}\vspace{3mm}
\includegraphics[scale=0.24]{Figures/Exclusion_fermion_dijet_2.pdf}
\caption{Left panel: experimental upper limits at $95\%$ C.L. on the total cross section for a $qg$ resonance measured by the CMS (blue) and ATLAS (red) experiments and cross sections predicted with the Lagrangian \eqref{eq2} for the heavy quark $U$, as functions of the resonance mass; Right panel: upper bound at $95\%$ C.L. on the coupling $c_{G}$ as function of the resonance mass for $\Lambda=3$ TeV, assuming $BR\(S\to qg\)\approx 1$.}
\label{Fermion_3}%
\end{figure}
\begin{figure}[t!]
\centering
\includegraphics[scale=0.26]{Figures/Exclusion_fermion_gammajet.pdf}
\caption{Upper limits on $c_{B}$ as functions of the heavy quark mass $M_{U}$ for $\Lambda=3$ TeV assuming $BR\(S\to q\gamma\)\ll BR\(S\to qg\)$.}
\label{Fermion_4}
\end{figure}%
In Fig.~\ref{Fermion_4} we also show a preliminary estimation of the senstivity of the early LHC to the coupling $c_{B}$ relevant for the $\gamma+jet$ final state. From this figure we can see that the NDA coupling $1/4\pi$ can be excluded, in a wide region of $M_{U}$, with an integrated luminosity of about $1$ to $10$ fb$^{-1}$.
\section{Acknowledgments}
This work was done in collaboration with Riccardo Barbieri. I want to thank Michelangelo Mangano, Ennio Salvioni and Paolo Francavilla for useful discussions. This work was partially supported by the European Program {\it Unification in the LHC Era}, contract PITN-GA-2009-237920 (UNILHC) and by the European Program {\it Advanced Particle Phenomenology in the LHC Era}, contract PITN-GA-2010-264564 (LHCPhenoNet).
\deffdp{fdp}
\bibliographystyle{JHEP}
|
\section{Introduction}
Spin coherence is of key importance to quantum devices utilizing spin degrees of freedom, e.g., a weak field magnetometer based on spin-1 Bose-Einstein condensates (BECs), which requires the spin coherence time to be as long as possible in order to improve the device sensitivity~\cite{Vengalattore07}. However, due to the ferromagnetic spin exchange interaction in the $^{87}$Rb Bose condensates, the intrinsic dynamical instability causes the unstable collective modes to grow exponentially into a spin-domain or spin-texture structure and eventually sabotages the spin coherence of the condensates~\cite{Chang05, Zhang2005PRL, Gu07, Vengalattore08, Kronjager10, Matuszewski10}. It is highly demanded to find a practical way to enhance the spin coherence in spinor Bose condensates in order to develop a higher precision magnetometer. In fact, the spin coherence is enhanced in a $^87$Rb spin-1 condensate by localizing the spin dynamics, which was proposed to unmask other interesting effects induced by the weak magnetic dipolar interactions~\cite{Zhang2010, Vengalattore08}.
Some experimental techniques may be employed to manipulate the atomic spin dynamics of a spinor condensate. One is the magnetic radio-frequency pulses, which can rotate the spin in an arbitrary angle along a certain direction~\cite{Bloch06,Vengalattore08}. Unfortunately, the spin exchange interaction in spin-1 condensates is rotationally invariant so that the magnetic pulses are not able to control the spin-exchange-interaction induced effects. The other is the optical Feshbach resonance technique, which can adjust both the sign and the magnitude of the spin exchange coefficient $c_2 = 4\pi\hbar^2 (a_2-a_0)/ 3m $ in a certain parameter range~\cite{Hamley09}, where $a_{0,2}$ is the $s$-wave scattering length and $m$ the mass of the atom. Theoretically, it has been shown that the dynamical instability of the $^{87}$Rb spin-1 condensate can be suppressed by periodically changing the spin exchange coefficient in a cosine form~\cite{Zhang2010}.
From the viewpoint of quantum control theory, other forms of modulating the spin exchange coefficient $c_2(t)$ are not only easy to realize in experiments but also powerful to suppress the spin decoherence. Recently, Uhrig proposed an optimal sequence with $N$ $\pi$-pulses to dynamically decouple a qubit from its boson environment in a spin-boson model~\cite{Uhrig07}. The Uhrig dynamical decoupling (UDD) sequence has also been applied in a system of a center spin coupled with many nuclear spins~\cite{Lee08, Yang08}. The results~\cite{Yang08} show that the UDD sequence can be used to remove the dephasing terms up to $\mathcal{O}(T^{N+1})$, where $T$ is the sequence period. This has aroused considerable experimental interests to demonstrate the advantages of UDD (see, e.g.,~\cite{Biercuk09, Du09, Lange10}). Moreover, the concatenated dynamical decoupling (CDD) sequence has also been demonstrated to be powerful in suppressing the decoherence in quantum systems~\cite{Khodjasteh05, Zhang07a, Yao07}. Despite of the success of these dynamical decoupling (DD) sequences, they have never been tested in a Bose-Einstein condensate, where the decoherence is dominated by a very different bath consisting of unstable collective modes.
In this paper, we theoretically investigate periodic dynamical decoupling (PDD), CDD, and UDD sequences on preserving the spin coherence of a homogenous $^{87}$Rb spin-1 Bose condensate. By comparing the performance of these sequences, we identify the optimal sequence in suppressing the dynamical instability of the spin-1 Bose condensate. Our results show that all sequences suppress the dynamic instability and the wave vector of the most unstable collective modes $k_-$ is pushed to a larger value. Although the CDD sequences behave better than the PDD or UDD sequences at the same number of pulses, the same law $k_{-}T^{1/2} = c$, with $c$ being a sequence-dependent constant, is followed for all the three sequences. Moreover, in a scalar Bose condensate, our analytical result for a PDD sequence further verifies this law between $k_{-}$ and $T$.
The paper is organized as follows. In Sec.~\ref{sec:dd}, we formulate the spin dynamics of a spin-1 condensate subjecting to a dynamical decoupling pulse sequence and describe our metric to measure the performance of the sequence. In Sec.~\ref{sec:num}, we present numerical results for a spin-1 condensate and compare the performances of PDD, UDD, and CDD sequences. To understand our numerical results, we analytically investigate, in Sec.~\ref{sec:scalar}, a simple case of a scalar condensate subjecting to a PDD sequence. The conclusion is given in Sec.~\ref{sec:con} and a brief derivation of our analytical results is provided in Appendix~\ref{sec:apd}.
\section{Spin dynamics of a spin-1 condensate subjecting to dynamical decoupling sequences}
\label{sec:dd}
In the mean-field theory, the evolution of a homogeneous three-component spin-1 Bose condensate is described by three coupled Gross-Pitaevski equations~\cite{Ho1998, Ohmi1998, Law1998}
\begin{eqnarray}
\label{Eq:1}
i\hbar\frac{\partial}{\partial t}\Phi_{\pm1} &=& [\mathcal{H}_0+c_{2}(n_{\pm1}+n_{0}-n_{\mp1})]\Phi_{\pm1}+c_{2}\Phi^{2}_{0}\Phi^{*}_{\mp1} \nonumber \\
i\hbar\frac{\partial}{\partial t}\Phi_{0} &=& [\mathcal{H}_0+c_{2}(n_{+}+n_{-})]\Phi_{0}+2c_{2}\Phi_{+1}\Phi^{*}_{0}\Phi_{-1},
\end{eqnarray}
where $\mathcal{H}_0=-(\hbar^{2}/2m)\nabla^{2}+V_{ext}+c_{0}n$ is the spin-independent Hamiltonian. In the above equations, $c_{0}=4\pi\hbar^{2}(a_{0}+2a_{2})/(3m)$ represents the spin-independent interaction and $n_i({\bf r})=|\Phi_i({\bf r})|^{2}$ is the total density of the condensate with $i=0,\pm 1$ and $\Phi_i$ the condensate wave function of the $i$th component.
The spin dynamics of the spin-1 condensate is determined by the rest spin exchange terms containing $c_2$. Depending on the sign of $c_2$, the spin exchange interaction can be either ferromagnetic ($c_2 < 0$, e.g., $^{87}$Rb condensates) or antiferromagnetic ($c_2 > 0$, e.g., $^{23}$Na condensates)~\cite{Ketterle1998,Stamper1998, Chapman2001, Sengstock04, Chang04, Kuwamoto04, Artur05,Lett07}. It was shown that a ferromagnetically interacting spin-1 Bose condensate is dynamically unstable while an antiferromagnetically interacting one is dynamically stable~\cite{Zhang2005PRL}. In other words, the spin coherence of a ferromagnetically interacting spin-1 condensate is lost, since some collective modes grow exponentially due to the intrinsic dynamical instability.
According to the Bogoliubov approximation, the time dependent equations of motion for the collective modes in a condensate can be cast in a matrix form as follows~\cite{Maldonado1993}:
\begin{equation}
\label{Eq:2}
i\hbar\frac{\partial{{\bf x}}}{\partial{t}}=M\cdot{\bf x},
\end{equation}
where ${\bf x} = (\delta\Psi_{+1},\delta\Psi_{0}, \delta\Psi_{-1}, \delta\Psi^{*}_{+1}, \delta\Psi^{*}_{0},\delta\Psi^{*}_{-1})^T$ denotes the amplitude of the collective modes and $M$ is an effective Hamiltonian. In general, $M$ has a matrix form as
\begin{equation}
\label{Eq:3}
M=\left(\begin{array}{cc}A & B \\-B^{*} & -A^{*}\end{array}\right),
\end{equation}
where for a spin-1 condensate
\begin{eqnarray}
A &=& \varepsilon_{k}I+c_{0}A_{0}+c_{2}A_{2}, \nonumber \\
B &=& c_{0}B_{0}+c_{2}B_{2}. \nonumber
\end{eqnarray}
Here, $\varepsilon_{k} = \hbar^2 k^2 / (2m)$ is the kinetic energy of the collective mode in a plane wave form with a wave vector $k$, $I$ is a $3\times3$ identity matrix, and $A_{0,2}, B_{0,2}$ are also $3\times3$ matrices given by
\begin{eqnarray}
A_0 &=& \left(
\begin{array}{ccc} n_{+1} & \Phi_{0}^{*}\Phi_{+1} & \Phi_{-1}^{*}\Phi_{+1} \\
\Phi_{0}\Phi_{+1}^{*} & n_{0} & \Phi_{0}\Phi_{-1}^{*}\\
\Phi_{-1}\Phi_{+1}^{*} & \Phi_{0}^{*}\Phi_{-1} & n_{-1}\\
\end{array}\right), \nonumber \\
A_2 &=& \left(
\begin{array}{ccc} n_{+1}+n_{0} & \Phi_{0}\Phi_{-1}^{*} & -\Phi_{-1}^{*}\Phi_{+1} \\
\Phi_{0}^{*}\Phi_{-1} & n_{+1}+n_{-1} & \Phi_{0}^{*}\Phi_{+1}\\
-\Phi_{-1}\Phi_{+1}^{*} & \Phi_{0}\Phi_{+1}^{*} & n_{-1}+n_{0}\\
\end{array}\right), \nonumber \\
B_0 &=& \left(
\begin{array}{ccc} n_{+1} & \Phi_{0}\Phi_{+1} & \Phi_{-1}\Phi_{+1} \\
\Phi_{0}\Phi_{+1} & n_{0} & \Phi_{0}\Phi_{-1}\\
\Phi_{-1}\Phi_{+1} & \Phi_{0}\Phi_{-1} & n_{-1}\\
\end{array}\right), \nonumber \\
B_2 &=& \left(
\begin{array}{ccc} n_{+1} & \Phi_{0}\Phi_{+1} & n_{0}-\Phi_{-1}\Phi_{+1} \\
\Phi_{0}\Phi_{+1} & 2\Phi_{-}\Phi_{+} & \Phi_{0}\Phi_{-1}\\
n_{0}-\Phi_{-1}\Phi_{+1} & \Phi_{0}\Phi_{-1} & n_{-1}\\
\end{array}\right). \nonumber
\end{eqnarray}
Obviously, matrix $M$ is neither Hermitian nor symmetric although $A_{0,2}$ are Hermitian and $B_{0,2}$ are symmetric. However, the matrix $M$ has paired eigenvalues with the same amplitude but different sign, due to its particular form~\cite{Maldonado1993}. The dynamical instability of a condensate is manifested as the existence of at least one pair of complex eigenvalues~\cite{Zhang2005PRL}.
By employing the optical Feshbach resonance technique, the spin exchange coefficient $c_{2}$ can be tuned not only in amplitude but also in sign~\cite{Hamley09}. It occurs naturally to utilize such a technique to suppress the dynamical instability in a ferromagnetically interacting spin-1 Bose condensate. Previous theoretical investigation has indeed demonstrated the suppression effect by modulating $c_2(t)$ in a simple cosine form~\cite{Zhang2010}.
From the viewpoint of the control theory, there are many more complicated DD pulse sequences which may outperform the simple cosine modulation of $c_2(t)$ in suppressing the dynamical instability and enhancing the spin coherence. In addition, these pulse sequences are easier to implement in experiments. Hereafter, applying a dynamical decoupling pulse means alternating the sign of $c_2$ but keeping its amplitude unchanged.
A PDD sequence denotes flipping the sign of $c_2$ back and forth periodically [see top panel in Fig.~\ref{Fig:1}]. By concatenating PDD sequences at different levels, we obtain the CDD sequences [see middle panel in Fig.~\ref{Fig:1}]. A UDD sequence [see bottom panel in Fig.~\ref{Fig:1}] containing $N$ pulses is defined as~\cite{Uhrig07, Lee08, Yang08}
\begin{equation}
\label{Eq:4}
\tau_{j}=\frac{T}{2} \left[\cos\left(\frac{\pi(j-1)}{N+1}\right) - \cos\left(\frac{\pi{j}}{N+1}\right)\right],
\end{equation}
where $\tau_{j}$ is the delay between the $(j-1)$th and $j$th pulses and $T=\sum_{j=1}^{N+1}\tau_{j}$ is the period of the total sequence. We notice that the CDD$_1$ (CDD at one-level concatenation), UDD$_1$ (one-pulse UDD), and PDD sequence are the same, and CDD$_2$ and UDD$_2$ are also the same.
\begin{figure}
\includegraphics[width=3.25in]{fig1.eps}
\caption{Various PDD (top), CDD (middle), and UDD (bottom) sequences. The vertical short lines mark the time of each applied pulse.}
\label{Fig:1}
\end{figure}
There are various choices of metric to measure the pulse sequence effect. We choose the wave vector of the most unstable mode, $k_-$, since this mode dominates in the dynamics of the spin-1 condensate~\cite{Zhang2005PRL}. The larger the value of $k_-$ is, the better the performance of the pulse sequence will be. For a spin-1 condensate under many DD pulses, we adopt a numerical method (se below) to obtain the $k_-$.
By applying a sequence of $N$ DD pulses, the time evolution operator $U_{T}$ of the collective modes of the spin-1 Bose condensate is
\begin{eqnarray}
U_{T} &\equiv & {\mathcal{T}}\exp\left[-\frac{i}{\hbar}\int^{T}_{0} dt' M(t')\right] \nonumber \\
& = & \exp[-iM_{\pm}\tau_{N+1}] \ldots\exp[-iM_{\mp}\tau_{2}] \exp[-iM_{\pm}\tau_{1}], \nonumber \\
\label{Eq:5}
\end{eqnarray}
where ${\mathcal{T}}$ denotes the time-ordering operator, and $M_{\pm}$ denotes the $M$ matrix with $c_{2}$ being either positive or negative. By diagonalizing $U_T$ and reexpressing it as $U_{T}=V^{\dagger}\exp(-iTF)V$ with $F$ a $6\times 6$ diagonal matrix, we identify the wave vector of the most unstable mode $k_-$ by searching the largest imaginary part of $F$ for different wave vector $k$~\cite{Zhang2010}.
\section{Numerical results for a spin-1 Bose condensate}
\label{sec:num}
\begin{figure}
\includegraphics[width=3.4in]{fig2a.eps}
\includegraphics[width=3.3in]{fig2b.eps}
\caption{(Color online) Dependence of the imaginary part of the Floquet operator $F$ on the wave vector $k$ and the period $T$ for sequences (a) PDD$_3$, (b) UDD$_5$, and (c) CDD$_3$. Lower panels show the $k$ dependence for (d) PDD$_3$, (e) UDD$_5$, and (f) CDD$_3$ at $T=T_{0}/20$ which is marked by the vertical white lines in panel (a), (b), and (c), respectively. Here, we use parameters~\cite{Pu1999,Zhang2010} $k_0=1$ and $T_0=\pi/\left|c_{2}n\sqrt{n_{0}(1-n_{0})}\cos(\theta_{0}/2)\right|=\pi/(0.4|c_2|n)$, where the initial condition is $\theta_{0}=0$ and $n_0= 0.8$. In calculations we have set $c_0=1$, $n=1$, and $c_2=-0.0046$ which are the parameters of $^{87}$Rb spin-1 condensates.}
\label{Fig:2}
\end{figure}
Following the procedure described at the end of Sec.~\ref{sec:dd}, we numerically obtain the dependence of the unstable modes on the DD sequence periods for a spin-1 Bose condensate. For the numerical simplicity, we set $m=1$ and $\hbar=1$. The results are presented in Fig.~\ref{Fig:2} and Fig.~\ref{Fig:3}(a). Several interesting observations are listed as follows:
\newcounter{Lcount}
\begin{list}{\roman{Lcount})}
{\usecounter{Lcount}}
\item Compared to the free evolution ($T=T_0$), all the DD sequences suppress the dynamical instability of the spin-1 condensate in three ways, narrowing the area of the unstable region, shifting $k_-$ to a larger value, and decreasing the value of $F_-=Im(F)|_{k_-}$ (the exponential growth rate of the most unstable mode).
\item Comparing the performances of the same DD sequence at different period $T$, $F_-$ is almost the same but $k_-$ is shifted to a larger value when decreasing $T$.
\item At the same number of pulses ($N=5$) and the same sequence period, the UDD sequence shifts $k_-$ the least, while the CDD and PDD sequences behave almost the same (but both better than the UDD) [see Fig.~\ref{Fig:3}(a)]. Moreover the CDD sequence suppresses the growth rate $F_-$ more than the PDD [see Figs.~\ref{Fig:2}(d) and ~\ref{Fig:2}(f)].
\item The relationship between $k_-$ and $T$ for all PDD, UDD, and CDD sequences follows exactly the same power law: $k_- T^{1/2} = c$, where $c$ is a sequence-dependent constant and can be obtained by numerically fitting the slope of the lines in Fig.~\ref{Fig:3}(a)\label{lcout4}.
\end{list}
The above results clearly indicate that the dynamical instability in a ferromagnetically interacting spin-1 Bose condensate is essentially inhibited once the wavelength of the most unstable mode $2\pi/k_-$ is less than the spin healing length $\xi=h/\sqrt{2m|c_2|n}$. Our results shows that CDD sequences most effectively suppress the dynamical instability and they are the best choice to enhance the spin coherence in a spin-1 Bose condensate.
It is not easy to fully understand all the above numerical results for a spin-1 Bose condensate in a simple picture, especially the result in (\romannumeral4), because it is difficult to obtain analytical results with either the effective Hamiltonian $M$ matrix [Eq.~(\ref{Eq:3})] or the $F$ matrix. To circumvent these difficulties, we investigate the effect of the same DD sequences but in a scalar Bose condensate, where it becomes possible for us to obtain the same behaviors but in an analytical way.
\section{results for scalar Bose condensate}
\label{sec:scalar}
Similar to the case of a spin-1 Bose condensate, we also focus on the change of $k_{-}$ for the most unstable mode of a scalar condensate subjecting to DD pulses, which alternates the sign of the interaction coefficient $g=4\pi\hbar^2 a_s/m$, with $a_s$ the $s$-wave scattering length. The interaction between atoms is repulsive if $g>0$ and attractive if $g<0$~\cite{Dalfovo}. For a homogeneous scalar condensate with attractive interaction, the system is intrinsically unstable~\cite{Kagan}. By periodically flipping the sign of $g$, it is possible to suppress the instability of an attractively interacting Bose condensate~\cite{Zhang2010}. Fortunately, both the sign and the amplitude of $g$ can be precisely tuned in experiments via the Feshbach resonance technique~\cite{Roberts1998,Roberts2001}.
The time-dependent effective Hamiltonian $M(t)$ matrix of the collective modes in the scalar condensate modulated by a PDD$_1$ sequence is given by
\begin{equation}
M(t) =\left\{\begin{array}{cc}
M_+, & t \in \left[0, \frac T 2\right), \\
M_-, & t \in \left[\frac T 2, T\right),
\end{array} \right. \nonumber
\end{equation}
where
\begin{equation}
M_{\pm }=\left(\begin{array}{cc}
\varepsilon_{k}\pm gn & \pm gn \\
\mp gn & -(\varepsilon_{k}\pm gn)
\end{array}\right).
\label{Eq:6}
\end{equation}
Following exactly the same procedures in Eq.~(\ref{Eq:5}) for the above $M$ matrix, we first diagonalize the corresponding time evolution operator $U_T$, then find out the eigenvalues of the Flouqet operator $F$, and finally extract the mode containing largest imaginary part $F_-$ and the corresponding wave vector $k_-$. For a scalar Bose condensate, the matrix $M(t)$ is much simpler than the one in a spin-1 condensate, so it is possible for us to obtain analytical results for a PDD$_1$ sequence. For convenience, we reform the matrix $M_\pm$ in Pauli matrices as
\begin{equation}
M_{\pm}=(\varepsilon_{k}\pm gn)\cdot\sigma_{z}\pm ign\cdot\sigma_{y}.
\label{Eq:7}
\end{equation}
By utilizing the properties of Pauli matrices, it is straightforward to obtain the time evolution operator
\begin{eqnarray}
U_{\pm} &\equiv & \exp(-i\tau M_{\pm}) \nonumber \\
&=& \cos(\tau{x_{\pm}}) - \frac{i}{x_{\pm}} [(\varepsilon_{k}\pm{gn})\cdot\sigma_{z}\pm ign\cdot\sigma_{y}] \sin(\tau x_{\pm}), \nonumber
\end{eqnarray}
where $x_{\pm}=\sqrt{\varepsilon_{k}(\varepsilon_{k}\pm 2gn)}\;$ and $\tau=T/2$. The propagator in a complete period is
\begin{equation}
U_{T}=U_{-}U_{+}=\left(\begin{array}{cc}P-iQ_{z} & -Q_{x}+iQ_{y}\\
-Q_{x}+iQ_{y} & P+iQ_{z} \end{array}\right),
\label{Eq:8}
\end{equation}
and the corresponding eigenvalues of $U_{T}$ is
\begin{equation}
\lambda=P\pm\sqrt{Q_{x}^{2}+Q_{y}^{2}-Q_{z}^{2}},
\label{Eq:9}
\end{equation}
where
\begin{eqnarray*}
P&=&\cos(\tau x_{-})\cos(\tau x_{+}) - \frac{\varepsilon_{k}^{2}}{x_{-}x_{+}} \sin(\tau x_{-}) \sin(\tau x_{+}), \\
Q_{x}&=&\frac{2\varepsilon_{k}gn}{x_{-}x_{+}} \sin(\tau x_{-})\sin(\tau x_{+}),\\
Q_{y}&=&\frac{gn}{x{-}}\sin(\tau x_{-})\cos(\tau x_{+}) - \frac{gn}{x_{+}} \cos(\tau x_{-})\sin(\tau x_{+}),\\
Q_{z}&=&\frac{\varepsilon_{k}-gn}{x_{-}} \sin(\tau x_{-})\cos(\tau x_{+}) \\
&& + \frac{\varepsilon_{k}+gn}{x_{+}} \cos(\tau x_{-})\sin(\tau x_{+}).
\end{eqnarray*}
Since $\lambda=\exp(-iTF)$, the imaginary part of the Flouquet operator follows as
\begin{equation}
Im(F)=\frac{1}{T}\ln{|\lambda|}.\label{Eq:10}
\end{equation}
\begin{figure}
\includegraphics[width=3.25in]{fig3a.eps}
\includegraphics[width=3.25in]{fig3b.eps}
\caption[]{\label{Fig:3}(Color online) Numerical results of power law relationship between the wave vector $k_-$ of the most unstable mode and the sequence period $T$ for (a) spin-1 condensates and (b) scalar condensates: open squares --- CDD$_3$, crosses --- PDD$_3$, filled circles --- UDD$_5$, filled triangles --- PDD$_1$. The specific fitting results are shown in Table~\ref{tab1}. The parameters for the scalar condensates are $g=-0.0046$ and $n=1$.}
\end{figure}
We notice that the kinetic energy $\varepsilon_k$ of the most unstable mode is much larger than the interaction energy $gn$ if the PDD$_1$ sequence period $T$ is small. By assuming $gn \ll \varepsilon_{k}$ and keeping to the second order of $gn / \varepsilon_{k}$, we analytically obtain the relationship between the wave vector $k_{-}$ of the most unstable mode and the sequence period $T$ from Eqs.~(\ref{Eq:8})-~(\ref{Eq:10}) as
\begin{equation}
k_{-} T^{1/2} = c,
\label{Eq:11}
\end{equation}
with $c=\sqrt{2\pi}$. For a detailed derivation of Eq.~(\ref{Eq:11}), see Appendix~\ref{sec:apd}.
\begin{table
\caption{\label{tab1}Numerical fitting values of the slope, intercept and calculated $c^2/2\pi$ for lines in Fig.~\ref{Fig:3} for spin-1 condensates [denoted below as (a)] and scalar condensates [denoted as (b)]. In calculations, $c^2 = \pi \exp(2y_0) /\left[|c_2|n\sqrt{n_0(1-n_0)}\right]$ for spin-1 condensates and $c^2 = 2\pi \exp(2y_0)/(|g|n)$ for scalar condensates.}
\begin{ruledtabular}
\begin{tabular}{cccc}
& slope & intercept ($y_0$) & $c^2/2\pi$ \\
\colrule
(a) PDD$_1$& -0.473 & -3.051 & 0.608\\
(a) UDD$_5$& -0.499 & -2.799 & 1.006\\
(a) PDD$_3$& -0.495 & -2.583 & 1.551\\
(a) CDD$_3$& -0.495 & -2.583 & 1.551\\
\toprule
(b) PDD$_1$& -0.498 & -2.681 & 1.020\\
(b) UDD$_5$& -0.499 & -2.341 & 2.013\\
(b) PDD$_3$& -0.499 & -2.137 & 3.027\\
(b) CDD$_3$& -0.499 & -2.137 & 3.027\\
\end{tabular}
\end{ruledtabular}
\end{table}
It is difficult to obtain the analytical results for other more complex sequences such as UDD$_5$, PDD$_3$ and CDD$_3$, so we carry out numerical calculations for the scalar condensates. The results are shown in Fig.~\ref{Fig:3}(b). It is clearly shown in Fig.~\ref{Fig:3} that no matter in a spin-1 condensate or in a scalar condensate, all sequences exhibit the same relationship between $k_-$ and $T$: $k_{-} T^{1/2} = c$, with $c$ a sequence-dependent constant (see Table~\ref{tab1}).
\section{Conclusion}
\label{sec:con}
In conclusion, we have theoretically investigated the PDD, CDD and UDD pulse sequences on suppressing the unstable collective modes in a spin-1 Bose Einstein condensate. Our numerical results show that the three DD sequences can be used to effectively protect the coherence of the spin-1 condensate. The wave vector $k_-$ of the most unstable mode is shifted to a much larger value, which in fact enhances the spin coherence of the condensate. Furthermore, using the analytical results in a scalar condensate, we reveal an interesting relation between $k_{-}$ and sequence period $T$: $k_- T^{1/2} = c$, where $c$ is a sequence-dependent constant.
\acknowledgments
BYN and WZ acknowledge support by the National Natural Science Foundation of China Grant No. 10904017, NCET, Specialized Research Fund for the Doctoral Program of Higher Education of China under Grant No. 20090071120013, and Shanghai Pujiang Program under Grant No. 10PJ1401300. JQY and WZ acknowledge support by National Basic Research Program of China Grant No. 2009CB929300.
|
\section{\label{}}
Quantum dot (QD) nanostructures formed by strain driven self-assembly are ideal for solid state quantum optics experiments due to their discrete optical spectrum, strong interaction with light and robust quantum coherence for both interband polarization\cite{Borri01,Borri03} and spin\cite{Greilich09}. The ease with which such nanostructures can be embedded into electrically active devices allows for tuning of the transition frequency and control of charge occupancy \cite{Warburton00}. Self-assembly provides a natural way to realize few dot systems via vertical stacking to produce more sophisticated nanostructures with coherent inter-dot coupling due to carrier tunneling\cite{Stinaff2006, PhysRevLett.97.076403, Doty2008, Kim2010,PhysRevLett.94.057402, Bracker2006, Scheibner2007biex, Scheibner2008}. When combined with the potential to coherently manipulate excitons over ultrafast timescales using precisely timed laser and electrical control pulses \cite{Zrenner02,Zecherle2010,Vasconcellos10} such systems raise exciting prospects for the operation of small scale few qubit systems in a solid-state device. Very recently, conditional quantum dynamics for a single resonantly driven QD-molecule (QDM)\cite{Robledo2008} and spin dependent quantum jumps have been observed\cite{Kim2010, Atatuere2010}.
In this paper we employ photocurrent (PC) absorption, photoluminescence (PL) emission and PL-excitation (PLE) spectroscopy to trace the spectrum of ground and excited state transitions of an individual self-assembled QD-molecule as their character is electrically tuned from spatially direct to indirect. PC absorption allows us to identify the spatially direct neutral exciton transitions in both the upper ($X_{ud}$) and lower ($X_{ld}$) dots in the molecule. A number of excited state transitions are identified in PLE $\sim8-16$ meV above $X_{ud}$. These excited states exhibit pronounced anticrossings (energy splitting $\Delta E \sim 3.2-3.5$ meV) as the electric field $F$ is tuned. Excited state transitions are identified from voltage dependent PLE measurements to correspond to transitions between these hybridized electronic states and different \textit{hole} orbitals in the upper dot. By performing a multi-color experiment where the QDM is simultaneously excited with different frequency lasers, we demonstrate how the resonant excitation of indirect excitons or excitons in the lower quantum dot can be used to suppress the resonant excitation of the upper quantum dot, due to interdot Coulomb interactions. An on-off gating contrast up to 88$\%$ is observed, demonstrating a \textit{conditional} optical response of an artificial molecule.
\begin{figure}
\includegraphics[width=1\columnwidth]{Figure_1}
\caption{\label{fig:Figure_1}
(Color online)
Combined result of an electric field dependent PL and PC measurement: The PL of the QDM as a function of the applied electric field is shown as a contour plot from 0 to 130cps from white to black on a logarithmic scale. The dashed red lines indicate the uncoupled neutral exciton in the upper dot and the indirect exciton as depicted schematically in the inset. The PC resonances are shown as red squares and blue triangles for the neutral exciton transitions in the upper and lower dot, respectively.}
\end{figure}
The sample consists of vertically stacked pairs of QDs separated by a 10nm thick GaAs spacer and embedded within the intrinsic region of a GaAs Schottky photodiode \cite{PhysRevLett.94.057402}. Typical PL and PC measurements recorded at T=4.2K are presented in Fig.1. For the PL measurement the sample was excited in the wetting layer at 1.49 eV. Typical electric field dependent PL from 18 to 32 kV/cm are presented in a greyscale contour plot representation in Fig.1. The measurements show an anticrossing of two transitions arising from spatially direct and indirect excitons in the QDM where the hole is located in the upper dot \cite{PhysRevLett.94.057402}. These optical transitions are depicted schematically in a single particle picture in the inset of Fig.1: the direct exciton in the upper dot $X_{ud}$ and the indirect exciton $X_{ind}$ with the hole in the upper dot and the electron in the lower dot. The indirect exciton exhibits a strong Stark shift due to the large static dipole. As the energies of the two states are tuned to resonance, electron mediated tunnel coupling occurs that results in the formation of molecular bonding (lower energy) and antibondig (higher energy) orbitals \cite{ab-b} and the observed anticrossing \cite{PhysRevLett.97.076403}. As $F$ increases beyond $\sim 25$ kV/cm, the intensity of the luminescence reduces as charge carriers escape from the QDM via tunneling and PC measurements can be performed with resonant optical excitation. Two prominent resonances are observed in PC, examples of which are presented in the left panel of Fig.1. The energy of these peaks are plotted in the main panel of Fig.1 for F=25-31 kV/cm. The transitions observed in PC arise from charge neutral excitons and show that the QD molecule exhibits \textit{two} neutral exciton absorption resonances. As $F$ decreases these two resonances clearly evolve into the two clear peaks observed in PL, labeled $X_{ld}$ and $X_{ud}$ on Fig.1. The state marked as $X_{ld}$ is attributed to an exciton in the lower quantum dot of the molecule (depicted in blue in the inset of Fig.1) while $X_{ud}$ is the direct exciton in the upper dot, assignments that are confirmed by the results presented below.
\begin{figure}
\includegraphics[width=1\columnwidth]{Figure_2}
\vspace{-15pt}
\caption{\label{fig:Figure_2}
(Color online)
(a) Typical PLE scan detecting on $X_{B}$. The inset compares PLE recorded at F= 23.2kV/cm detecting either $X_B$ (black) or $X_{AB}$ (red), respectively.
(b) Spectrum of QDM transitions determined by PL (black symbols) and PLE (blue, red symbols) States with bonding (antibonding) character are plotted as triangles (circles). (c) Energy separation ($\Delta E$) between corresponding anticrossing states, fits using equation \ref{the_equation}.}
\end{figure}
We conducted detailed PLE measurements to track the evolution of the excited state spectrum of the QDM as a function of $F$.
A typical PLE scan detecting on $X_{ud}$ is presented in Fig.2(a) with $F$ fixed close to the anticrossing (21.6 kV/cm). Several discrete electronic resonances are observed in this region, the first four of which are labeled $e_{0B}-h_1$, $e_{0B}-h_2$, $e_{0AB}-h_1$ and $e_{0AB}-h_2$, in Fig.2(a). This assignment anticipates the nature of these excited states corresponding to the electron being in the bonding (B) or antibonding (AB) molecular orbital while the hole occupies the first ($h_1$) or second ($h_2$) excited orbital state in the upper QD. These assignments are now justified by examining the electric field dependence of the excited state resonances.
Fig.2(b) shows the energy of the molecular ground states determined via PL and the first four excited states as a function of $F$ in the range 20-25 kV/cm measured using PLE. The first four excited states consist of two different pairs of lines colour coded by the red and blue symbols in Fig.2(b), each of which anticross at an electric field close to 22 kV/cm. To analyze the observed excited state anticrossings in more detail and compare to the anticrossing of $X_{B}$ and $X_{AB}$ observed in PL, we plot the energy separation ($\Delta E$) between bonding and antibonding state of the ground states and the two excited states anticrossings in Fig.2(c). Thereby, the energy difference is plotted using the same colour coding as the corresponding anticrossing in Fig.2(b).
\begin{table}
\begin{tabular}{|c|c|c|c|}
\hline
Hole state & $2V_{ee}$ (meV) & $F_0$ (kV/cm) & d (nm) \\ \hline
$h_0$ & $3.4 \pm 0.1$ & $23.1 \pm 0.1$ & $15.3 \pm 0.1$ \\ \hline
$h_1$ & $3.2 \pm 0.1$ & $22.4 \pm 0.1$ & $15.8 \pm 0.2$ \\ \hline
$h_2$ & $3.5 \pm 0.1$ & $22.2 \pm 0.1$ & $15.9 \pm 0.2$ \\ \hline
\end{tabular}
\caption{Results of the fits of $\Delta E$ from Fig.2(c) with eq.1}
\label{table1}
\end{table}
For all three anticrossings $\Delta E$ shows a similar hyperbolic behaviour that can be fitted with
\begin{equation}
\Delta E = \sqrt{(2V_{ee})^2 + (ed(F-F_0))^2}
\label{the_equation}
\end{equation}
where $V_{ee}$ denotes the interdot tunnel coupling strength, $F_0$ the field at which the states anticross and $ed$ is the equivalent static dipole moment of the indirect exciton. ($d$ is the distance between the centers of the electron and hole envelope functions.) Fits to the three anticrossings are presented as lines in Fig.2(c) and the extracted values of $F_0$ and $d$ are summarized in table \ref{table1}. For each anticrossing, the extracted values of $d$ vary only slightly and are fully consistent with the expected electron-hole separation for the indirect exciton, since the dots have a typical height of 5 nm and a nominal separation of 10 nm for this sample. Both $V_{ee}$ and $F_0$ remain unchanged for the transitions involving $h_0$, $h_1$ and $h_2$, providing evidence that they arise from transitions between the same electron mediated anticrossing and different hole levels. This expectation is confirmed by PLE measurements performed close to $F_0$, detecting on either $X_{AB}$ or $X_B$ respectively. Typical results are presented in Fig.2(a)(inset). When detecting on $X_{AB}$ (red curve), transitions are only observed for $e_{0AB}-h_1$ and $e_{0AB}-h_2$, whilst $e_{0B}-h_1$ and $e_{0B}-h_2$ are absent. This arises since the electron populates the lower energy bonding level and thermal activation into the higher energy bonding level is unlikely since $2V_{ee}>>k_BT$ \cite{Nakaoka2006}. In contrast, upon exciting states with bonding electron character all four resonances are observed due to phonon mediated thermalization from antibonding to bonding electron states.\cite{Nakaoka2006} The small differences between $F_0$ arise from the Coulomb interactions between the various hole orbital states. Compared to the ground state with $F_0 = 23.1\pm0.1$ kV/cm the critical field of the excited state transitions are shifted by $-0.7\pm0.2$ kV/cm and $-0.9\pm0.2$ kV/cm. This can be converted to an energy difference of $1.1\pm0.3$ meV and $1.4\pm0.4$ meV using $d=15.8 \pm 0.2$ nm and $d=15.9 \pm 0.2$ nm, respectively. These values correspond to less than $10 \%$ of the total attractive e-h Coulomb interaction that has been estimated to be $\sim 22$ meV for similar samples.\cite{PhysRevLett.94.057402} Thus, the Coulomb shifts between the different valence band states represent a small pertubation and we conclude that these observations provide strong evidence that the four excited states shown in Fig.2(b) take place between an electron in either the lowest energy bonding or antibonding levels and different hole states in the upper dot, as depicted in the inset of Fig.2(b). The energy splitting between the lowest energy hole state and the first two excited states is 9.5 and 10.5 meV.
\begin{figure}
\includegraphics[width=1\columnwidth]{Figure_3}
\caption{\label{fig:Figure_3}
(Color online)
(a) PL intensity of $X_{ud}$ for the excitation of a resonance with direct character (pump, red), indirect character(block, blue) and both lasers (black). PL from the excitation of the direct resonance is quenched due to the presence of the block laser.
(b) Intensity of $X_{ud}$ as a function of the block laser energy. Whenever the block laser hits an excited state with indirect character the PL from the upper dot is decreased.}
\end{figure}
Away from resonance for $F < F_0$ the bonding ground state as well as the first two excited states $e_{0B}-h_1$ and $e_{0B}-h_2$ have predominant \textit{direct} character, while the third and fourth excited states $e_{0AB}-h_1$ and $e_{0AB}-h_2$ have \textit{indirect} character. Therefore, exciting $e_{0B}-h_1$ and $e_{0B}-h_2$ is expected to populate $X_{ud}$ whilst exciting $e_{0AB}-h_1$ and $e_{0AB}-h_2$ generates an indirect exciton $X_{indir}$. We devised an experimental scheme using the direct and indirect character of the excited states to test whether the QDM exhibits a conditional optical response. The inset of Fig.3(a) illustrates schematically the principle of the measurement. $F$ is chosen such that $e_{0B}-h_1$ and $e_{0B}-h_2$ have predominant \textit{direct} excitonic character and $e_{0AB}-h_1$ and $e_{0AB}-h_2$ predominantly \textit{indirect} character. The system is resonantly excited by either one laser or two lasers simultaneously. In the two laser experiment, the first laser termed \emph{pump} is resonant with $e_{0B}-h_1$, as indicated by the red arrow on the inset of Fig.3(a). As discussed above, an exciton created by laser absorption will primarily relax to the $X_{ud}$ ground state before PL is measured via phonon mediated processes \cite{PhysRevB.54.11532}. A second laser, termed \emph{block}, is tuned into resonance with excited states that have a predominantly \textit{indirect} character in order to generate excitons with indirect character $X_{ind}$. If the absorption of the blocking laser is more efficient than that of the pump then the QDM will be driven into an indirect exciton state and the absorption of the pump laser is suppressed since absorption shifts to a biexcitonic state of the system. Typical results of such a measurements are presented in Fig.3(a) that shows the PL spectrum recorded from $X_{ud}$ for $F=21.05$ kV/cm when the pump laser only (0.5 kW/cm$^2$) is applied (red curve). This is compared with the situation when the system is excited only by the blocking laser (5 kW/cm$^2$ - blue curve) and when both lasers are applied simultaneously (black curve).
\begin{figure}[t]
\includegraphics[width=1\columnwidth]{Figure_4}
\caption{\label{fig:Figure_4}
(Color online)
PL of the upper and lower quantum dot as a function of the blocking laser power.
Resonant excitation of an exciton in the upper dot is blocked by the presence of an exciton in the lower dot due to the Coulomb interaction. (b) Intensity of the PL of the upper and lower dot and (c) suppression ratio $\rho$ as a function of the blocking laser power.}
\end{figure}
The intensity of $X_{ud}$ clearly reduces strongly when both lasers are applied simultaneously. We scanned the energy of the blocking laser over the spectrum of excited states from 1322 to 1326 meV. The result of this experiment is shown in Fig.3(b) comparing the intensity of $X_{ud}$ as a function of the blocking laser energy for the pump laser only (red curve), the blocking laser only (blue curve) and with both blocking and pump lasers applied simultaneously (black curve). Three resonances, labeled $R_1$, $R_2$ and $R_3$ in Fig.3(b) can be clearly seen. At these resonances PL is observed from $X_{ud}$ following excitation with the blocking laser only. The PL signal for excitation with both lasers shows a series of dips for an excitation with the pump laser only. $R_1$ and $R_2$ coincide precisely with the PLE resonances $e_{0AB}-h_1$ and $e_{0AB}-h_2$, presented in Fig.2(b) demonstrating that the blocking laser can be used to suppress the absorption of the pump beam.
After establishing the spectrum of interband transitions of the molecule and their electric field dependence we performed an experiment whereby an excitation in one of the two dots forming the molecule was used to block absorption in the other. The scheme for this experiment is illustrated schematically in the inset of Fig.4(a). An exciton with predominantly direct character can be excited in the upper dot via its first excited hole state (red arrow). If, in addition to the pump laser the blocking laser is tuned to a direct exciton transition in the lower dot (blue arrow) the energy of the upper dot absorption shifts to that of the spatially separated biexciton. Thus, the absorption of the pump laser is switched off and, thus, PL from $X_{ud}$ can be optically gated on and off. The result of such a measurement is presented in Fig.4(a) that shows the PL intensity of $X_{ud}$ and $X_{ld}$ for $F=20.8$ kV/cm, a pump laser power density of 1 kW/cm$^{2}$ and different blocking laser power densities from 0 to 1.5 kW/cm$^{2}$.
In the absence of the blocking laser, only PL from $X_{ud}$ can be seen with an intensity of $98\pm4$ cps. However, upon increasing the power of the blocking laser we observe a pronounced decrease in the intensity of $X_{ud}$, while emission from $X_{ld}$ emerges and gradually increases in intensity. These observations clearly demonstrate a conditional optical response, blocking of $X_{ud}$ induced by direct excitation of $X_{ld}$. To quantitatively analyze the effect of the blocking laser on the PL signal from the QDM we define the intensity of $X_{ud}$ subject to the combined pump and blocking lasers as $I_{block, pump}$: there $I_{1,1}$ corresponds to the intensity measured when blocking and pump lasers are switched on, $I_{0,1}$ is the intensity with the pump laser etc. These three intensities are plotted together with the intensity of the lower dot in Fig. 4(b) as a function of the blocking laser power. We measure $I_{0,1} = 98\pm4$ cps (red triangles) and $I_{1,1} = 19\pm1$ cps (green squares) as the intensity of the blocking laser is increased from 0 to 4 kW/cm. When only the blocking laser is applied we measure $I_{1,0}=9\pm1$ cps (blue triangles) from which we can obtain the suppression $\rho = \frac{I_{0,1}-I_{1,1}}{I_{0,1}-I_{1,0}} \sim 88\pm 2\%$. The dependence of $\rho$ on the blocking laser power is presented in Fig.4(c). The reason why the blocking laser induced PL from $X_{ud}$ results from the fact that the resonance used for efficiently exciting $X_{ld}$ is 22.3 meV higher than the ground state of $X_{ld}$. Therefore, relaxation from this excited state to the ground state of $X_{ud}$ is possible due to tunneling of both charge carriers.
In summary, we probed the spectrum of ground and excited state transitions in an individual, electrically tunable artificial molecule. Excited state transitions were identified between hybridized electron states having bonding or antibonding character and different excited hole states. By simultaneously pumping different discrete optical transitions we demonstrated a conditional optical response with an on-off gating fidelity of $88 \pm 2 \%$. The results demonstrate an
electrically tunable, discrete coupled quantum system with a conditional optical response.
We gratefully acknowledge financial support of the DFG via SFB-631 / Nanosystems Initiative Munich and the EU via SOLID.
|
\subsection{#1}\Plots{#1}}
\newcommand{\MySubSubSection}[1]{\subsubsection{#1}\Plots{#1}}
\newcommand{\MyTableReference}[1]{ (see Table \ref{#1}) }
\newcommand{\Caption}[1]{The first level graph presented to the #1 attack profile.}
\newcommand{\Plots}[2]{
\begin{figure*}
\centering
\includegraphics[width=5.0in,angle=-90]{Images/Examples/#1/Rplot001.ps}
\MyCaption{\Caption{#1}}{#2}{fig:#1}
\end{figure*}
}
\newcommand{\Elements}[1]{\cellcolor[gray]{0.5} #1}
\newcommand{\Turns}[1]{\cellcolor[gray]{0.8} #1}
\newcommand{\MyPageSetup}[0]{
\newlength{\myrightmargin}
\newlength{\myleftmargin}
\newlength{\mytopmargin}
\newlength{\mybottommargin}
\setlength{\myrightmargin}{1.0in}
\setlength{\myleftmargin}{1.0in}
\setlength{\mytopmargin}{1.0in}
\setlength{\mybottommargin}{1.0in}
\setlength{\oddsidemargin}{0.0in}
\setlength{\evensidemargin}{0 in}
\setlength{\marginparsep}{0 in}
\setlength{\marginparwidth}{0 in}
\setlength{\hoffset}{\myleftmargin - 1.0in}
\setlength{\textwidth}{8.5in -\myleftmargin -\myrightmargin -\oddsidemargin}
\setlength{\voffset}{\mytopmargin -1.0in}
\setlength{\topmargin}{0 in}
\setlength{\headheight}{12pt}
\setlength{\headsep}{20pt}
\setlength{\footskip}{36pt}
\setlength{\textheight} {11.0in-\mytopmargin-\mybottommargin-\headheight-\headsep-\footskip}
}
\newcommand{\MyFloats}[0] {
\renewcommand{\topfraction}{0.9}
\renewcommand{\bottomfraction}{0.8}
\setcounter{topnumber}{2}
\setcounter{bottomnumber}{2}
\setcounter{totalnumber}{4}
\setcounter{dbltopnumber}{2}
\renewcommand{\dbltopfraction}{0.9}
\renewcommand{\textfraction}{0.01}
\renewcommand{\floatpagefraction}{0.7}
\renewcommand{\dblfloatpagefraction}{0.7}
}
\section{Graph attack profiles}\label{sec:attackProfiles}
\subsection{Comparison of errors and attacks}
Errors and attacks remove components from a system. The distinguishing characteristic between the two
types of losses is how components are selected. This characteristic can be explained by using a computer network as a
graph. The network is a graph where vertices are represented by routers, switches and computers. While edges
are represented by the connections between the vertices, either wired or wireless connections.
The loss of a router through hardware failure, or mis-configuration,
or the severing of the communications links to the router can be considered to be accidental.
An error is the accidental loss of a component from a system.
The simultaneous loss of a set of routers, perhaps without a readily apparent reason, could be considered to
be an attack. An attack is the deliberate loss of components, or a component from a system.
The survivability of a graph to error or attack depends on the underlying structure of the graph (for example scale-free or exponential).
Scale-free graphs are very robust in the face of random failures, but are very susceptible to attacks \cite{Albert2000}. Where
exponential graphs have just the opposite behavior.
\subsection{Selection of graph component to attack}
Ultimately there are only two graph components that an attacker can attack, edges or vertices. The selection of which
of these components to attack has to be based on some metric rather than random selection. Holme and Kim \cite{holme2002attack}
looked at how an attacker could maximize the damage to a graph by one of two approaches. The approaches being:
\begin{enumerate}
\item To remove the vertex with the highest initial degree (ID)
\MyEquation{degreeUndirectedDegree}
\item Or, the vertex with the highest in-betweenness
centrality (IB) \MyEquation{betweennessVertex}
\end{enumerate}
Their idea about betweenness can be extended to include removing the edge with the highest in-betweenness centrality
\MyEquation{betweennessEdge}.
Lee et al. in \cite{lee2006rnt} put forth failures in a network as being either \textbf{node}, \textbf{link}, or \textbf{path} related.
Their \textbf{node} corresponds to our \textbf{vertex}. Their \textbf{link} to our \textbf{edge}. And, their \textbf{path} to our
\textbf{betweenness}. The betweenness of a component is a measurement of the component's contribution to all the shortest paths
\MyEquationInline{pathLHS} in the graph. The higher the betweenness value, the more shortest paths use that component.
In the following subsections, we will use a sample graph to show the effects of an attacker's limited knowledge of the global graph
on which component to remove.
\subsubsection{Size of subgraph to evaluate}
An attacker has to select a graph component to attack, and identifying which component to remove is
based on the attacker's knowledge of some portion of the graph. The attacker's knowledge can
range from a single component to complete knowledge of the graph. One approach to gaining knowledge of a graph's organization
is to identify a vertex and then determine those vertices that are at a path length distance of 1 edge from
the initial vertex. This process is repeated again and again until the attacker decides to stop increasing the
path length \MyFigureReference{fig:neighbor}.
In Figure \ref{fig:neighbor}, vertex 5 is the source vertex and is colored red. The path length is initially
set to 1 and the attacker now knows about the vertex set \{4, 5, 6, 8, 9\} \MyFigureReference{fig:neighbor:1}. All
attacker discovered vertices are colored pink. As the
path length increases from 2 \MyFigureReference{fig:neighbor:2} to 4 \MyFigureReference{fig:neighbor:4}, more and more
of the global graph becomes known. As readers, we know what the global graph looks like because we have an omnipotent view point.
The attacker does not enjoy this view and must blindly continue to work outwards from his initial vertex.
The attacker must expend time and energy to increase his knowledge of the graph, until at some point
he will have spent ``enough'' and believes that sending additional time will not be worth the effort.
The attacker uses this limited local knowledge of the global graph to select the
component whose removal will cause the greatest damage to the graph. If the path length is increased
enough, the entire graph will be discovered. Barab{\'a}si hypothesized that the entire INTERNET
could be discovered with a path length of 19 \cite{barabsi:world_wide_web_forms_a_large_directed_graph}. The
resources for attempting to conduct such a discovery may be too large to be practical.
\begin{figure}
\centering
\MyLocalSubfigure{Path length = 1, discovered diameter = 2 }{fig:neighbor:1}{Images/neighborSize-001.eps}
\MyLocalSubfigure{Path length = 2, discovered diameter = 4}{fig:neighbor:2}{Images/neighborSize-002.eps}
\MyLocalSubfigure{Path length = 3, discovered diameter = 6}{fig:neighbor:3}{Images/neighborSize-003.eps}
\MyLocalSubfigure{Path length = 4, discovered diameter = 7}{fig:neighbor:4}{Images/neighborSize-004.eps}
\MyCaption{The effects of different path lengths starting from a fixed vertex in discovering the global graph.}
{Vertex 5 is the center vertex. Each sub-figure shows the subgraph that is discovered based on the path length
from the center vertex as the path length increments from 1 to 4. The diameter of the discovered subgraph is at most twice the
path length. As the path length increases, more and more of the global graph is discovered.}
{fig:neighbor}
\end{figure}
\subsubsection{Edge selection}\label{sec:edgeSelection}
The selection of an edge to remove from the graph is based on how much of the graph that the attacker has
discovered. As the discovered graph becomes larger and larger (as measured by the path length from a initial/central)
vertex to the rest of the graph \MyFigureReference{fig:neighbor}, the more accurate the computed value betweenness value
of the edge is to the edge's betweenness value for the entire graph. The edge betweenness value for all edges in the
global graph and for the discovered subgraph is shown in Table \ref{tbl:neighborhood:edge}.
In the table, the first two columns are the vertices that are connected by an edge. The third column is the edge betweenness
for that edge based on the global graph. The remaining columns show the edge betweenness value as the path length from
the central vertex gets longer and longer. In those cases where the discovered subgraph has not discovered a particular
vertex in the global graph, the edge betweenness value is marked with a --- indicating no value possible. It is interesting to
see how the value of an edge changes as the size of the graph changes. In most cases the value of an edge decreases as graph size
increases.
\begin{table}
\input{edgeTable}
\MyCaption{Comparing the betweenness of edges based on the neighborhood discovered from a central vertex.}
{The size of the neighborhood increases from 1 to 4 based around vertex 5 \MyFigureReference{fig:neighbor}.
As the size of the neighborhood gets closer and closer to the global graph, the betweenness values
get closer and closer to the global values. Those edges that have not been discovered because they belong
to a portion of the global graph that has not been discovered are marked with a ---.}
{tbl:neighborhood:edge}
\end{table}
\subsubsection{Vertex selection}
The selection of a vertex to remove from the graph is based on how much of the graph that the attacker has
discovered. As the discovered graph becomes larger and larger (as measured by the path length from a initial/central)
vertex to the rest of the graph \MyFigureReference{fig:neighbor}, the more accurate the computed value betweenness value
of the vertex is to the vertex's betweenness value for the entire graph. The betweenness value for all vertices in the
global graph and for the discovered subgraph is shown in Table \ref{tbl:neighborhood:vertex}.
In the table, the first column is the vertex number. The second column is the
vertex's betweenness value based on the global graph. The remaining columns show the vertex betweenness value as the path length from
the central vertex gets longer and longer. In those cases where the discovered subgraph has not discovered a particular
vertex in the global graph, the vertex betweenness value is marked with a --- indicating no value possible. It is interesting to
see how the value of an vertex changes as the size of the graph changes. In most cases the value of an vertex decreases as graph size
increases. One notable exception is the vertex 2. As the graph size increases, that vertex's
betweenness increase and decreases and yet in the global graph, its value is less than in some of the subgraphs.
\begin{table}
\input{vertexTable}
\MyCaption{Comparing the betweenness of vertices based on the neighborhood discovered from a central vertex.}
{The size of the neighborhood increases from 1 to 4 based around vertex 5 \MyFigureReference{fig:neighbor}.
As the size of the neighborhood get closer and closer to the global graph, the betweenness values
get closer and closer to the global values. Those vertices that have not been discovered because they belong
to a portion of the global graph that has not been discovered are marked with a ---. The betweenness values have been
normalized to the range (0,1) to allow comparisons across different sized graphs.}
{tbl:neighborhood:vertex}
\end{table}
\subsubsection{Degree selection}
Discovering the degree of a node is based on the idea that the nodes exchange messages between themselves and
that the attacker can intercept these messages. As the attacker intercepts more and more messages; a node's neighbors
(a.k.a., degree) can be determined. The degree of a node can be used as a criterion to determine if the node is worthy of
attack.
The degrees for the discovered graph based on differing path lengths is shown in Table \ref{tbl:neighborhood:degree}. The first column
is the vertex number. The second column is the vertex's global degree. The remaining columns show the degree of the each
of the discovered vertices as the path length increases. If the vertex has not been discovered based on a particular path length
then the marker --- is used to indicate that no data is available. It is interesting to note that the degree of a vertex
always increases as the path length increases until the global degree value is reached. Once the global value is reached, it remains constant.
\begin{table}
\input{degreeTable}
\MyCaption{Comparing the degreeness of each vertex based on the neighborhood discovered from a central vertex.}
{The size of the neighborhood increases from 1 to 4 based around vertex 5 \MyFigureReference{fig:neighbor}.
As the size of the neighborhood get closer and closer to the global graph, the betweenness values
get closer and closer to the global values.}
{tbl:neighborhood:degree}
\end{table}
\subsection{Attack Profile Notation}
An attacker can target any graph component for removal based on the damage estimate or other criteria and
whether to use the highest, or lowest valued component based on those criteria. We introduce
the notation \MyAttackNotation{C}{V} as a short hand way to identify a specific profile.
The first subscript in \MyAttackNotation{C}{V} is the metric that is being used to select
a component $C\in\{E,V,D,*\}$ for \textit{edge, vertex, degree or any} respectively.
The second subscript is the value of the metric that is being used
$V\in\{L,M,H,R,*\}$ for \textit{low, medium, high, random or any} respectively.
The notation
\MyAttackNotation{D}{H} means that the attacker is using a profile that targets nodes based on their
degree $D$ and choose the highest $H$ valued one.
\subsection{Effectiveness of different attack profiles}
The damage to a graph by fragmentation can be calculated \MyEquationReference{equ:damage} using the fragmented graph and
approximating the graph without fragmentation.
\MyEquationLabeled{damage-02.tex}{equ:damage}
An unfragmented graph is created from the fragmented graph by adding an edge between
each of the highest degreed nodes of each fragment. As each edge is added to coalesce the fragments into a larger and larger connected
component, the highest degreed node may change based on the order in which the fragments are coalesced. Therefore the highest
degreed node in the coalescing component must be evaluated after each fragment addition. At the end of the collation process,
there will be a single connected component containing the same number of nodes as the fragmented graph and one additional edge for
each of the original fragments.
As the original graph becomes more and more fragmented, its AIPL will decrease. The AIPL of the unfragmented approximation will decrease and
the \MyEquationInline{damageLHS} will increase as well. This behavior is readily apparent when edges are removed from the
original graph in order to create the fragments. When vertices are removed, the behavior is similar, until the last vertex is removed.
In the limiting case,
AIPL of the fragmented graph with one fragment and one node in that fragment, is the same as the AIPL of a connected component with one node.
Using Equation \ref{equ:damage} results in a value of 0 meaning that the graph is undamaged.
\subsubsection{Edge selection}
The attacker can compute the betweenness of any edge in the subgraph that he has discovered \MyTableReference{tbl:neighborhood:edge}. Based on these
computed betweenness values, the attacker can select either the highest or lowest valued edge to remove. After the removal
of this edge, the betweenness values can be recomputed for the newly modified subgraph and the process repeated again and again
until there are no edges left in the discovered graph (the discovered graph is totally destroyed).
Figures \ref{fig:delete:edge:low} and \ref{fig:delete:edge:high} show the effects of repeatedly applying attack \MyAttackNotation{E}{L} or
\MyAttackNotation{V}{L} profile to the discovered subgraph of path length 3. In each figure, the betweenness value of each edge is written on the edge. The
edge with the lowest \MyFigureReference{fig:delete:edge:low} or highest \MyFigureReference{fig:delete:edge:high} betweenness value
is highlighted in red, prior to it being removed. After the removal of the edge, the betweenness values of all the
remaining edges is computed shown in the next subfigure, along with the next edge that has been selected for removal. The four subfigures in Figures \ref{fig:delete:edge:low} and \ref{fig:delete:edge:high} show this process. When two or more edges have the same betweenness value, the
selection of which edge to remove it totally random.
Attack profile \MyAttackNotation{E}{L} tends to attack the periphery of the graph. While profile \MyAttackNotation{E}{H} tends
to attack the core of the graph. Either profile will result in a fully disconnected graph with the same number of
removals, selecting the highest valued edge causes more damage quicker.
\begin{figure}
\centering
\MyLocalSubfigure{First lowest has been identified}{fig:delete:edge:low:0}{Images/deleteEdge-L-003-0}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:delete:edge:low:1}{Images/deleteEdge-L-003-1}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:delete:edge:low:2}{Images/deleteEdge-L-003-2}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:delete:edge:low:3}{Images/deleteEdge-L-003-3}
\MyCaption{The effects of the \MyAttackNotation{E}{L} attack profile on the sample graph.}
{Vertex 5 is the center vertex and is marked in red. The discovered graph is at a path length of 3 from the center vertex
and is marked in pink.
The edge with the lowest
betweenness value is marked in red. After each deletion, all edge betweenness values are recomputed because the
graph has changed. Some of the edges are unlabeled because the attacker has not ``discovered'' them.}
{fig:delete:edge:low}
\end{figure}
\begin{figure}
\centering
\MyLocalSubfigure{First highest has been identified}{fig:delete:edge:high:0}{Images/deleteEdge-H-003-0}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:edge:high:1}{Images/deleteEdge-H-003-1}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:edge:high:2}{Images/deleteEdge-H-003-2}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:edge:high:3}{Images/deleteEdge-H-003-3}
\MyCaption{The effects of the \MyAttackNotation{E}{H} attack profile on the sample graph.}
{Vertex 5 is the center vertex and is marked in red. The discovered graph is at a path length of 3 from the center vertex
and is marked in pink. The edge with the highest
betweenness value is marked in red. After each deletion, all edge betweenness values are recomputed because the
graph has changed. Some of the edges are unlabeled because the attacker has not ``discovered'' them.}
{fig:delete:edge:high}
\end{figure}
Table \ref{tbl:damage:edge} lists the computed damage to the discovered subgraph after the removal of either the highest or lowest
betweenness valued edge. Figure \ref{fig:damage:edge} shows the damage plotted against the deletion.
There are 16 edges in the discovered subgraph and damage is total upon the removal of the last edge.
\begin{table}
\input{damageEdge}
\MyCaption{Damage to the discovered subgraph of path length 3 based on \MyAttackNotation{E}{*} attack profiles.}
{The betweenness of each edge is recomputed after the
removal of either the highest or lowest betweenness valued edge. The process is repeated again and again until all edges
are removed.}
{tbl:damage:edge}
\end{table}
\begin{figure}
\centering
\includegraphics[width = 4.5in, angle = -90]{Images/damageEdge.eps}
\MyCaption{Damage to the discovered graph of path length 3 based on \MyAttackNotation{E}{*} attack profiles.}{
The ``local'' values are those that come from the discovered graph, while the
global values are from the total graph. Damage inflicted on the discovered graph when
using the high edge betweenness value and the resulting impact on the total
graph are show in black and red respectively. In a similar manner, damage caused by choosing the low
betweenness is shown in the green and blue lines respectively. The betweenness of each edge is recomputed after the
removal of either the highest or lowest betweenness valued edge. The process is repeated again and again until all edges
are removed.}
{fig:damage:edge}
\end{figure}
\subsubsection{Vertex selection}
The attacker can compute the betweenness of any vertex in the subgraph that he has discovered \MyTableReference{tbl:neighborhood:vertex}. Based on these
computed betweenness values, the attacker can select either the highest or lowest valued vertex to remove. After the removal
of this vertex, the betweenness values can be recomputed for the newly modified subgraph and the process repeated again and again
until there are no vertices left in the discovered graph (the discovered graph is totally destroyed).
Figures \ref{fig:delete:vertex:low} and \ref{fig:delete:vertex:high} show the effects of repeatedly applying
\MyAttackNotation{V}{L} or \MyAttackNotation{V}{L} profile to the
discovered subgraph of path length 3. In each figure, the betweenness value of each vertex is written in the vertex. The
vertex with the lowest \MyFigureReference{fig:delete:vertex:low} or highest \MyFigureReference{fig:delete:vertex:high} betweenness value
is highlighted in yellow, prior to it being removed. After the removal of the vertex, the betweenness values of all the
remaining vertices are computed and shown in the next subfigure, along with the next vertex that has been selected for removal.
The four subfigures in Figures \ref{fig:delete:vertex:low} and \ref{fig:delete:vertex:high} show this process. When two or more vertices have the same betweenness value, the
selection of which edge to remove it totally random.
Attack profile \MyAttackNotation{V}{L} tends to attack the periphery of the subgraph. While attack
profile \MyAttackNotation{V}{H} tends
to attack the core of the graph. While both selection choices will result in a fully disconnected graph with the same number of
removals, selecting the highest valued vertex causes more damage quicker.
The betweenness computation, removal and damage computation process is shown in Table \ref{tbl:damage:vertex} and Figure \ref{fig:damage:vertex}.
The global high line in Figure \ref{fig:damage:vertex} goes flat after the fifth deletion
while the global low line continues to increase. This behavior is explained by looking
at Figures \ref{fig:delete:vertex:high:4} and \ref{fig:delete:vertex:low:4}. By the fifth
high deletion, the discovered and global graphs are disconnected and further local
deletions do not affect the global graph. In Figure \ref{fig:delete:vertex:high:4},
the discovered and global graphs are still connected and local deletions will affect the
global graph.
\begin{figure}
\centering
\MyLocalSubfigure{First lowest has been identified}{fig:vertex:edge:low:0}{Images/deleteVertex-L-003-0}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:vertex:edge:low:1}{Images/deleteVertex-L-003-1}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:vertex:edge:low:2}{Images/deleteVertex-L-003-2}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:vertex:edge:low:3}{Images/deleteVertex-L-003-3}
\MyCaption{The effects of an \MyAttackNotation{V}{L} attack profile on the sample graph.}
{Vertex 5 is the center vertex and is shown in red. The discovered graph, in pink is at a distance of 3 from the center vertex.
Each vertex is labeled with the number of shortest paths that go use that vertex.
The vertex with the lowest betweenness is drawn in yellow. Each time, the lowest valued vertex is removed from the
discovered graph and all betweenness values for the discovered graph are recomputed. If there is more
than one vertex with the same low value, one is selected at random for removal.
Some of the vertices are unlabeled because the attacker has not ``discovered'' them.}
{fig:delete:vertex:low}
\end{figure}
\begin{figure}
\centering
\MyLocalSubfigure{First highest has been identified}{fig:delete:vertex:high:0}{Images/deleteVertex-H-003-0}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:vertex:high:1}{Images/deleteVertex-H-003-1}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:vertex:high:2}{Images/deleteVertex-H-003-2}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:vertex:high:3}{Images/deleteVertex-H-003-3}
\MyCaption{The effects of an \MyAttackNotation{V}{H} attack profile on the sample graph.}
{Vertex 5 is the center vertex. The discovered graph is at a distance of 3 from the center vertex.
The vertex with the highest betweenness is drawn in yellow. Each time, the highest valued vertex is removed from the
discovered graph and all betweenness values for the discovered graph are recomputed. If there is more
than one vertex with the same high value, one is selected at random for removal.
Some of the vertices are unlabeled because the attacker has not ``discovered'' them.}
{fig:delete:vertex:high}
\end{figure}
\begin{table}
\input{damageVertex}
\MyCaption{Damage to the discovered subgraph of path length 3 based on \MyAttackNotation{V}{*} attack profiles.}
{The betweenness of each vertex is recomputed after the
removal of either the highest or lowest betweenness valued vertex. The process is repeated again and again until all vertices
are removed.}
{tbl:damage:vertex}
\end{table}
\begin{figure}
\centering
\includegraphics[width = 4.5in, angle = -90]{Images/damageVertex.eps}
\MyCaption{Damage to the discovered subgraph of path length 3 based on \MyAttackNotation{V}{*} attack profiles.}{
The ``local'' values are those that come from the discovered graph, while the
global values are from the total graph. Damage inflicted on the discovered graph when
using the high vertex betweenness value and the resulting impact on the total
graph are show in black and red respectively. In a similar manner, damage caused by choosing the low
betweenness is shown in the green and blue lines respectively.
The betweenness of each vertex
is recomputed after the
removal of either the highest or lowest betweenness valued vertex. The process is repeated again and again until all vertices
are removed. Damage to the global graph is flat from deletion 4 through 11, while the local damage increases due to the selection
of the particular high valued vertices to remove. The low betweenness option does not show this type of behavior.
The system of graphs for high and low selection is shown in Figure \ref{fig:damage:vertex:explain}. }
{fig:damage:vertex}
\end{figure}
\begin{figure}
\centering
\MyLocalSubfigure{Results of \MyAttackNotation{V}{H}}{fig:delete:vertex:high:4}{Images/deleteVertex-H-003-4}
\MyLocalSubfigure{Results of \MyAttackNotation{V}{L}}{fig:delete:vertex:low:4}{Images/deleteVertex-L-003-4}
\MyCaption{Markedly different graphs resulting from the differences in choosing \MyAttackNotation{V}{H} or
\MyAttackNotation{V}{L} attack profiles.}
{Both subfigures show the sample graph after 4 deletions based on \MyAttackNotation{V}{H} or \MyAttackNotation{V}{L}
attack profiles.
Continued deletions in the discovered graph (in pink) in the high betweenness case \MyFigureReference{fig:delete:vertex:high:4},
will have only marginal effect on the global graph (the union of pink and green). Deletions
in the discovered graph in low betweenness case \MyFigureReference{fig:delete:vertex:low:4} will continue
to affect the union of the pink and the green nodes because the two graphs (pink and green) are still connected.
Some of the vertices are unlabeled because the attacker has not ``discovered'' them.}
{fig:damage:vertex:explain}
\end{figure}
\subsubsection{Degree selection}
The attacker can compute the degreeness of any vertex in the subgraph that he has discovered \MyTableReference{tbl:neighborhood:degree}. Based on these
values, the attacker can select either the highest or lowest valued vertex to remove. After the removal
of this vertex, the degreeness values can be recomputed for the newly modified subgraph and the process repeated again and again
until there are no vertices left in the discovered graph (the discovered graph is totally destroyed).
Figures \ref{fig:delete:degree:low} and \ref{fig:delete:degree:high} show the effects of repeatedly applying attack \MyAttackNotation{D}{L} or
\MyAttackNotation{D}{L} profiles to the
discovered subgraph of path length 3. In each figure, the degreeness value of each vertex is written in the vertex. The
edge with the lowest \MyFigureReference{fig:delete:degree:low} or highest \MyFigureReference{fig:delete:degree:high} betweenness value
is highlighted in yellow, prior to it being removed. After the removal of the vertex, the degreeness values of all the
remaining vertices are computed shown in the next subfigure, along with the next vertex that has been selected for removal. The four subfigures in Figures \ref{fig:delete:degree:low} and \ref{fig:delete:degree:high} show this process. When two or more vertices have the same degreeness value, the
selection of which edge to remove it totally random.
Attack profile \MyAttackNotation{D}{L} tends to attack the periphery of the subgraph. While attack
profile \MyAttackNotation{D}{H} tends to attack the core of the graph.
While both selection choices will result in a fully disconnected graph with the same number of
removals, selecting the highest valued vertex causes more damage quicker.
The betweenness computation, removal and damage computation process is shown in Table \ref{tbl:damage:degree} and Figure \ref{fig:damage:degree}.
The Global High line in Figure \ref{fig:damage:degree} goes flat after the fifth deletion while the Global Low line continues to
increase. This behavior is explained by looking at Figures \ref{fig:delete:degree:high:5} and \ref{fig:delete:degree:low:5}. Using a
\MyAttackNotation{D}{H} profile, the discovered and global graphs are disconnected and further local deletions do not affect the
global graph. Using \MyAttackNotation{D}{L} profile in Figure \ref{fig:delete:degree:low:5} results in the
discovered and global graphs still being connected, so any deletions on the discovered graph affect the global graph.
the fifth deletion the discovered and global graphs are
\begin{figure}
\centering
\MyLocalSubfigure{First lowest has been identified}{fig:delete:degree:low:0}{Images/deleteDegree-L-003-0}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:delete:degree:low:1}{Images/deleteDegree-L-003-1}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:delete:degree:low:2}{Images/deleteDegree-L-003-2}
\MyLocalSubfigure{Previous lowest has been removed, new lowest identified}{fig:delete:degree:low:3}{Images/deleteDegree-L-003-3}
\MyCaption{The effects of an \MyAttackNotation{D}{L} attack profile on the sample graph.}
{Vertex 5 (marked in red) is the center vertex. The discovered graph is at a distance of 3 from the center vertex. The vertex with the
lowest degree is marked in yellow. In the case where multiple vertices have the same degree value \MyFigureReference{fig:delete:degree:low:1},
random choice is used to select one vertex as the next one to be removed.
Removal of a vertex causes a reduction in the degree values of all of the removed vertex's neighbors. This change in the
degreeness of potentially many vertices requires that the relative order of the vertices be evaluated after each removal.
Some of the vertices are unlabeled because the attacker has not ``discovered'' them.}
{fig:delete:degree:low}
\end{figure}
\begin{figure}
\centering
\MyLocalSubfigure{First highest has been identified}{fig:delete:degree:high:0}{Images/deleteDegree-H-003-0}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:degree:high:1}{Images/deleteDegree-H-003-1}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:degree:high:2}{Images/deleteDegree-H-003-2}
\MyLocalSubfigure{Previous highest has been removed, new highest identified}{fig:delete:degree:high:3}{Images/deleteDegree-H-003-3}
\MyCaption{The effects of on \MyAttackNotation{D}{H} attack profile on the sample graph.}
{Vertex 5 (marked in red) is the center vertex. The discovered graph is at a distance of 3 from the center vertex. The vertex with the
highest degree is marked in yellow. In the case where multiple vertices have the same degree value \MyFigureReference{fig:delete:degree:high:2},
random choice is used to select one vertex as the next one to be removed.
Removal of a vertex causes a reduction in the degree values of all of the removed vertex's neighbors. This change in the
degreeness of potentially many vertices requires that the relative order of the vertices be evaluated after each removal.
Some of the vertices are unlabeled because the attacker has not ``discovered'' them.}
{fig:delete:degree:high}
\end{figure}
\begin{table}
\input{damageDegree}
\MyCaption{Damage to the discovered subgraph of path length 3 based on \MyAttackNotation{D}{*} attack profiles.}
{The degree of each vertex is computed after each deletion. A vertex's degree value will change if one of it's
immediate neighbor vertices has been removed. The removal of a neighbor will reduce the degreeness of all its
neighbors by one. This change in the degreeness of all neighboring vertices may affect the relative order of all vertices
based on their respective degreeness.
The process is repeated again and again until all edges
are removed.}
{tbl:damage:degree}
\end{table}
\begin{figure}
\centering
\includegraphics[width = 4.5in, angle = -90]{Images/damageDegree.eps}
\MyCaption{Damage to the discovered subgraph of path length 3 by based on \MyAttackNotation{D}{*} attack profiles.}
{The degree of each vertex is computed after each deletion. A vertex's degree value will change if one of it's
immediate neighbor vertices have been removed. The removal of a neighbor will reduce the degreeness of all its
neighbors by one. This change in the degreeness of all neighboring vertices may affect the relative order of all vertices
based on their respective degreeness.
The process is repeated again and again until all vertices
are removed. The flat area on the Global High line
is related to the discovered and global graphs becoming
disconnected \MyFigureReference{fig:damage:degree:explain}.}
{fig:damage:degree}
\end{figure}
\begin{figure}
\centering
\MyLocalSubfigure{High degree}{fig:delete:degree:high:5}{Images/deleteDegree-H-003-5}
\MyLocalSubfigure{Low degree}{fig:delete:degree:low:5}{Images/deleteDegree-L-003-5}
\MyCaption{The sample graph after removing the fifth discovered node using
\MyAttackNotation{D}{*} attack profiles.}
{The undiscovered graph is drawn in green. The central vertex, where it remains
is drawn in red \MyFigureReference{fig:delete:degree:low:5}. The vertex that will be deleted next
is drawn in yellow. While each graph shows the effects of five deletions, selecting the
highest degreed node to delete results in a graph that is disconnected \MyFigureReference{fig:delete:degree:high:5}.
Focusing on the lowest degreed node results in damage to the periphery and a graph that is
still connected \MyFigureReference{fig:delete:degree:low:5}. Some of the vertices are unlabeled because the attacker has not ``discovered'' them.}
{fig:damage:degree:explain}
\end{figure}
\subsection{Attack profile conclusions}
All node based attacks (\MyAttackNotation{V}{*}, \MyAttackNotation{D}{*}) will
totally destroy the discovered graph. All edge based attacks \MyAttackNotation{E}{*}
will cause the discovered graph to be totally disconnected. The two attack
philosophies differ in their efficacy and are summarized in Table \ref{tbl:attack:efficacy}.
If the attacker's goal is to disconnect the sample graph by repeated use of the same
attack profile, then the most effective profiles in order are: \MyAttackNotation{E}{H},
\MyAttackNotation{V}{H} and \MyAttackNotation{D}{H}.
\begin{table}
\centering
\setlength{\extrarowheight}{3pt}
\begin{tabular}{|c|p{2.75in}|}
\MyHline
\multicolumn{1}{|p{0.75in}}{\textbf{Attack Profile}} &
\multicolumn{1}{|c|}{\textbf{Efficacy}} \\
\hline
\MyAttackNotation{D}{H} & Tends to attack the core of the graph\\
\MyAttackNotation{D}{L} & Tends to attack the periphery of the graph\\
\MyAttackNotation{E}{H} & Tends to attack the core of the graph\\
\MyAttackNotation{E}{L} & Tends to attack the periphery of the graph\\
\MyAttackNotation{V}{H} & Tends to attack the core of the graph\\
\MyAttackNotation{V}{L} & Tends to attack the periphery of the graph\\
\MyHline
\end{tabular}
\MyCaption{Efficacy of various attack profiles.}
{In general, regardless of the attack profile utilized,
attacking the highest valued component is the most destructive.}
{tbl:attack:efficacy}
\end{table}
\section{Comparison of connected and disconnected graph metrics}\label{sec:metrics}
Within this paper the following terms and ideas are used:
\begin{enumerate}
\item A graph \MyEquationInline{graph} is an ordered pair of disjoint sets
(\MyEquationInline{vertexLHS},\MyEquationInline{edgeLHS}) such that \MyEquationInline{edgeLHS} is a subset of $\MyEquationInline{vertexLHS}^{2}$ of the unordered pairs of
\MyEquationInline{vertexLHS} \cite{bollobas1998modern}.
\item The terms \emph{vertex} and \emph{node} are used interchangeably and mean the
same thing.
\item The term \emph{connected} means that there is a series \emph{edges} between
any arbitrary nodes source \emph{s} and terminus \emph{t} that can be used to get from
node \emph{s} to \emph{t}\cite{bollobas1998modern}.
A graph is \emph{disconnected} when nodes \emph{s} and \emph{t}
cannot be reached by any series of edges.
\item The term \emph{directed} means that the edge connecting nodes \emph{s} and \emph{t}
is unidirectional. \emph{t} is an immediate neighbor to \emph{s} because they are
separated by one edge and it takes more than one edge for \emph{t} to reach \emph{s}.
\item The term \emph{undirected} means that the edge connecting nodes \emph{s} and \emph{t}
is bidirectional. \emph{t} is an immediate neighbor to \emph{s} because they are
separated by one edge and the same edge connects \emph{t} to \emph{s}.
\item The term \emph{simple} means that there is only one \emph{edge} between
any adjacent nodes.
\item The terms \emph{fragment}, \emph{cluster} or \emph{component} are used
interchangeably and mean a set of \emph{nodes} (there may be only 1 node) that
are connected to each other. A graph \emph{G} may have more than one \emph{component}.
\item The difference between a \emph{graph} and a \emph{network} is the assignment
of different \emph{weights} to each \emph{edge} in the graph. By default, all
\emph{edges} in a \emph{graph} have a weight of 1. While, \emph{edges} in a \emph{network}
may have different weights.
\item A node could have an edge that started and ended at the same source node. These edges are
called \emph{self loops}.
\end{enumerate}
The graphs in this paper are: \emph{undirected}, \emph{simple}, \emph{self loops} are not permitted
and may have more than one \emph{component}.
\subsection{Connected graph metrics} \label{sec:connected}
Here we review a collection of characteristic metrics for connected graphs.
In many cases the characteristic does not have meaning, or a computable value when the
graph is not connected.
\begin{description}
\item Path length \cite{bollobas1998modern}. \\
The number of edges in a path $P$ from a starting node $u$ to terminating node $v$.
\MyEquationLabeled{distanceEdge}{equ:00c}
\input{apl}
\MyIgnore{\input{aipl}}
\item Centrality, betweenness of an edge \cite{koschutzkicentrality}. \\
The proportion of shortest paths between nodes $s$ and $t$ that use edge $e$.
\MyEquationLabeled{betweennessEdge}{equ:03c}
\item Centrality, betweenness of an edge relative to all edges in a graph.\\
The edge that has the highest centrality of all edges is the edge that is
most used by all shortest paths in the graph.
\MyEquationLabeled{betweennessGlobalEdge}{equ:14c}
\item Centrality, betweenness of a vertex \cite{koschutzkicentrality}. \\
The proportion of shortest paths between nodes $s$ and $t$ that use vertex $v$.
\MyEquationLabeled{betweennessVertex}{equ:04c}
\item Centrality, betweenness of a vertex relative to all vertices in a graph.\\
The vertex that has the highest centrality of all vertices is the vertex that is used
by the most shortest paths in the graph.
\MyEquationLabeled{betweennessGlobalVertex}{equ:15c}
\MyIgnore{
\item Centrality, closeness of a vertex \cite{csardi2006igraph}. \\
How close (fewest
number of edges) $u$ is to all other vertices.
\MyEquationLabeled{closenessNormalized}{equ:05c}
\item Centrality, degree. \\
The number of edges incident to a vertex. \MyEquationLabeled{degreeUndirectedDegree}{equ:06c}
}
\item Clustering coefficient \cite{brinkmeier211,watts:collective_dynamics}.\\
The likelihood that two neighbors of $v$ are connected.
\MyEquationLabeled{clusteringCoefficient}{equ:07c}
\item Degree of a node.\\
The number of edges incident to a node.
\MyEquationLabeled{degree}{equ:12c}
\item Diameter of a graph \cite{brinkmeier211}.\\
The maximal shortest path between any vertices $u$ and $v$.
\MyEquationLabeled{diameter}{equ:08c}
\item Eccentricity of a node \cite{brinkmeier211,west2001introduction}.\\
The maximal distance between vertex $u$ and any other vertex $v$.
\MyEquationLabeled{eccentricity}{equ:09c}
\item Eccentricity of a graph. \\
The maximal eccentricity of all nodes $u$ in $G$.
\MyEquationLabeled{eccentricityGlobal}{equ:11c}
\item Radius of a graph \cite{brinkmeier211,west2001introduction}.\\
The minimal eccentricity of all vertices in $G$.
\MyEquationLabeled{radius}{equ:10c}
\item Triangles based on a node \cite{brinkmeier211}.\\
The number of subgraphs of the graph $G$ that have exactly three nodes and three edges and one
of the nodes is $v$.
\MyEquationLabeled{triangleLocal}{equ:13c}
\end{description}
Equations \ref{equ:01c}, \ref{equ:08c} and \ref{equ:09c} are directly related to the
length of the path between nodes $u$ and $v$ \MyEquationReference{equ:00c}. Equations \ref{equ:03c},
\ref{equ:14c}, \ref{equ:04c}, \ref{equ:15c}, \ref{equ:07c}, \ref{equ:09c} and \ref{equ:10c} are
indirected related to the path length.
\subsection{Disconnected graph metrics}\label{sec:disconnected}
Here we review a collection of characteristic metrics for disconnected graphs.
In many cases the connected graph characteristic does not have meaning, or is not computable
when the graph is disconnected.
\begin{description}
\input{aplConstrained}
\input{aipl}
\end{description}
Equation \ref{equ:01d} is an constrained APL as
compared to a un-constrained APL \MyEquationReference{equ:01c} that restricts the
path lengths between nodes to those whose path length is not $\infty$. Equation \ref{equ:02c} at first appears to be
dependent on a path length, but in fact, it does not. If a path does not exist between nodes $u$ and $v$
then, by definition, the path length is infinite $\infty$. Any number divided by $\infty$ is defined
to be 0.
\subsection{The effect of directivity and self loops}
\input{edges}
\section{Derivation of various Albert, Jeong and Barab{\'a}si related estimations}\label{sec:derivation}
Often papers have only the solution to a problem or perhaps only the first and last steps. What follows is
a collection of all the equations and their derivations for the solutions in Table \ref{tbl:s2}.
Table \ref{tbl:s2} is repeated here for convenience. In most cases this is basic algebra and the equations are
here because sometimes it is hard to remember how an answer was derived when only an answer is given.
\begin{table}[ht]
\input {tableS2}
\end{table}
\MyExplanation{\frac{n}{2}}{1}{equ:3-3}
\begin{eqnarray}
\frac{n-1}{\frac{n}{2}-1} & \approx & \frac{n}{\frac{n}{2}} \nonumber \\
& \approx & 2 \label{equ:3-3}
\end{eqnarray}
\MyExplanation{\frac{n}{2}}{\frac{n}{2}}{equ:3-4}
\begin{eqnarray}
\frac{n-\frac{n}{2}}{\frac{n}{2}-1} &\approx &\frac{\frac{n}{2}}{\frac{n}{2}} \nonumber \\
& \approx & 1 \label{equ:3-4}
\end{eqnarray}
\MyExplanation{\frac{n}{2}}{\frac{n}{j}}{equ:3-5}
\begin{eqnarray}
\frac{n-\frac{n}{j}}{\frac{n}{2}-1} &\approx &\frac{n(1-\frac{1}{j})2}{n} \nonumber \\
& \approx& 2-\frac{2}{j} \label{equ:3-5}
\end{eqnarray}
\MyExplanation{\frac{n}{2}}{n-1}{equ:3-6}
\begin{eqnarray}
\frac{n-(n-1)}{\frac{n}{2}-1} & \approx & \frac {1}{\frac{n}{2}} \nonumber \\
& \approx & \frac {2}{n} \label{equ:3-6}
\end{eqnarray}
\MyExplanation{\frac{n}{j}}{1}{equ:4-3}
\begin{eqnarray}
\frac{n-1}{\frac{n}{j}-1} & \approx & \frac {n}{\frac{n}{j}} \nonumber \\
& \approx & j \label{equ:4-3}
\end{eqnarray}
\MyExplanation{\frac{n}{j}}{\frac{n}{2}}{equ:4-4}
\begin{eqnarray}
\frac{n - \frac{n}{2}}{\frac{n}{j}-1} &\approx &\frac{n(1-\frac{1}{2})j}{n} \nonumber \\
& \approx & \frac {j}{2} \label{equ:4-4}
\end{eqnarray}
\MyExplanation{\frac{n}{j}}{\frac{n}{j}}{equ:4-5}
\begin{eqnarray}
\frac{n-\frac{n}{j}}{\frac{n}{j}-1} & \approx & \frac{n-\frac{n}{j}} {\frac{n}{j}} \nonumber \\
& \approx & \frac{j(n-\frac{n}{j})}{n} \nonumber \\
& \approx & \frac{jn(1-\frac{1}{j})}{n} \nonumber \\
& \approx & j(1-\frac{1}{j}) \nonumber \\
& \approx & j - 1 \label{equ:4-5}
\end{eqnarray}
\MyExplanation{\frac{n}{j}}{n-1}{equ:4-6}
\begin{eqnarray}
\frac{n-(n-1)}{\frac{n}{j}-1} & = & \frac {1}{\frac{n}{j}} \nonumber \\
& = & \frac{j}{n} \label{equ:4-6}
\end{eqnarray}
\MyExplanation{n-1}{1}{equ:5-3}
\begin{eqnarray}
\frac{n-1}{(n-1)-1} & = & \frac{n-1}{n-2} \nonumber \\
&\approx & \frac{n-1}{n-1} \nonumber \\
&\approx & 1 \label{equ:5-3}
\end{eqnarray}
\MyExplanation{n-1}{\frac{n}{2}}{equ:5-4}
\begin{eqnarray}
\frac{n - \frac{n}{2}}{(n-1)-1} & = & \frac{\frac{n}{2}}{n-2} \nonumber \\
& \approx &\frac{\frac{n}{2}}{n} \nonumber \\
& \approx & \frac{1}{2} \label{equ:5-4}
\end{eqnarray}
\MyExplanation{n-1}{\frac{n}{j}}{equ:5-5}
\begin{eqnarray}
\frac{n-\frac{n}{j}}{(n-1)-1} & = & \frac {n-\frac{n}{j}}{n-2} \nonumber \\
& \approx & \frac{n(1 - \frac{1}{j})}{n} \nonumber \\
& \approx & 1 - \frac {1}{j} \label{equ:5-5}
\end{eqnarray}
\MyExplanation{n-1}{n-1}{equ:5-6}
\begin{eqnarray}
\frac{n-(n-1)}{(n-1)-1} & = & \frac{1}{n-2} \nonumber \\
& \approx & \frac{1}{n} \label{equ:5-6}
\end{eqnarray}
\MyExplanation{n}{\frac{n}{2}}{equ:6-4}
\begin{eqnarray}
\frac{n - \frac{n}{2}}{n-1} & \approx & \frac {n(1 - \frac{1}{2})}{n} \nonumber \\
& \approx & 1 - \frac {1}{2} \label{equ:6-4}
\end{eqnarray}
\MyExplanation{n}{\frac{n}{j}}{equ:6-5}
\begin{eqnarray}
\frac{n-\frac{n}{j}}{n-1} & \approx & \frac {n(1 - \frac{1}{j})}{n} \nonumber \\
& \approx & 1 - \frac {1}{j} \label{equ:6-5}
\end{eqnarray}
\MyExplanation{n}{n-1}{equ:6-6}
\begin{eqnarray}
\frac{n-(n-1)}{n-1} & = & \frac {1}{n-1} \nonumber \\
& \approx & \frac {1}{n} \label{equ:6-6}
\end{eqnarray}
\section{Introduction}\label{sec:introduction}
The likelihood of a graph, or network to remain functional in the face of random failures and
directed attacks has been the interest to many different authors. In attempting
to understand the problem and their root causes; we reviewed
\cite{Albert2000, crucitti2004error, goh:classification_of_scale-free_networks,
dekker2004simulating, holme2002attack, wang2003cns,criado2005effective,netotea2006evolution,beygelzimer2005improving,zio-modeling,lee2006rnt,newth2004evolving,lee-attack,cohen2001breakdown}.
Our desire is to have a single value that could be used across graphs as an indicator of the graph's
``damage,'' ``robustness,'' or ``general health.'' This value would be applicable whether or not
the graph was \emph{connected} or \emph{disconnected}.
The paper documents the approach used to arrive at a single metric that can be used to
report the damage caused to the graph by the removal of either an
edge or a vertex. The complement of this damage estimate would be the ``health'' of
a graph by the addition of an edge.
Included are
the supporting equations, the data
used to test main stream and ``corner'' test cases and an analysis of the results.
Our sense is that a graph may still perform most of its duties (i.e., communicate
between nodes, maintain data in a node, respond to queries, etc.) even when it
may not be able to perform those functions between any arbitrary nodes $u$ and $v$.
In this sense, a graph may be \emph{connected} or \emph{disconnected}.
There are a number of metrics that can be used to quantify different aspects of a graph
(see sections \ref{sec:connected} and \ref{sec:disconnected}).
Most of these metrics do not have a meaning when the graph
is disconnected. But still; our intuition says that a disconnected graph may still be able
to perform most of its functions. Our intuition is captured in Figure \ref{fig:notionalGraphs}
where representative connected and disconnected graphs are presented. If an attacker
intent on disrupting the functions of a graph, then it is probably reasonable to assume that
the attacker would not be content with simply the disconnection of the graph. The attacker
would probably want to cause greater damage.
\begin{figure}
\centering
\subfigure[Representative connected graph]{\label{fig:notationalConnected} \includegraphics[width=2.0in, angle=-90]{Images/notionalOriginal}}
\subfigure[Representative disconnected graph]{\label{fig:notationalDeleted} \includegraphics[width=2.0in, angle=-90]{Images/notionalDeleted}}
\MyCaption{Representative graphs.}
{The graph in Figure \ref{fig:notationalConnected} is connected, while the one in Figure \ref{fig:notationalDeleted} is not.
Yet, our intuition is that both graphs can perform much of their functions and meet
most of their responsibilities even though they may not be able to meet all of them. }
{fig:notionalGraphs}
\end{figure}
Our quest is to derive a metric for a graph (that is either connected or disconnected)
that reports the inability of a graph to perform its functions. The inverse of this metric
would report its ability to perform its functions, thusly how healthy the graph is. We intend for
this metric to form the basis for a ``game'' where an attacker selects a graph component (either an
edge or a vertex) for removal based on the amount of damage that the removal will cause to
the graph. Additionally, the graph will be able to ``repair'' itself through the addition
of new edges in between selected nodes that would result in a ``less damaged'' graph. The metric
could be used as part of a ``game'' where an attacker and the graph alternated turns. The attacker
could be given the equivalent of a some number of ``bullets'' to damage a graph and then
the graph would be given the same number of repair opportunities to ``repair'' itself.
Section \ref{sec:relwork} presents related work by Albert, Jeong and Barab{\'a}si, and Criado, Flores, et al., and
Holme and Kim among others. A brief synopsis of relevant papers from these authors is given and how
the metric that we intuit exists is different than that put forth by the authors.
Section \ref{sec:alt} presents the criteria that our metric must exhibit.
Section \ref{sec:comparison} investigates how a collection of different metrics performs against
a series of sample graphs. These proposed metrics are use against a small sample graph to
triage candidate metrics. Once our metric is identified, we introduce a series of larger graphs
and continue our investigation. Section \ref{sec:analysis} provides a summary analysis of
Albert, Jeong and Barab{\'a}si's paper. Section \ref{sec:conclusion} contains our conclusion.
Appendix \ref{sec:metrics} contains
a comparison of various graph related metrics applicable to both connected and disconnected
graphs.
Appendix \ref{sec:derivation} contains a more detailed analysis of
Albert, Jeong and Barab{\'a}si's paper.
Appendix \ref{sec:attackProfiles} has a series of profiles that an attacker could
use when seeking to damage a graph. These profiles contain techniques that can focus on either
edges or vertices and then summarizes which profile is most effective.
\section{Related work}\label{sec:relwork}
Albert, Jeong and Barab{\'a}si's (AJB) paper \cite{Albert2000} looks at the effect on the average (or expected)
path length for a graph (specifically snapshots of the Internet and the WWW) when the highest degreed
node (be it an Internet router, or a well connected HTML page) is removed from the graph.
Within their context, the Internet is a graph where routers equate to nodes and communications links
equate to edges. Also the WWW is a graph where pages equate to nodes and HTML links equate to
edges.
They
proposed a tuple metric $(LCC, S, s)$ based on the proportion of the graph represented by the ratio of
largest connected component
$LCC$ to the entire graph \MyLargeS
and the mean size of all remaining fragments \MyLittleS.
Klau and Weiskircher \cite{klau2005robustness} formalized AJB's idea
into a two argument tuple $(S, s)$.
Holme and Kim et al. \cite{holme2002attack} took AJB's paper and
expanded it by introducing the idea of using the \emph{average inverse path length} (AIPL) as an
approach to measure the vulnerability of a graph to different types of attacks.
Crucitti, Latora, et al. \cite{crucitti2004error} published a paper
with the same title as AJB's, dealing with the same general topic, but
proposing a metric they called \emph{global efficiency}. Their global efficiency is
AIPL, but with a different name.
Notetea and Pongor
\cite{netotea2006evolution} proposed measuring the ``robustness'' of a network by computing the
AIPL before and after a change is made to a graph under consideration. If the robustness
of the graph is improved, then the change becomes permanent. If the robustness decreases then
the change is reverted.
Criado, Flores et al. in \cite{criado2005effective} propose to quantify the vulnerability of a graph
based on the number of nodes, number of edges and the standard deviation of the degrees of the
nodes. Ideas from these and other authors are expanded upon in the following sections.
\subsection{Ideas from Albert, Jeong and Barab{\'{a}}si}
Equations \ref{equ:n} through \ref{equ:s} were derived from Albert, Jeong and Barab{\'a}si \cite{Albert2000},
and are the basic definitions for the number of nodes $n$ in the graph at any point in time.
At that point in time, there is a set of clusters \MyLittleS in the graph. If the graph is connected then
there is one cluster.
In \cite{Albert2000}, the node with the highest degree is removed (along with
its adjacent edges) and all values are computed again.
$n$ starts at an initial value and is decremented at each time step until all nodes are disconnected.
Equation \ref{equ:m} is the number of clusters (components) in the set of clusters $c$.
Equation \ref{equ:LCC} identifies the size of the largest connected component $LCC$ in $c$.
Equation \ref{equ:S} is the ratio (percentage) of the size of $LCC$ to the current $n$.
Equation \ref{equ:s} is the mean size of all the remaining clusters (i.e., less the $LCC$) in the graph.
The minimal values of \MyLittleS under differing
conditions (\MyLittleS$=f(n, LCC, m)$) are shown in Table \ref{tbl:s2}.
\begin{equation}
n \stackrel{\mbox{\tiny{def}}}{=} \mbox{ number of nodes in } G
\label{equ:n}
\end{equation}
\begin{equation}
c \stackrel{\mbox{\tiny{def}}}{=} \mbox{ set of clusters in } G
\label{equ:c}
\end{equation}
\begin{equation}
m \stackrel{}{=} \mathclose \mid c \mathclose \mid
\label{equ:m}
\end{equation}
\begin{equation}
LCC \stackrel{}{=} max(\mid \mathclose{<}c \mathclose{>} \mathclose \mid)
\label{equ:LCC}
\end{equation}
\begin{equation}
S \stackrel{}{=} \frac{\ensuremath {\mathclose \mid LCC \mathclose \mid}~}{n}
\label{equ:S}
\end{equation}
\begin{equation}
\ensuremath{ \MyLittleS \stackrel{}{=} \frac{n- \ensuremath {\mathclose \mid LCC \mathclose \mid}~}{m-1}}~
\label{equ:s}
\end{equation}
The various characteristics in equations \ref{equ:n} through \ref{equ:s}
are subject to some mathematical constraints. These constraints are:
\begin{equation}
1 \leq \ensuremath {\mathclose \mid LCC \mathclose \mid}~ \leq n
\label{equ:lcc}
\end{equation}
\begin{equation}
m_{min}= \left \{ \begin{array}{ll}
1 & \mbox{when } \ensuremath {\mathclose \mid LCC \mathclose \mid}~ == n \\
2 & \mbox{otherwise}
\end{array}
\right.
\end{equation}
\begin{equation}
m_{max}= \left \{ \begin{array}{ll}
1 & \mbox{when } \ensuremath {\mathclose \mid LCC \mathclose \mid}~ == n \\
n-LCC & \mbox{otherwise}
\end{array}
\right.
\end{equation}
\begin{equation}
m_{min} \leq m \leq m_{max}
\label{equ:M}
\end{equation}
\begin{equation}
1 \leq j \leq m
\label{equ:j}
\end{equation}
In addition to the mathematical constraints, there are a series of logical constraints. These constraints are:
\begin{enumerate}
\item \MyLittleS$ < \ensuremath {\mathclose \mid LCC \mathclose \mid}~$\label{cont:5} \MyEquationReference{equ:s}
\item \MyLargeS will always be in the range $\frac {1}{n} \leq S \leq 1$ \label{cont:6} \MyEquationReference{equ:S}
\item If \ensuremath {\mathclose \mid LCC \mathclose \mid}~ == 1 then $\forall c: \mid c_{i} = 1 \Longrightarrow m = n$ meaning that anytime where $m == n$
and \ensuremath {\mathclose \mid LCC \mathclose \mid}~ $\neq 1$~is a contradiction and can not happen.\label{cont:1}
\item If $\ensuremath {\mathclose \mid LCC \mathclose \mid}~ == \frac{n}{2} \Longrightarrow m_{max} = \frac {n}{2} $ where $ \forall c_{i}:\mathclose \mid c_{i} \mathclose \mid == 1$.\label{cont:2}
\item If $\ensuremath {\mathclose \mid LCC \mathclose \mid}~ == \frac{n}{j} \Longrightarrow m_{max} = \frac {n}{j} $ where $ \forall c_{i}:\mathclose \mid c_{i} \mathclose \mid == 1$.\label{cont:4}
\item If $\ensuremath {\mathclose \mid LCC \mathclose \mid}~ == (n-1) \Longrightarrow m = 2$.\label{cont:3}
\end{enumerate}
Constraint \ref{cont:6} limits \ensuremath {\mathclose \mid LCC \mathclose \mid}~ between $n$ and 1. The \ensuremath {\mathclose \mid LCC \mathclose \mid}~~ will equal $n$
when the graph is connected (i.e., the graph has not been fragmented).
$LCC$ will equal 1 when the graph is totally disconnected (i.e., the graph is composed of only nodes and no edges). Equation \ref{equ:M} limits
the number of fragments $m$ to
between 1 and $n$.
Equation \ref{equ:j} limits the number of fragments to the greater of 1 (when the graph is totally connected;
i.e. one cluster) or $n$ (when the graph is totally disconnected).
AJB were interested in the fraction $f$ of their graphs that had to be removed to cross a
percolation threshold that would cause the graph to become severely fragmented. We are interested
in the continuum of the graph's performance while it is connected and after it
is disconnected. The percolation threshold is of passing interest, while the ideas that they espouse
serve as starting point for our investigation.
\subsection{Ideas from Criado, Flores, et al.}
Criado, Flores et al. in \cite{criado2005effective} propose to quantify the vulnerability of a graph
based on the number of nodes, number of edges and the standard deviation of the degrees of the
nodes. Perhaps most importantly, they define the attributes of a vulnerability function
in terms of the graph.
Their definition is:
\begin{itemize}
\item[]Let $\mathcal{G}$ be the set of all possible graphs with a finite number of vertices. A
\emph{vulnerability function} $v$ is a function $v:\mathcal{G}\rightarrow [0,1]$~ verifying the following
properties:
\begin{enumerate}
\item $v$ is invariant under isomorphisms.
\item $v(G') \leq v(G)$ if $G'$ is obtained from $G$ by adding edges.
\item $v(G)$ is computable in polynomial time with respect to the number
of vertices of $G$.
\end{enumerate}
\end{itemize}
The equation they present to meet their definitions is:
\MyEquationLabeled{vulnerability}{vulnerability}
Supported by:
\MyEquation{vulnerabilityDegreeStdDev}
Equation \ref{vulnerability} evaluates to the interval [0,1]. A value of
0 means that the graph is very robust (low vulnerability), while a value of 1
means that the graph is very vulnerable (not robust). Using equation \ref{vulnerability}
before and after a modification to a graph can be used as a way to measure
what effect the change has had on the graph's vulnerability. If the
vulnerability increases, then probably the change should not be finalized.
While their system of equations meets their requirements, the equations do not report
the type of damage that we are interested in measuring. Their definition of the attributes
of a metric are in harmony with our intuition.
\subsection{Ideas from Holme and Kim}
Holme and Kim in \cite{holme2002attack} looked at how an attacker could maximize the damage
to a graph by following one of two approaches:
\begin{enumerate}
\item To remove the vertex with the highest initial degree (ID)
\MyEquation{degreeUndirectedDegree}
\item Or, the vertex with the highest normalized in-betweenness
centrality (IB) \MyEquation{betweennessVertex}
\end{enumerate}
For these approaches, they allowed the attacker two different options. The options are:
\begin{enumerate}
\item To attack the graph (remove a vertex) based on the ordering of the vertices when a series of attacks started, or
\item To recompute ID and IB after a vertex has been removed.
\end{enumerate}
This second option took into account that the characteristics of the graph change when a vertex
is removed and therefore the ID and IB ordering would change. Recomputed ID and IB were called
RD and RB respectively.
They used their ID, IB, RD and RB attack profiles on the hep-lat e-print archive,
a snapshot of the Internet autonomous system connections over a 24 hour period and
Erd{\"{o}s-R{\'{e}nyi random, Watts-Strogatz small-world, and
Barab{\'{a}si-Albert scale-free graphs. They concluded that each of the different types of
graphs respond (as in how the AIPL responds) differently and that the attacker
should use the RB approach to maximize the impact as measured by AIPL.
Holme and Kim used AIPL as their metric to assess the functionality of the current
graph. They did not use AIPL to assess how the most recent
attack affected the graph's ability to perform.
\subsection{Ideas from Crucitti, Latora, Marchiori and Rapisarda}
Crucitti et al. in \cite{crucitti2004error} look at the behavior of a network (i.e., a graph
that has a measurable flow along an edge) when a node or an edge is removed. Their
premise is that the flow between nodes will always take the lowest cost path. In their
models, each edge has a capacity and a tolerance factor. As edges/nodes are removed, the
flow that was going through the removed component is spread out to other edges. The
removal of a critical edge (high flow) and the redistribution of the flow through
adjacent edges can result in a cascade of failures as the increased flow causes additional
edges to reach saturation.
They investigated these phenomena for Erd{\"{o}s-R{\'{e}nyi random graphs and
Barab{\'{a}si-Albert scale-free graphs using the same ideas of ID, IB, RD and RB as introduced
in by Holme in \cite{holme2002attack}. Crucitti introduces the idea of
\emph{global efficiency} that has the same form and character as AIPL.
\MyEquation{globalEfficiency2}
Crucitti computes global
efficiency after a node or an edge is removed, but they do not compare the current
efficiency versus a connected graph's efficiency.
\subsection{Ideas from Netotea and Pongor}
Netotea and Pongor in \cite{netotea2006evolution} focus on the evolution of a graph towards
a new organization that is more robust or efficient. Their definition of efficiency $E$ is
AIPL and their definition of robustness $R$ is the ratio of the current efficiency $E_{t}$
divided by the
the previous efficiency $R = \frac{E_{t}}{E}$.
Netotea and Pongor use a genetic algorithm that starts with a random graph (100 nodes and 120 edges)
and mutates and crossovers the graph until it reaches a ``steady state'' condition. A steady
state was achieved when the goals of $E$, $R$ and the maximum percentage of periphery nodes (those
nodes with a degreeness of 1) was reached. $E_{t}$ was computed after either 1 or 5 of the
highest betweenness nodes were removed.
Netotea and Pongor's idea of \emph{robustness} $R$ comes close to capturing our idea of a
single number that measures the \emph{health} of a graph. Health is the inverse
of our idea of damage.
\subsection{Ideas from others}
Lee and Kim in \cite{lee-attack} look at the effects of node and path failure on the Internet and
report on the percentage of nodes that are required for disconnection. While they model failure of
the graph, they do not report on how damaged the graph is when attempting to perform its functions.
Cohen et al. in \cite{cohen2001breakdown} focus on the modeling the failure of the Internet
when the most connected routers (highest degreed nodes) are removed. While they look towards
quantifying the percolation value $p$ where the Internet and scale-free graphs become disconnected,
they do not report on the graph's ability to perform.
Newth and Ash in \cite{newth2004evolving} look at
cascading failures in a complex network. They extend the work of Crucitti et al. in \cite{crucitti2004error}
by manipulating their graph by: (1) adding a new edge, or (2) deleting an existing edge, or (3)
changing one end of an existing edge. If the graph becomes disconnected during any of these
operations, the change in rejected.
Beygelzimer et al. in \cite{beygelzimer2005improving} use AIPL
as their metric for the robustness of a graph. They take an existing graph,
rewire it using a number of different schemes and look at the robustness after each modification.
They disallow any rewiring that would disconnect the graph.
Zio and Sansavini in \cite{zio-modeling} look at how the failure of a node or an edge may
cause a failure in adjacent components as the load of the failed component cascades to its neighbors.
These failures may be the result of random acts or targeted attacks. They do not use
transfer of load as a metric of the damage done to the graph.
Lee et al. in \cite{lee2006rnt} look at how the topology of the graph affects which type of attack
profile would be most effective. They propose a new metric, called \emph{attack power} to
quantify the effect of any of their attack profiles. They measure damage to their graph
using degree distribution, average path length and vertex cover. They enumerate some interesting
attack profiles, but their approach does not address a disconnected graph.
Klau and Ren{\'e} Weiskircher
in \cite{klau2005robustness} (a chapter in \cite{brandes2005network})
provide a very nice survey of robustness and resilience metrics
and ideas that have been advocated by various authors. None of the approaches provide
a single unit-less value that describes the damage inflicted on a graph by the removal
of an edge or node and the possible disconnection of the graph.
Dekker in \cite{dekker2004simulating} introduces the idea of \emph{intelligence} of a graph related
to the quality of a sensor and the time delay associated with the data from the sensor. The intelligence
of the graph starts to loose its meaning when the graph becomes disconnected. While the idea of
intelligence in the graph is appealing to our sense that a graph can still perform when it is
fragmented, Dekker's metric does not speak to the total graph.
{\'A}goston et al. in \cite{agoston2005multiple} enumerate a series of attack profiles including:
\begin{enumerate}
\item \textit{Complete knockout} --- meaning the removal of a node and its adjacent edges,
\item \textit{Partial knockout} --- meaning the removal of a set of edges (but not all) adjacent to a node,
\item \textit{Attenuation} --- meaning that the amount of traffic that an edge can support is decreased,
subsequently, the total cost of a path that uses that edge from a source node \MyLittleS to a terminus node $t$ is increased,
\item \textit{Distributed knockout} --- meaning that a set of edges, not sharing a common node, are removed,
\item \textit{Distributed attenuation} --- meaning that the amount of traffic that the set of edges can support
is decreased.
\end{enumerate}
These attack profiles are used in simulated attacks on \textit{Escherichia coli} and
\textit{Saccharomyces cerevisiae} transcriptional regulatory networks. Their conclusion is that
multiple partial attacks causes more damage. Our interests are slightly different because edges
in our network of DOs are really communications links vice edges that have a measurable
capacity. DOs in our network can either send messages via these communications links or they can not.
This difference in edge utilization and modeling eliminates the attenuation and distributed attenuation
profiles. In our network, a DO exists or it does not and therefore all of its adjacent edges (communications
links) are valid, or not. This approach matches {\'A}goston's complete knockout profile.
We view partial and distributed knockouts as being repeated application of removing
single edges in our network.
Yin et al. in \cite{yan2008multiple} take the ideas from {\'A}goston in \cite{agoston2005multiple} and apply
them to scale-free and random graphs. Yin et al. apply weights to the edges in their graphs and use
AIPL as a metric to quantify the effect of each attack profile. Their results confirm that scale-free
networks are relatively immune to random attacks, but very sensitive to targeted attacks. While both
random and targeted attacks on random graphs have relatively the same effect.
Lee et al. in \cite{lee2006rnt} use the autonomic system (AS) connectivity graphs from National Laboratory for Applied
Network Research as their test graph. Based on this graph, they apply weights to each of the
edges in the graph based on the amount of traffic along that edge. They then focus on three different
types of failures. \textit{Node failure} where an AS is lost due to some sort of hardware failure (i.e.,
power supply failure, accidental or deliberate misconfiguration, etc.). \textit{Link failure} where
adjacent ASes are not able to communicate because of hardware failure (such as the cutting of a cable), or
electronic failure (such as DNS hacking, routing table poisoning, etc.). \textit{Path failure} including
DoS and routing table loops, resulting in a flooding of the path with packets to the extent that the
communications links are unusable. Lee et al. then create different attack profiles based on
these types of failures. Their attack profiles are:
\begin{enumerate}
\item \textit{Random AS attack} --- randomly choose an AS and and remove it,
\item \textit{Min-degree AS attack} --- order the ASes by their degree connectivity and then start removing
them from low degree to high degree order,
\item \textit{Max-degree AS attack} --- order the ASes by the degree connectivity and then start removing
them from high degree to low degree order,
\item \textit{Random edge attack} --- randomly choose an edge and remove it,
\item \textit{Min-weight edge attack} --- order the edges by their weight and then start removing
them from low weight to high weight order,
\item \textit{Max-weight edge attack} --- order the edges by their weight and then start removing
them from high weight to low weight order,
\item \textit{Random path attack} --- randomly choose a path and remove it,
\item \textit{Max-weight edge attack} --- order all paths by weight and then remove paths in order
from heaviest to lightest, and
\item \textit{Max-length path attack} --- order all paths by length and then remove paths in order
from longest to shortest.
\end{enumerate}
After each attack, the effect on the graph is quantified by a metric they labeled as ``attack power''
Attack power reports the effect of each attack on the number of components that fail in the system. We treat
Lee's path failure as a limited case of our edge failure \MySectionReference{sec:edgeSelection}. Path failure is based on the path at the start of the
attack where the path meets some sort of criteria and then a series of edges are removed based on these criteria.
The limitation is that the set of criteria used to identify the path in the first place, may not be valid after
the removal of the first edge in the path. We select an edge based on some criteria, remove the edge and then
reevaluate the entire graph to select the next edge. We do not base future actions on information that
may be stale or obsolete.
Latora et al. in \cite{latora2005vulnerability} look at the vulnerability of complex networks to three different attack
profiles and then provide a method to reduce the vulnerability of the network by the addition
of edges between selected nodes. Their attack profiles are: \textit{loss of a single cable connection} (loss of an
edge), \textit{loss of a single Internet router} (loss of a single node) and \textit{loss of two Internet routers}
(loss of two nodes). They assume that for the system $S$ there exists a \emph{performance} metric
$\MyEquationInline{latoraPerformanceLHS}>0$ that characterizes
the performance of the graph and that this metric increases in value when the graph is damaged $D$. Therefore
\MyEquationLabeled{latoraVulnerability}{equ:latora:vuln}
Where \MyEquationInline{latoraDamageWorst} is the worst possible damage that can happen to the graph
based on a specific attack profile. They use \MyEquationInline{latoraVulnerabilityLHS} as a metric
to quantify the efficacy of an attack.
The same metric is used to evaluate the effect of adding a communications link (an edge) between any two nodes
in order to improve (i.e., reduce the \emph{vulnerability}) of the system.
Our approach is different in that we we are explicit about the metric that we will use
to measure the ``performance'' of the graph and we are currently focusing on attacking
the graph vice repairing it. Our approach could be used to evaluate graph repair alternatives.
\section{An alternative approach}\label{sec:alt}
After looking at the different approaches in Section \ref{sec:relwork} and thinking about what it is that the \emph{damage} metric
is trying to capture, we do not feel that individually any of them fit the bill.
The attributes of the \emph{damage} metric should be:
\begin{enumerate}
\item Different fragmentation cases should result in different numerical value,
\item Test cases where the size of the fragments have been scaled,
and the entire graph (for instance, increased by a factor of 10 or 0.1) should result in the same value
\item The value should be useful without additional information about the graph (i.e., the
metric is graph independent and does not require knowledge of the graph in a
different state),
\item The metric should be \emph{unitless}. The approach and equations from
AJB's paper \cite{Albert2000} have some function of
$node$. The units of \MyLargeS or \ensuremath {\mathclose \mid LCC \mathclose \mid}~ and \MyLittleS is $nodes$. \MyEquationInline{fScoreLHS} \MyEquationReference{equ:fscore} and the generalized \MyEquationInline{fScoreBetaLHS}
\MyEquationReference{equ:fscorebeta} metrics have units of $nodes$. Geometric \MyEquationReference{equ:geomean} and quadratic
mean \MyEquationReference{equ:quadmean} and ratio (\MyLittleS$/$\MyLargeS) are unit-less and therefore attractive.
\end{enumerate}
The desire/need to have a unit-less and scale-free description of the fragmentation and damage of a graph points to using a different
type of metric. One that appears popular is based on the average inverse average path length (AIPL)
\MyEquationReference{equ:02c}. There are a couple of variations on Equation \ref{equ:02c},
such as Equation \ref{equ:avi2} from \cite{netotea2006evolution} and Equation \ref{equ:avi3} from
\cite{crucitti2004error}. Equation \ref{equ:avi2} is applicable to a graph that has directed edges
and permits self loops. Equation \ref{equ:avi3} is applicable to a graph that has directed edges
and does not permit self loops.
\MyEquationLabeled{globalEfficiency}{equ:avi2}
\MyEquationLabeled{globalEfficiency2}{equ:avi3}
AIPL equations are used to compute the AIPL between any
pair of nodes in a graph, even if the graph is disconnected. Use of the AIPL can be counter intuitive, in that a larger AIPL is
better than a smaller AIPL because a smaller AIPL means that the average path length is increasing.
At the core of the \MyEquationInline{damageLHS} metric is the ratio of two AIPLs. One
of the damaged/fragmented graph and the other an unfragmented artificial graph.
\MyEquationLabeled{damage-02}{equ:damageLocal}
The unfragmented artificial graph is constructed by sorting the original graph fragments
by their size and repeatedly connecting the nodes of two largest fragments with the highest centrality
value \MyEquationReference{equ:04c} until the graph is connected.
Conceptually, the artificial graph could have been existed in the fragmented graph's past and the
current fragmented graph is the result from losing edges. The edges could have been lost due
to error or attack.
\section{Comparison and evaluation of various metrics}\label{sec:comparison}
\subsection{Small test case}
We create a small graph with 21 nodes and 27 edges \MyFigureReference{fig:original},
and use it to show the effects
on ``classical'' graph metrics by using different attack profiles.
Damage will be inflected on the
graph by targeting either
the edge (\MyAttackNotation{E}{*}) or the vertex (\MyAttackNotation{V}{*}) based on its betweenness centrality measurement.
The betweenness centrality measurement is a count (or normalized value)
of the number
of geodesic paths that use either an edge or a vertex, hence
edge or vertex centrality to the graph.
For \MyAttackNotation{E}{*} or \MyAttackNotation{V}{*}, the appropriate centrality measurement
is computed and the component (edge or vertex as applicable) is removed from the graph.
Various graph metrics are computed and reported after each removal.
This targeted attack is repeated until the graph becomes disconnected.
After
targeting the edges, the graph will be restored to its initial condition
prior to targeting the vertices.
Removing graph components (either an edge or a vertex) may result in the
graph becoming disconnected, or fragmented. Sometimes this fragmentation
will result in a graph that is divided in half and whose $LCC$ is approximately
the same size as the non $LCC$. A different choice in which component to remove
(a different attack criteria), might result in a graph whose $LCC$ contains
all the remaining edges and all but one node.
\begin{figure}
\centering
\includegraphics*[viewport = 180mm 220mm 20mm 55mm, clip, width = 5.0in, angle = 90]{Images/original.eps}
\MyCaption{Small test graph used to show the effects of different attack profiles.}
{The graph has 21 nodes and 27 edges. It clearly shows 2 groupings of nodes that are
connected by 2 separate sets of edges.}
{fig:original}
\end{figure}
\newcommand{\MyToyTabular}[0]{\begin{tabular}{|*{8}{p{0.50in}|}}}
\newcommand{\MyToyHeader}[1] {&APL&AIPL&{#1}& Clustering coefficient & Diameter & Eccentricity & Radius\\}
\subsubsection{Removal of vertices}
\input{removalOfVertex}
\subsubsection{Removal of edges}
\input{removalOfEdge}
\subsubsection{Comparing \MyAttackNotation{E}{H} and \MyAttackNotation{V}{H} profiles}
Both the \MyAttackNotation{E}{H} and the \MyAttackNotation{V}{H} profiles result
in a disconnected graph after to removals. But the two profiles result in different
graphs at time point of disconnection (compare Figure \ref{fig:vertex:4} and Figure \ref{fig:edge:3}).
\subsection{A change in notation}
Figure \ref{fig:original} is small and sparse enough that it is practical to draw
and label the complete graph and still be able to understand its structure. As graphs
get larger, and more interesting it is not practical to draw and label every component.
Therefore, we introduce a different notation style that is more in keeping with the
aspects of the graph that are if interest to our research.
We are interested in how the graph functions, its connectivity as it becomes more
and more fragmented. The internal connectivity (how many edges are in a fragment)
is of less interest than the fact that the graph is fragmented, and that the numbers
and relative sizes of these fragments can be used as a metric to describe how well
the fragmented graph ``operates'' when compared to the unfragmented graph.
Specific graph instances will have names such as \MyShortNameB whose \ensuremath {\mathclose \mid LCC \mathclose \mid}~ and number and
size of any fragments are shown in Table \ref{tbl:cases}. Tables \ref{tbl:circles.1} through \ref{tbl:circles.4}
provide notational diagrams of the graph instances.
\subsection{Larger test cases}
A series of test cases were constructed to exercise the different approaches
proposed by Albert, Jeong and Barab{\'{a}}si and ourselves.
Each test case consists of some
number of fragments (a.k.a., components) between 1 and 11. The test cases are intuitively ordered from least to most
damaged. The test cases are described numerically in Table \ref{tbl:cases}, and shown
diagrammatically in Tables \ref{tbl:circles.1} through \ref{tbl:circles.4}.
\begin{table}
\input{fragsDescriptionNew}
\end{table}
\input{comparisonTable}
\input{graphsNew}
\subsection{Comparison equations}
Now that we have the basic definitions and constraints out of the way, we can begin to look at how AJB's \MyLargeS and \MyLittleS will
be evaluated. A set of equations was selected that seemed like they might be of use. The set includes:
\begin{enumerate}
\item The median value of all the fragments, except the $LCC$.
\item The average size of all the fragments, except the $LCC$.
\item The standard deviation of all the fragments, except the $LCC$.
\item The harmonic mean of all the fragments, except the $LCC$.
\item The geometric mean of all the fragments, except the $LCC$.
\item A variation on the information retrieval (IR) metric \MyEquationInline{fScoreLHS}
\MyEquationReference{equ:fscore} (a 2 value harmonic mean). We selected \MyEquationInline{fScoreLHS}
because it had been used in other applications and we thought that it might be useful.
In the IR world, \MyEquationInline{fScoreLHS} traditionally
operates on the values of \textit{precision} and \textit{recall}.
For the purposes of analysis \MyLargeS was treated as \textit{precision} and \MyLittleS was
treated as \textit{recall}.
\MyEquationLabeled{fScore}{equ:fscore}
\item A generalized \MyEquationInline{fScoreBetaLHS} \MyEquationReference{equ:fscorebeta} metric that incorporates a value $\beta$
that is used to weight \textit{precision} relative to
\textit{recall}.
\MyEquationLabeled{fScoreBeta}{equ:fscorebeta}
\item A simple arithmetic mean of \MyLargeS and \MyLittleS.
\item A geometric mean of \MyLargeS and \MyLittleS \MyEquationReference{equ:geomean}.
\MyEquationLabeled{geometricMean}{equ:geomean}
\item A quadratic mean of \MyLargeS and \MyLittleS \MyEquationReference{equ:quadmean}.
\MyEquationLabeled{quadraticMean}{equ:quadmean}
\item Ratio of \MyLittleS to \MyLargeS.
\item \MyLargeS raised to the \MyLittleS power.
\item \MyLittleS raised to the \MyLargeS power.
\end{enumerate}
In equations \ref{equ:fscore} through \ref{equ:quadmean}, $x_{1} = S$ and $x_{2}=s$.
\section{Analysis}\label{sec:analysis}
The interaction between \MyLargeS, $LLC$ and \MyLittleS is of interest and is summarized in Table \ref{tbl:s2}. Various cells in Table \ref{tbl:s2} are have different colors and color is significant.
Cells that are filled with \color{cyan} cyan \color{black}~ violate some basic mathematical operation.
Cells that are filled with \color{orange} orange \color{black}~ violate some some logical restriction on $LCC$ \MyConstraintReference{equ:lcc}.
Cells that are filled with \color{red} red \color{black}~ violate some logical restriction on $m$ \MyConstraintReference{equ:m}. There may be cases when a combination of $j$, $m$, \MyLargeS, $LLC$ and \MyLittleS
violates more than one constraint, in those cases the fill color will be chosen at random. Cells that are not filled, do not violate any constraints.
There is a limited range of values for $m$ and \ensuremath {\mathclose \mid LCC \mathclose \mid}~~ that do not violate some sort of mathematical or
logical constraint when attempting to compute \MyLittleS. These limits are in keeping with the values
computed in AJB's paper.
The test cases from Table \ref {tbl:cases} were subjected to a series of mathematical investigations looking to
identify and quantify a metric that was near 0 for ``undamaged'' graphs and near 1 for
``damaged'' ones. Table \ref{tbl:case01} shows the various mean and standard
deviation values for the test cases. These approaches produced values that had
no discernible relationship to their state of damage.
Table \ref{tbl:case02} showed some useful information, but each of these more
sophisticated approaches, had some sort of ``hump'' or ``swale'' in the computed values.
Values produced by using these approaches would initially trend in the right direction
(from low to high) as the case numbers increased, but then the values would change direction
and start to go the other way. Some of the exponentiation cases, created values that were too
large for the computer to handle reasonably. While these computational limitations could
be overcome, there does not seem to be any reason to expend the effort to do so when
the data that was available was not well behaved.
None of the approaches in Tables \ref{tbl:case01} and
\ref{tbl:case02} showed the desired property of continuous directed change.
The investigation into a unit-less metric for assessing the ``damage'' inflicted upon a graph by fragmentation, led to writing an R script
that could produce three different types of graphs; random, small world and scale free. These graph types were selected
because they are felt to represent approximately the extremes of the fundamental graph types.
The R script takes as an a argument the fragments that make up the test case \MyTableReference{tbl:cases}. Two graphs
are created based on the fragments. The first graph is a simple connected graph whose size
is equal to the sum of the fragments. The second graph is a simple disconnected graph whose
size is equal to the sum of the fragments.
The average inverse path length (AIPL) \MyEquationReference{equ:02c} for the two graphs is
computed and then the ratio of the AIPLs is reported. The hoped for behavior (a value near zero when the
graph is not too damaged, and near unity when severely damage) is exhibited by the ratio of the AIPLs \MyTableReference{tbl:case03}.
The ratio of the AIPLs metric for the test cases does range from 1.0 to 0.0
\MyTableReference{tbl:case03} fitting our intuition. Now the question becomes, does that
metric continue (within reasonable bounds) as the size of the graph changes, this
is in keeping with the desirable behavior of the metric as listed in section \ref{sec:alt}.
The base size of the graph was increased by factors of 2, 4, 8 and 10,
the ratio was computed and reported \MyTableReference{tbl:case04}. Data in Table \ref{tbl:case04}
shows the metric starts at 1.0 for a non-fragmented graph and decreases towards 0.00 as the
graph becomes more fragmented and \ensuremath {\mathclose \mid LCC \mathclose \mid}~ becomes smaller. Data in the table for
totally fragmented graphs does not reach 0.000 as the graph becomes larger possibly
because the round offs when computing all the paths and their inverses start to accumulate.
Where the expected value should be 0.000, it is in fact 0.0. Computing all shortest paths
in a graph using the Floyd-Warshall algorithm can take $\Theta(V^{3})$ time \cite{cormenintroduction},
so larger
graphs were not fully analyzed.
\begin{table}
\input{tableS2}
\MyCaption{Analysis of \MyLittleS based on possible values of \ensuremath {\mathclose \mid LCC \mathclose \mid}~~ and $m$.}{\ensuremath{ \MyLittleS \stackrel{}{=} \frac{n- \ensuremath {\mathclose \mid LCC \mathclose \mid}~}{m-1}}~ is the average size of all fragments
in the graph, less the $LCC$. The table summarizes the lower limit on \MyLittleS based on the maximum number of
fragments $m$ there can be in the graph based on \ensuremath {\mathclose \mid LCC \mathclose \mid}~. Where the value in the cell is not obvious (i.e.,
how it was derived, what assumptions were made, etc.), the (E\#) refers to a set of equations
that show how the value was obtained. In some cells there is a constraint logical violation. These constraint
violations are shown as (C\#).}{tbl:s2}
\end{table}
\input{simpleTable}
\input{lessSimple}
\input{bridgeTable}
\input{largerTable}
\section{Conclusion}\label{sec:conclusion}
Considerable time was spent examining the equation Albert, Jeong and Barab{\'a}si \ensuremath{ \MyLittleS \stackrel{}{=} \frac{n- \ensuremath {\mathclose \mid LCC \mathclose \mid}~}{m-1}}~
from \cite{Albert2000} to see how it could be used to quantify the ``damage'' to a
graph when the graph becomes fragmented. This investigation was spurred on by the
equation's use in \cite{Albert2000,brandes2005network} and the
belief that there was more information there that could be of use.
The equation was analyzed and limits (both mathematical and logical)
were identified. These limits fit nicely with the graphs in both references.
Because of the limitations experienced using the tuple (\MyLittleS, $m$, \MyLargeS) from AJB
and the desire to have a unit-less
metric that reflects the efficiency of the graph; a different approach was identified.
Netotea and Pongor in \cite{netotea2006evolution} and Crucitti et al. in
\cite{crucitti2004error} proposed
the use of the average inverse path length (AIPL) as a way of quantifying the efficiency of a graph.
We used the equations from Crucitti to compute the AIPL of
a connected graph that is equal to the sum of all
the fragments and the original disconnected graph consisting of the fragments.
A ratio was computed using these AIPLs. This ratio has the desired effect of being: (1) unit-less,
(2) independent of graph size, and (3) does not require a priori knowledge of
the graph.
The ratio of the average inverse path lengths of a connected and a disconnected graph
can be used as a metric about the health of a graph. The damage (i.e., the converse of health) of a fragmented graph can be
computed using \MyEquationInline{damage-02.tex}.
\input{include_acknowledgement}
\bibliographystyle{abbrv}
\section{Acknowledgment}\label{sec:acknowledgement}
This work supported in part by the NSF, Project 370161.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.